Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-25T05:31:06.329Z Has data issue: false hasContentIssue false

Part I - Challenges: Time and Memory

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. 11 - 142
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

Agar, M. (1973). Ripping and Running: A Formal Ethnography of Urban Heroin Addicts. New York: Seminar.Google Scholar
Agar, M. (2004). We have met the other and we’re all nonlinear: ethnography as a nonlinear dynamic system. Complexity, 10(2), 1624.Google Scholar
Agar, M. (2007). Dope Double Agent: The Naked Emperor on Drugs. Lulu.com.Google Scholar
Anderson, B. (1983). Imagined Communities: Reflections on the Origin and Spread of Nationalism. New York: Verso.Google Scholar
Aristotle. Meteorology. Book 1, Part 14.Google Scholar
Bailey, G. (1981). Concepts, time-scales and explanations in economic prehistory, in Economic Archaeology, Sheridan, A. and Bailey, G., eds., British Archaeological Reports, pp. 97–117 (International Series) 96, Oxford.Google Scholar
Bailey, G. (2007). Time perspectives, palimpsests and the archaeology of time. Journal of Anthropological Archaeology, 26, 198223.Google Scholar
Barnett, H. G. (1953). Innovation: The Basis of Cultural Change. New York: McGraw-Hill.Google Scholar
Barthel, S., Crumley, C., & Svedin, U. (2013a). Bio-cultural refugia: safeguarding diversity of practices for food security and biodiversity. Global Environmental Change, 23(5), 1142–52.Google Scholar
Barthel, S., Crumley, C., & Svedin, U. (2013b). Bio-cultural refugia: combating the erosion of diversity in landscapes of food production. Ecology and Society, 18(4), 71.Google Scholar
Bateson, G. (1972). Steps to an Ecology of Mind. New York: Random House.Google Scholar
Bateson, G. (1979). Mind and Nature – A Necessary Unity. New York: Bantam Books.Google Scholar
Beekman, C. S. & Baden, W. S., eds. (2005). Nonlinear Models for Archaeology and Anthropology. Aldershot (Hampshire), UK: Ashgate Press.Google Scholar
Bengtsson, J., Angelstam, P., Elmqvist, Y., Emanuelsson, U., Folke, C., Ihse, M., Moberg, F., & Nyström, M. (2003). Reserves, resilience and dynamic landscapes. Ambio, 32(6), 389–96.Google Scholar
Berry, W. B. N. (1987). Growth of a Prehistoric Time Scale: Based on Organic Evolution. Palo Alto, CA: Blackwell Scientific Publications.Google Scholar
Binford, S. & Binford, L., eds. (1968). New Perspectives in Archaeology. Chicago: Aldine.Google Scholar
Bintliff, J., ed. (1991). The ‘Annales’ School and Archaeology. New York: New York University Press.Google Scholar
Boettiger, C. & Hastings, A. (2012). Early warning signals and the prosecutor’s fallacy. Proceedings of the Royal Society B, 279, 4734–9.Google ScholarPubMed
Bourgois, P., Lettiere, M., & Quesada, J. (1997). Social misery and the sanctions of substance abuse: confronting HIV risk among homeless heroin addicts in San Francisco. Social Problems, 44, 155–73.Google Scholar
Brand, F. (2005). Ecological Resilience and its Relevance within a Theory of Sustainable Development. UFZ Centre for Environmental Research Leipzig-Halle in the Helmholtz-Association. Department of Ecological Modelling UFZ-Report 0 3/2005 ISSN 0948-9452.Google Scholar
Braudel, F. (1972). The Mediterranean and the Mediterranean World in the Age of Philip II. Originally published in 1949. New York: Harper.Google Scholar
Carlstein, T. (1982). Time Resources, Society, and Ecology: On the Capacity for Human Interaction in Space and Time. London; Boston: Allen & Unwin. ISBN 0043000827. OCLC 7946554.Google Scholar
Carpenter, S. R., Cole, J. J., Pace, M. L., Batt, R., Brock, W. A., Cline, T., Coloso, J., Hodgson, J. R., Kitchell, J. F., Seekell, D. A., Smith, L., & Weidell, B. (2011). Early warnings of regime shifts: a whole-ecosystem experiment. Science, 332, 1079–82.Google Scholar
Chase, A. F. & Scarborough, V. L., eds. (2014). The Resilience and Vulnerability of Ancient Landscapes: Transforming Maya Archaeology through IHOPE. Archaeological Papers of the American Anthropological Association, no. 24. Hoboken, NJ: Wiley.Google Scholar
Clarke, D. L. (1968). Analytical Archaeology. London: Methuen.Google Scholar
Clements, F. E. (1928). Plant Succession and Indicators. New York: H. W. Wilson.Google Scholar
Crumley, C. L. (2005). Remember how to organize: heterarchy across disciplines. In Beekman, C. S. and Baden, W. S., eds., Nonlinear Models for Archaeology and Anthropology, pp. 3550. Aldershot (Hampshire), UK: Ashgate Press.Google Scholar
Cumming, G. S. & Norberg, J. (2008). Scale and complex systems. In Norberg, J. and Cumming, G. S., eds., Complexity Theory for a Sustainable Future, pp. 246–76. New York: Columbia University Press.Google Scholar
Darwin, C. (1859). The Origin of Species by Means of Natural Selection. London: John Murray.Google Scholar
Davies, P. (2014). That mysterious flow. Scientific American: A Matter of Time, 23, 813. Published online: 23 October 2014. doi:10.1038/scientificamericantime1114-8.Google Scholar
DeLanda, M. (1997). A Thousand Years of Nonlinear History. New York: Zone Books, New York.Google Scholar
Diamond, J. (2005). Collapse – How Societies Choose to Fail or Succeed. New York: Viking Penguin.Google Scholar
Ellen, R. (1982). Environment, Subsistence and System: The Ecology of Small-Scale Social Formations. Cambridge: Cambridge University Press.Google Scholar
Evans, J. (1872). The Ancient Stone Implements, Weapons and Ornaments of Great Britain. New York: D. Appleton and Company.Google Scholar
Evans, J. (1881). The Ancient Bronze Implements, Weapons, and Ornaments of Great Britain and Ireland. London: Longmans Green & Co.Google Scholar
Evans-Pritchard, E. E. (1954). Nuer Religion. Oxford: Oxford University Press.Google Scholar
Flannery, K. V. (1972). The cultural evolution of civilizations. Annual Review of Ecology and Systematics, 3, 399426.Google Scholar
Fraser Darling, F. (1969). The Impact of Man on the Biosphere. UK: BBC Reith Lectures.Google Scholar
Garnsey, E. & McGlade, J., eds. (2006). Complexity and Co-evolution, Continuity and Change in Socio-economic Systems. Cheltenham: Edward Elgar.CrossRefGoogle Scholar
Gibbard, P. & van Kolfschoten, T. (2004). The Pleistocene and Holocene epochs. In Gradstein, F. M., Ogg, J. G., & Smith, A. G., eds., A Geologic Time Scale 2004. Cambridge: Cambridge University Press.Google Scholar
Giddens, A. (1979). Central Problems in Social Theory: Action, Structure and Contradiction in Social Analysis. London: MacMillan Press.Google Scholar
Gladwell, M. (2000). The Tipping Point. Boston: Little, Brown and Company.Google Scholar
Gleick, J. (1987). Chaos – Making a New Science. London: Sphere Books.Google Scholar
Goodrum, M. R. (2008). Questioning thunderstones and arrowheads: the problem of recognizing and interpreting stone artifacts in the seventeenth century. Early Science and Medicine, 13(5), 482508. doi:10.1163/157338208X345759.CrossRefGoogle Scholar
Gould, S. J. (2002). The Structure of Evolutionary Theory. Cambridge, MA: Harvard University Press.Google Scholar
Gräslund, B. (1987). The Birth of Prehistoric Chronology: Dating Methods and Dating Systems in Nineteenth-Century Scandinavian Archaeology. Cambridge: Cambridge University Press.Google Scholar
Grey, S. A., Gray, S., De Kok, J. L., Helfgott, A. E. R., O’Dwyer, B., Jordan, R., & Nyaki, A. (2015). Using fuzzy cognitive mapping as a participatory approach to analyze change, preferred states, and perceived resilience of social-ecological systems. Ecology and Society, 20(2), 11.Google Scholar
Gunderson, L. H. & Holling, C. S. (eds.). (2001). Panarchy: Understanding Transformations in Human and Natural Systems. Washington, DC: Island Press.Google Scholar
Hägerstrand, T. (1970). What about people in regional science? Papers of the Regional Science Association, 24(1), 621. doi:10.1007/BF01936872.CrossRefGoogle Scholar
Hägerstrand, T. (1983). In search for the sources of concepts. In Buttimer, A., ed., The Practice of Geography, pp. 238–56. London, New York: Longman.Google Scholar
Haury, L. R., McGowan, J. A., & Wiebe, P. H. (1978). Patterns and processes in the time-space scales of plankton distribution. In Steele, J. H., ed., Spatial Pattern in Plankton Communities, pp. 277327. New York: Springer Science+Business Media.Google Scholar
Holling, C. S. (1973). Resilience and stability of ecological systems. Annual Review of Ecology and Systematics, 4, 123.CrossRefGoogle Scholar
Homer-Dixon, T., Walker, B., Biggs, R., Crépin, A.-S., Folke, C., Lambin, E. F., Peterson, G. D., Rockström, J., Scheffer, M., Steffen, W., & Troell, M. (2015). Synchronous failure: the emerging causal architecture of global crisis. Ecology and Society, 20(3), 6.CrossRefGoogle Scholar
Hutchinson, G. E. (1957). Concluding remarks. Cold Spring Harbor Symposia on Quantitative Biology, 22, 415–27. Reprinted (1991): Classics in Theoretical Biology. Bulletin of Mathematical Biology, 53, 193–213.Google Scholar
Hutchinson, G. E. (1957–93). A Treatise on Limnology. Vol. I Geography, Physics and Chemistry (1957) Wiley. Vol. II Introduction to Lake Biology and the Limnoplankton (1967) Wiley. Vol. III Limnological Botany (1975) Wiley. Vol. IV The Zoobenthos (1993). New York: Wiley.Google Scholar
Johnson, N. (2007). Simply Complexity – A Clear Guide to Complexity Theory. Oxford: Oneworld, Oxford University Press.Google Scholar
Kauffman, S. (1993). The Origins of Order: Self Organization and Selection in Evolution. Oxford: Oxford University Press.Google Scholar
Kauffman, S. (1995). At Home in the Universe: The Search for Laws of Self-Organization and Complexity. Oxford: Oxford University Press.Google Scholar
Kidambi, P. (2011). Time, temporality and history. In Gunn, S. and Faire, L., eds., Research Methods for the Arts and the Humanities: Research Methods for History, pp. 220–37. Edinburgh, GBR: Edinburgh University Press. ProQuest ebrary. Web. 2 November 2014.Google Scholar
King, D. (2005). Finding Atlantis: A True Story of Genius, Madness, and an Extraordinary Quest for a Lost World. New York: Harmony Books.Google Scholar
Knapp, B., ed. (1992). Archaeology, Annales and Ethnohistory. Cambridge: Cambridge University Press.Google Scholar
Lane, D., Pumain, D., van der Leeuw, S., & West, G., eds. (2009). Complexity Perspectives in Innovation and Social Change. Springer.Google Scholar
Le Roy Ladurie, E. (1988). Times of Feast, Times of Famine: A History of Climate since the Year 1000. New York: Noonday Press – Farrar, Straus & Giroux.Google Scholar
Lenton, T. M. (2013). What early warning systems are there for environmental shocks? Environmental Science & Policy, 27, S60S75.Google Scholar
Levin, S. A. (1992). The problem of pattern and scale in ecology. Ecology, 73, 1943–67.Google Scholar
Levin, S. A. (1998). Ecosystems and the biosphere as complex adaptive systems. Ecosystems, 1, 431–6.Google Scholar
Linnaeus, C. (1735). Systema naturae, sive regna tria naturae systematice proposita per classes, ordines, genera, & species. Leiden: Haak.Google Scholar
Lubbock, J. (1865). Pre-historic Times as Illustrated by Ancient Remains, and the Manners and Customs of Modern Savages. London & Edinburgh: Williams and Norgate.Google Scholar
Lucas, G. (2005). The Archaeology of Time. London: Routledge.Google Scholar
Lyell, C. (1830). Principles of Geology. 1st edition. London: John Murray.Google Scholar
May, R. M. (1977). Threshold and breakpoints in ecosystems with a multiplicity of states. Nature, 267(5628), 471–7.Google Scholar
Mbiti, J. S. (1990). African Religions and Philosophy. London: Heinemann.Google Scholar
McGlade, J. (1995). Archaeology and the ecodynamics of human-modified landscapes. Antiquity, 69(262), 113–32.Google Scholar
McGlade, J. (1999). The times of history: archaeology, narrative and non linear causality. In Murray, T., ed., Time and Archaeology, pp. 139–63. London: Routledge.Google Scholar
McGlade, J. (2006). Ecohistorical regimes and la longue durée: an approach to mapping long term societal change. In Garnsey, E. and McGlade, J., eds., Complexity and Co-evolution, Continuity and Change in Socio-economic Systems, pp. 77114. Cheltenham: Edward Elgar.Google Scholar
Meadows, D. H. (2008). Thinking in Systems: A Primer. White River Junction, VT: Chelsea Green Publishing.Google Scholar
Merrill, R. M. (2012). An Introduction to Epidemiology, 5th edition. Sudbury, MA: Jones & Bartlett.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
Murray, T. (1999). Time and Archaeology. London: Routledge.Google Scholar
Needham, J. (1954–9). Science and Civilization in China. Cambridge: Cambridge University Press.Google Scholar
Odum, E. P. & Odum, H. T. (1953). Fundamentals of Ecology. Philadelphia: W. B. Saunders.Google Scholar
Olsson, L., Jerneck, A., Thoren, H., Persson, J., & O’Byrne, D. (2015). Why resilience is unappealing to social science: theoretical and empirical investigations of the scientific use of resilience. Science Advantages, 1(4), e1400217.Google Scholar
Pace, M. L., Carpenter, S. R., & Cole, J. J. (2015). With and without warning: managing ecosystems in a changing world. Frontiers in Ecology and Environment, 13, 460–7.Google Scholar
Parrott, L. & Lange, H. (2013). An introduction to complexity science. In Messier, C., Klaus, J., Puettmann, K., & Coates, D., eds., Managing Forests as Complex Adaptive Systems: Building Resilience to the Challenge of Global Change. London: Earthscan, Routledge.Google Scholar
Peterson, G. D. (2002). Contagious disturbance, ecological memory, and the emergence of landscape pattern. Ecosystems, 5, 329–38.Google Scholar
Pielou, E. (1975). Ecological Diversity. New York: Wiley.Google Scholar
Redman, C. L. (2005). Resilience theory in archaeology. American Anthropologist, 107(1), 7077.Google Scholar
Robinson, K. S. (2009). Galileo’s Dream. London: HarperVoyager.Google Scholar
Rogers, E. (2003). Diffusion of Innovations, 5th edition. New York: Simon & Schuster.Google Scholar
Samford, P. A. (2007). Pits and the Archaeology of Slavery in Colonial Virginia. Tuscaloosa: University of Alabama Press.Google Scholar
Scheffer, M., Bascompte, J., Brock, W., Brovkin, V., Carpenter, S., Dakos, V., Held, H., van Nes, E. H., Rietkerk, M., & Sugihara, G. (2009). Early-warning signals for critical transitions. Nature, 461(3), 53–9.Google Scholar
Schneider, D. C. (2001). The rise of the concept of scale in ecology. BioScience, 51(7), 545–53.Google Scholar
Scholes, R. J., Reyers, B., Biggs, R., Spierenburg, M. J., & Duriappah, A. (2013). Multi-scale and cross-scale assessments of social-ecological systems and their ecosystem services. Current Opinion in Environmental Sustainability, 5, 1625.Google Scholar
Scott, J. C. (2009). The Art of Not Being Governed: An Anarchist History of Upland Southeast Asia. New Haven, CT: Yale University Press.Google Scholar
Soranno, P. A., Cheruvelil, K. S., Bissell, E., Bremigan, M. T., Downing, J. A., Fergus, C. E., Filstrup, C., Lottig, N. R., Henry, E. N., Stanley, E. H., Stow, C. A., Tan, P.-N., Wagner, T., & Webster, K. E. (2014). Cross-scale interactions: quantifying multi-scaled cause-effect relationships in macrosystems. Frontiers in Ecology and the Environment, 12(1), 6573.Google Scholar
Spengler, O. (1991) The Decline of the West, vol. 1:21–2. New York: Oxford University Press (original publication 1918).Google Scholar
Steno, N. (Nicholas Steensen) (1669). De solido intra solidum naturaliter contento dissertationis prodromus. Firenze: Biblioteca Nazionale Centrale.Google Scholar
Sun, Z. Y. & Ren, H. (2011). Ecological memory and its potential applications in ecology: a review. (in Chinese with English abstract). Chinese Journal of Applied Ecology, 22(3), 549–55.Google Scholar
Tainter, J. A. (2006). Social complexity and sustainability. Ecological Complexity, 3, 91103.Google Scholar
Tainter, J. A. & Crumley, C. L. (2007). Climate, complexity, and problem solving in the Roman empire. In Costanza, R., Graumlich, L. J., & Steffen, W., eds., Sustainability or Collapse? An Integrated History and Future of People on Earth, pp. 6175. Boston: MIT Press.Google Scholar
Tomich, D. (1958). The order of historical time: the longue durée and micro history. The longue durée and world systems analysis. Histoire et sciences sociales. Annales E.S.C. XIII, 4.Google Scholar
Trigger, B. (2006). A History of Archaeological Thought, 2nd edition. Oxford: Cambridge University Press.Google Scholar
Turchin, P. (2003). Historical Dynamics: Why States Rise and Fall. Princeton, NJ: Princeton University Press.Google Scholar
Van der Leeuw, S., ed. (1998). The Archaeomedes Project – Understanding the Natural and Anthropogenic Causes of Land Degradation and Desertification in the Mediterranean. Luxemburg: Office for Official Publications of the European Union.Google Scholar
Van der Leeuw, S. & McGlade, J. (1997). Archaeology: Time, Process and Structural Transformations. London: Routledge.Google Scholar
Vance, T. C. & Doel, R. E. (2010). Graphical methods and Cold War scientific practice: the Stommel diagram’s intriguing journey from the physical to the biological environmental sciences. Historical Studies in the Natural Sciences, 40(1), 147.Google Scholar
Vernadsky, V. (1929). La Biosphère. Paris: Librairie Félix Alcan.Google Scholar
Watson, P. J., LeBlanc, S., & Redman, C. (1971). Explanation in Archaeology: An Explicitly Scientific Approach. New York: Columbia University Press.Google Scholar
Webb, C. & Bodin, Ö. (2008). A network perspective on modularity and control of flow in robust systems. In Norberg, J. and Cumming, G. S., eds., Complexity Theory for a Sustainable Future, pp. 85118. New York: Columbia University Press.Google Scholar
Werners, S. E., Pfenniger, S., van Slobbe, E., Haasnoot, M., Kwakkel, J. H. & Swart, R. J. (2013). Thresholds, tipping and turning points for sustainability under climate change. Current Opinion in Environmental Sustainability, 5, 334–40.Google Scholar

References

Abel, W. (1980). Agricultural Fluctuations in Europe: From the Thirteenth to the Twentieth Centuries. London: Methuen.Google Scholar
Afifi, T. & Jäger, J. eds. (2010). Environment, Forced Migration and Social Vulnerability. Heidelberg: Springer-Verlag Berlin.Google Scholar
Alexander, P. (1987). Le Climate en Europe au moyen âge: Contribution à l’histoire des variations climatiques de 1000 à 1425, d’après les sources narratives de l’Europe occidentale. Paris: Éditions de l’École des hautes études en sciences sociales [in French].Google Scholar
Alley, R. B. (2000). The Younger Dryas cold interval as viewed from central Greenland. Quaternary Science Reviews, 19, 213–26.CrossRefGoogle Scholar
Andersen, B. G. & Borns, H. W. (1997). The Ice Age World: An Introduction to Quaternary History and Research with Emphasis on North America and Northern Europe during the Last 2.5 Million Years. Oslo: Scandinavian University Press.Google Scholar
Anderson, D. G., Maasch, K., & Sandweiss, D. H., eds. (2007). Climate Change and Cultural Dynamics: A Global Perspective on Mid-Holocene Transitions. London: Academic Press.Google Scholar
Anderson, R. W., Johnson, N., & Koyama, D. M. (2017). Jewish persecutions and weather shocks: 1100–1800. The Economic Journal, 127, 924–58.CrossRefGoogle Scholar
Annan, J. D. & Hargreaves, J. C. (2013). A new global reconstruction of temperature changes at the Last Glacial Maximum. Climate of the Past, 9, 367–76.Google Scholar
Appleby, A. B. (1980). Epidemics and famine in the Little Ice Age. The Journal of Interdisciplinary History, 10, 643–63.Google Scholar
Arjava, A. (2005). The mystery cloud of 536 CE in the Mediterranean sources. Dumbarton Oaks Papers, 59, 7394.Google Scholar
Arnold, D. (1988). Famine: Social Crisis and Historical Change. Oxford: Basil Blackwell.Google Scholar
Babst, F., Poulter, B. & Trouet, V., Tan, K., Neuwirth, B., Wilson, R., Carrer, M., Grabner, M., Tegel, W., Levanic, T., Panayotov, M., Urbinati, C., Bouriaud, O., Ciais, P., & Frank, D. (2013). Site- and species-specific responses of forest growth to climate across the European continent. Global Ecology and Biogeography, 22, 706–17.Google Scholar
Barlow, L. K., Sadler, J. P. & Ogilvie, A. E. J., Buckland, P. C., Amorosi, T., Ingimundarson, J. H., Skidmore, P., Dugmore, A., & McGowan, T. H. (1997). Interdisciplinary investigations of the end of the Norse Western Settlement in Greenland. The Holocene, 7, 489–99.CrossRefGoogle Scholar
Barnett, J. (2003). Security and climate change. Global Environmental Change, 13, 717.Google Scholar
Bartlein, P. J., Harrison, S. P. & Brewer, S., Connor, S., Davis, B. A. S., Gajewski, K., Joel, G., Harrison-Prentice, T. I., Henderson, A. P., Peyron, O., Prentice, I. C., Scholze, M., Seppä, H., Shuman, B., Sugita, S., Thompson, R., Viau, A. E., Williams, J. W., & Wu, H. (2011). Pollen-based continental climate reconstructions at 6 and 21 ka: a global synthesis. Climate Dynamics, 37, 775802.Google Scholar
Behringer, W. (1995). Weather, hunger and fear: the origins of the European witch persecution in climate, society and mentality. German History, 13, 127.Google Scholar
Behringer, W. (1999). Climate change and witch-hunting: the impact of the Little Ice Age on mentalities. Climate Change, 43, 335–51.Google Scholar
Behringer, W. (2010). A Cultural History of Climate. London: Polity Press.Google Scholar
Bender, M., Sowers, T., & Brook, E. (1997). Gases in ice cores. Proceedings of the National Academy of Sciences of the United States of America, 94, 8343–9.Google Scholar
Benson, L. V. (2010). Who provided maize to Chaco Canyon after the mid-12th-century drought?. Journal of Archaeological Science, 37, 621–9.Google Scholar
Benson, L. & Berry, M. S. (2009). Climate change and cultural response in the prehistoric American Southwest. Kiva, 75, 89119.Google Scholar
Benson, L., Kashgarian, M., Rye, R., Lund, S., Paillet, F., Smoot, J., Kester, C., Mensing, S., Meko, D., & Landström, S. (2002). Holocene multidecadal and multicentennial droughts affecting Northern California and Nevada. Quaternary Science Reviews, 21, 659–82.Google Scholar
Benson, L., Petersen, K., & Stein, J. (2007). Anasazi (pre-Columbian Native American) migrations during the middle-12th and late-13th centuries – Were they drought induced?. Climatic Change, 83, 187213.Google Scholar
Berger, A. & Loutre, M. F. (1991). Insolation values for the climate of the last 10 million years. Quaternary Science Reviews, 10, 297317.Google Scholar
Blinman, E. (2008). 2000 years of cultural adaptation to climate change in the Southwestern United States. AMBIO: A Journal of the Human Environment, 37, 489–97.Google Scholar
Bocinsky, R. K. & Kohler, T. A. (2014). A 2,000-year reconstruction of the rain-fed maize agricultural niche in the US Southwest. Nature Communications, 5, 5618. doi: 10.1038/ncomms6618.Google Scholar
Bond, G., Kromer, B. & Beer, J., Muscheler, R., Evans, M. N., Showers, W., Hoffmann, S., Lotti-Bond, R., Hajdas, I., & Bonani, G. (2001). Persistent solar influence on North Atlantic climate during the Holocene. Science, 294, 2130–6.Google Scholar
Bond, G., Showers, W. & Cheseby, M., Lotti, W., Almasi, P., Demenocal, P. B., Priore, P., Cullen, H., Hajdas, I., & Bonani, G. (1997). A pervasive millennial-scale cycle in North Atlantic Holocene and glacial climates. Science, 278, 1257–66.Google Scholar
Bond, G. C., Showers, W., Elliot, M., Evans, M., Lotti, R., Hajdas, I., Beosnan, G., & Johnson, S. (1999). The North Atlantic’s 1–2 kyr climate rhythm’s relation to Heinrich Events, Dansgaard/Oeschger Cycles and the Little Ice Age. In Clark, P. U., Webb, R. S., & Keigwin, L. D., eds., Mechanisms of Global Climate Change at Millennial Time Scales. Washington, DC: Geophysical Monograph Series. American Geophysical Union, pp. 3558.Google Scholar
Bradley, R. S. (1999). Paleoclimatology: Reconstructing Climates of the Quaternary. San Diego, CA: Academic Press.Google Scholar
Brázdil, R., Pfister, C., Wanner, H., von Storch, H., & Luterbacher, J. (2005). Historical climatology in Europe – The state of the art. Climatic Change, 70, 363430.Google Scholar
Breitenmoser, P., Beer, J., & Brönnimann, S. (2012). Solar and volcanic fingerprints in tree-ring chronologies over the past 2000 years. Palaeogeography, Palaeoclimatology, Palaeoecology, 313–14, 127–39.Google Scholar
Broecker, W. S. & Putnam, A. E. (2013). Hydrologic impacts of past shifts of Earth’s thermal equator offer insight into those to be produced by fossil fuel CO2. Proceedings of the National Academy of Sciences of the United States of America, 110, 16710–15.Google Scholar
Brooke, J. L. (2014). Climate Change and the Course of Global History: A Rough Journey. New York: Cambridge University Press.Google Scholar
Bryson, R. A. & Murray, T. J. (1977). Climates of Hunger: Mankind and the World’s Changing Weather. Madison: University of Wisconsin Press.Google Scholar
Buckland, P. C., Amorosi, T. & Barlow, L. K., Dugmore, A. J., Mayewski, P. A., McGovern, T. H., Ogilvie, A. E. J., Sadler, J. P., & Skidmore, P. (1996). Bioarchaeological and climatological evidence for the fate of Norse farmers in medieval Greenland. Antiquity, 70, 8896.Google Scholar
Buhaug, H. (2010). Climate not to blame for African civil wars. Proceedings of the National Academy of Sciences of the United States of America, 107 (38), 16477–82.Google Scholar
Büntgen, U., Myglan, V. & Ljungqvist, F. C. McCormic, M., Di Cosmo, N., Sigl, M., Jungclaus, J. H., Wagner, S., Krusic, P. J., Esper, J., Kaplan, J. O., De Vaan, M. A. C., Luterbacher, J., Wacker, L., Tegel, W., & Kirdyanov, . (2016). Cooling and societal change during the Late Antique Little Ice Age from 536 to around 660 AD. Nature Geoscience, 9, 231–6.Google Scholar
Campbell, B. M. S. (2016). The Great Transition: Climate, Disease and Society in the Late-Medieval World. Cambridge: Cambridge University Press.Google Scholar
Carey, M. (2012). Climate and history: a critical review of historical climatology and climate change historiography. Wiley Interdisciplinary Reviews: Climate Change, 3, 233–49.Google Scholar
Chen, J., Chen, F. & Feng, S., Huang, W., Liu, J., & Zhou, A. (2015). Hydroclimatic changes in China and surroundings during the Medieval Climate Anomaly and Little Ice Age: spatial patterns and possible mechanisms. Quaternary Science Reviews, 107, 98111.CrossRefGoogle Scholar
Chen, J., Rao, Z., Liu, J., Huang, W., Feng, S., & Dong, G. (2016). On the timing of the East Asian summer monsoon maximum during the Holocene – Does the speleothem oxygen isotope record reflect monsoon rainfall variability? Science China Earth Sciences, 59, 2328–38.CrossRefGoogle Scholar
Chen, Q. (2015). Climate shocks, dynastic cycles and nomadic conquests: evidence from historical China. Oxford Economic Papers, 67, 185204.Google Scholar
Cheyette, F. L. (2008). The disappearance of the ancient landscape and the climatic anomaly of the early Middle Ages: a question to be pursued. Early Medieval Europe, 16, 127–65.Google Scholar
Christiansen, B. & Ljungqvist, F. C. (2012). The extra-tropical Northern Hemisphere temperature in the last two millennia: reconstructions of low-frequency variability. Climate of the Past, 8, 765–86.Google Scholar
Christiansen, B. & Ljungqvist, F. C. (2017). Challenges and perspectives for large-scale temperature reconstructions of the past two millennia. Reviews of Geophysics, 55, 40–96.Google Scholar
Chu, C. Y. C. & Lee, R. D. (1994). Famine, revolt, and the dynastic cycle–population dynamics in historical China. Journal of Population Economics, 7, 351–78.Google Scholar
Compo, G. P., Whitaker, J. S., Sardeshmukh, P. D., Matsui, N., Allan, R. J., Yin, X., Gleason, Jr., B. E., Vose, R. S., Rutledge, G., Bessemoulin, P., Brönnimann, S., Brunet, M., Crouthamel, R. I., Grant, A. N., Groisman, P. Y., Jones, P. D., Kruk, M. C., Kruger, A. C., Marshall, G. J., Maugeri, M., Mok, H. Y., Nordli, Ø., Ross, T. F., Trigo, R. M., Wang, X. L., Woodruff, S. D., & Worley, S. J. (2011). The twentieth century reanalysis project. Quarterly Journal of the Royal Meteorological Society, 137, 128.Google Scholar
Cook, B. I., Smerdon, J. E., Seager, R., & Coats, S. (2014). Global warming and 21st century drying. Climate Dynamics, 43, 2607–27.Google Scholar
Cook, E. R., Anchukaitis, K. J., Buckley, B. M., Jacoby, G. C., & Wright, W. E. (2010). Asian monsoon failure and megadrought during the last millennium. Science, 328, 486–9.CrossRefGoogle ScholarPubMed
Cook, E. R., Briffa, K. R., Meko, D. M., Graybill, D. A., & Funkhouser, G. (1995). The ‘segment length curse’ in long tree-ring chronology development for palaeoclimatic studies. The Holocene, 5, 229–37.Google Scholar
Cook, E. R., Woodhouse, C. A., & Eakin, C. M. (2004). Long term aridity changes in the western United States. Science, 306, 1015–18.Google Scholar
Crumley, C. L. (1993). Historic ecotonal shifts. Ecological Applications, 3, 377–84.Google Scholar
Cullen, H., deMenocal, P. B. & Hemming, S., Hemming, G., Brown, F. H., Gulderson, T., & Sirocko, F. (2000). Climate change and the collapse of the Akkadian empire: evidence from the deep sea. Geology, 28, 379–82.Google Scholar
d’Alpoim Guedes, J. A., Crabtree, S. A., Bocinsky, R. K., & Kohler, T. A. (2016). Twenty-first century approaches to ancient problems: climate and society. Proceedings of the National Academy of Sciences of the United States of America, 113, 14483–91.Google Scholar
D’Arrigo, R., Frank, D., Jacoby, G., & Pederson, N. (2001). Spatial response to major volcanic events in or about AD 536, 934 and 1258: frost rings and other dendrochronological evidence from Mongolia and northern Siberia. Climatic Change, 49, 239–46.Google Scholar
deMenocal, P. B. (2001). Cultural responses to climate change during the Late Holocene. Science, 292, 667–73.Google Scholar
Diamond, J. (2005). Collapse: How Societies Choose to Fail or Succeed. New York: Penguin Books.Google Scholar
Diaz, H. & Trouet, V. (2014). Some perspectives on societal impacts of past climatic changes. History Compass, 12, 160–77.Google Scholar
Douglas, P. M. J., Demarest, A. A., Brenner, M., & Canuto, M. A. (2016). Impacts of climate change on the collapse of lowland Maya civilization. Annual Review of Earth and Planetary Sciences, 44, 613–45.Google Scholar
Dugmore, A. J., Keller, C., & McGovern, T. H. (2007). Norse Greenland settlement: reflections on climate change, trade, and the contrasting fates of human settlements in the North Atlantic islands. Arctic Anthropology, 44, 1236.Google Scholar
Dykoski, C. A., Edwards, R. L., Chen, H., Yuan, D., Cai, Y., Zhang, M., Lin, Y., An, Z., & Revenaugh, J. (2005). A high-resolution, absolute-dated Holocene and deglacial Asian monsoon record from Dongge Cave, China. Earth and Planetary Science Letters, 233, 7186.Google Scholar
Esper, J., Cook, E. R., Krusic, P. J., Peters, K., & Schweingruber, F. H. (2003). Tests of the RCS method for preserving low-frequency variability in long tree-ring chronologies. Tree-Ring Research, 59, 8198.Google Scholar
Esper, J., Krusic, P. J., Ljungqvist, F. C., Luterbacher, J., Carrer, M., Cook, E., Davi, N. K., Hartl-Marier, C., Kirdyanov, A., Konter, O., Myglan, V., Timonen, M., Treydte, K., Truet, V., Villalba, R., Yang, B., & Büntgen, U. (2016). Ranking of tree-ring based temperature reconstructions of the past millennium. Quaternary Science Reviews, 145, 134–51.Google Scholar
Fagan, B. M. (2000). The Little Ice Age: How Climate Made History, 1300–1850. New York: Basic Books.Google Scholar
Fairchild, I. J. & Baker, A. (2012), Speleothem Science: From Process to Past Environments. Oxford: Wiley.Google Scholar
Fan, K.-W. (2010). Climatic change and dynastic cycles in Chinese history: a review essay. Climatic Change, 101, 565–73.Google Scholar
Fan, K.-W. (2015). Climate change and Chinese history: a review of trends, topics, and methods. Wiley Interdisciplinary Reviews: Climate Change, 6, 225–38.Google Scholar
Fang, J. & Liu, G. (1992). Relationship between climatic change and the nomadic southward migrations in eastern Asia during historical times. Climatic Change, 22, 1511–69.Google Scholar
Finné, M., Holmgren, , Sundqvist, K., Weiberg, H. S., , E., & Lindblom, M. (2011): Climate in the eastern Mediterranean, and adjacent regions, during the past 6000 years – A review. Journal of Archaeological Science, 38, 3153–73.Google Scholar
Fritts, H. (1976). Tree Rings and Climate. London: Academic Press.Google Scholar
Fyfe, J. C., Meehl, G. A. & England, M. H., Mann, M. E., Santer, B. D., Flato, G. M., Hawkins, E., Gillet, N. P., Xie, S.-P., Kosaka, Y., & Swart, N. C. (2016). Making sense of the early-2000s warming slowdown. Nature Climate Change, 6, 224–8.Google Scholar
Ge, Q., Hao, Z., Zheng, J., & Shao, X. (2013). Temperature changes over the past 2000 yr in China and comparison with the Northern Hemisphere. Climate of the Past, 9, 1153–60.Google Scholar
Ge, Q., Zheng, J., Tian, Y., Wu, W., Fang, X., & Wang, W.-C. (2008). Coherence of climatic reconstruction from historical documents in China by different studies. International Journal of Climatology, 28, 1007–24.Google Scholar
Gill, R. B. (2000). The Great Maya Droughts: Water, Life, and Death. Albuquerque: University of New Mexico Press.Google Scholar
Glaser, R. (2008). Klimageschichte Mitteleuropas. 1200 Jahre Wetter, Klima, Katastrophen. Darmstadt: Wissenschaftliche Buchgesellschaft [in German].Google Scholar
Gleisner, H., Thejll, P., Christiansen, B., & Nielsen, J. K. (2015). Recent global warming hiatus dominated by low-latitude temperature trends in surface and troposphere data. Geophysical Research Letters, 42, 510–17. doi: 10.1002/2014GL062596.Google Scholar
Gray, L. J., Beer, J. & Geller, M., Haigh, J. D., Lockwood, M., Matthes, K., Cubasch, U., & Fleitmann, D. (2010). Solar influences on climate. Reviews of Geophysics, 48, RG4001. doi: 10.1029/2009RG000282.Google Scholar
Gräslund, B. & Price, N. (2012). Twilight of the gods? The ‘dust veil event’ of AD 536 in critical perspective. Antiquity, 86, 428–43.Google Scholar
Gunn, J. D., ed. (2000). The Years Without Summer. Tracing A.D. 536 and its Aftermath. Oxford: Archaeopress.Google Scholar
Halstead, P. & O’Shea, J., eds. (1989). Bad Year Economics: Cultural Responses to Risk and Uncertainty. New York: Cambridge University Press.Google Scholar
Hansen, J., Ruedy, R., Sato, M., & Lo, K. (2010). Global surface temperature change. Reviews of Geophysics, 48, RG4004. doi: 10.1029/2010RG000345.Google Scholar
Hansen, J., Sato, M. & Ruedy, R., Lo, K., Lea, D. W., & Medina-Elizade, M. (2006). Global temperature change. Proceedings of the National Academy of Sciences of the United States of America, 103, 14288–93.Google Scholar
Hao, Z., Zheng, J., Ge, Q., & Zhang, X. (2012). Spatial patterns of precipitation anomalies for 30-yr warm periods in China during the past 2000 years. Acta Meteorologica Sinica, 26, 278–88.Google Scholar
Hao, Z., Zheng, J. & Zhang, X., Liu, H., Li, M., & Ge, Q. (2016). Spatial patterns of precipitation anomalies in eastern China during centennial cold and warm periods of the past 2000 years. International Journal of Climatology, 26, 467–75.Google Scholar
Held, I. M. & Soden, B. J. (2006). Robust responses of the hydrological cycle to global warming. Journal of Climate, 19, 5686–99.Google Scholar
Hellmann, L., Nikolaev, A. & Ljungqvist, F. C., Churakova(Sidorova), O., Düthorn, E., Esper, J., Hülsmann, L., Kirdyanov, A. V., Moiseev, P., & Myglan, V. S. (2016). Diverse growth trends and climate responses across Eurasia’s boreal forest. Environmental Research Letters, 11, 074021. doi: 10.1088/1748–9326/11/7/074021.Google Scholar
Hiller, A., Boettger, T., & Kremenetski, C. (2001). Medieval climatic warming recorded by radiocarbon dated alpine tree-line shift on the Kola Peninsula, Russia. The Holocene, 11, 491–7.CrossRefGoogle Scholar
Hind, A., Zhang, Q., & Brattström, G. (2016). Problems encountered when defining Arctic amplification as a ratio. Scientific Reports, 6, 30469. doi: 10.1038/srep30469.Google Scholar
Holmes, J. A. (2008). How the Sahara became dry. Science, 320, 752–3.Google Scholar
Hsiang, S. M. & Burke, M. (2014). Climate, conflict, and social stability: what does the evidence say? Climate Change, 123, 3955.Google Scholar
Hsiang, S. M., Burke, M., & Miguel, E. (2013). Quantifying the influence of climate on human conflict. Science, 341, 1235367. doi: 10.1126/science.1235367.Google Scholar
Hsu, K. J. (1998). Sun, climate, hunger, and mass migration. Science in China Series D. Earth Sciences, 41, 449–72.Google Scholar
Hughes, M. K. & Graumlich, L. J. (1996). Multi-millennial dendroclimatic studies from the western United States. In Jones, P. D., Bradley, R. S., & Jouzel, J., eds., Climatic Variations and Forcing Mechanisms of the Last 2000 Years. Volume 141. NATO ASI Series, pp. 109–24.Google Scholar
Huang, S. P., Pollack, H. N., & Shen, P. Y. (2008). A late Quaternary climate reconstruction based on borehole heat flux data, borehole temperature data, and the instrumental record. Geophysical Research Letters, 35, L13703. doi: 10.1029/2008GL034187.Google Scholar
Hugo, G. & Currey, B. (1989). Famine: As a Geographical Phenomenon. Dordrecht, Springer.Google Scholar
Huntington, E. (1907). The Pulse of Asia: A Journey in Central Asia Illustrating the Geographic Basis of History. Boston: Houghton, Mifflin and Company.Google Scholar
Huntington, E. (1913). Changes of climate and history. American Historical Review, 19, 213–32.Google Scholar
IPCC. (2013). Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T. F., Qin, D., Plattner, G. K., Tignor, M., Allen, S. K., Boschung, J., Nauels, A., Xia, Y., Bex, V., & Midgley, P. M., eds.]. Cambridge, United Kingdom and New York, NY, USA: Cambridge University Press. doi: 10.1017/CBO9781107415324.Google Scholar
Issar, A. & Zohar, M. (2004). Climate Change: Environment and Civilization in the Middle East. Berlin: Springer.Google Scholar
Jones, P. (2016). The reliability of global and hemispheric surface temperature records. Advances in Atmospheric Sciences, 33, 269–82.Google Scholar
Jones, P. D., Briffa, K. R. & Osborn, T. J., Loughm, J. M., & van Ommen, T. D. (2009). High-resolution palaeoclimatology of the last millennium: a review of current status and future prospects. The Holocene, 19, 349.Google Scholar
Jouzel, J. (2013). A brief history of ice core science over the last 50 yr. Climate of the Past, 9, 2525–47.Google Scholar
Karlén, W. & Kuylenstierna, J. (1996). On solar forcing of Holocene climate: evidence from Scandinavia. The Holocene, 6, 359–36.Google Scholar
Karlén, W. & Larsson, L. (2007). Mid-Holocene climatic and cultural dynamics in Northern Europe. In Anderson, D. G., Maasch, K., & Sandweiss, D. H., eds., Climate Change and Cultural Dynamics: A Global Perspective on Mid-Holocene Transitions. London: Academic Press, pp. 407–34.Google Scholar
Klomp, J. & Bulte, E. (2013). Climate change, weather shocks, and violent conflict: a critical look at the evidence. Agricultural Economics, 44, 6378.Google Scholar
Krämer, D. (2015). Menschen Grasten nun mit dem Vieh. Die Letzte Grosse Hungerkrise der Schweiz 1816/17. Basel: Schwabe [in German].Google Scholar
Kullman, L. (2015). Higher-than-present Medieval pine (Pinus sylvestris L.) treeline along the Swedish Scandes. Landscape Online, 42, 114.Google Scholar
Kuper, R. & Kröpelin, S. (2006). Climate-controlled Holocene occupation in the Sahara: motor of Africa’s evolution. Science, 313, 803–7.Google Scholar
Laird, K. R., Fritz, S. C., Maasch, K. A., & Cumming, B. F. (1996). Greater drought intensity and frequency before AD 1200 in the northern Great Plains, USA. Nature, 384, 552–4.Google Scholar
Lamb, H. H. (1972–7). Climate: Present, Past and Future 1–2. London: Methuen.Google Scholar
Larsen, L. B., Vinther, B. M. & Briffa, K. R., Melvin, T. M., Clausen, H. B., Jones, P. D., Andersen, M. L. S., Hammer, C. U., Eronen, M., Grudd, H., Gunnarson, B. E., Hantemirov, R. M., Naurzbaev, M. M., & Nicolussi, K. (2008). New ice core evidence for a volcanic cause of the A.D. 536 dust veil. Geophysical Research Letters, 35, L04708. doi: 10.1029/2007GL032450.Google Scholar
Leduc, G., Schneider, R., Kim, J.-H., & Lohmann, G. (2010). Holocene and Eemian sea surface temperature trends as revealed by Alkenone and Mg/Ca paleothermometry. Quaternary Science Reviews, 29, 9891004.Google Scholar
Lee, H. F. (2014). Climate-induced agricultural shrinkage and overpopulation in late imperial China. Climate Research, 59, 229–42.Google Scholar
Lee, H. F. & Zhang, D. D. (2015). Quantitative analysis of climate change and human crises in history. In Kwan, M.-P., Richardson, D., Wang, D., & Zhou, C., eds., Space-Time Integration in Geography and GIScience: Research Frontiers in the US and China. Dordrecht: Springer, pp. 235–67.Google Scholar
Liu, Y., Cai, Q. F., Song, H. M., An, Z. S., & Linderholm, H. W. (2011). Amplitudes, rates, periodicities and causes of temperature variations in the past 2485 years and future trends over the central-eastern Tibetan Plateau. Chinese Science Bulletin, 6, 2986–94.Google Scholar
Ljungqvist, F. C. (2009). Global nedkylning: Klimatet och människan under 10 000 år. Stockholm: Norstedts [in Swedish].Google Scholar
Ljungqvist, F. C. (2011). The spatio-temporal pattern of the Mid-Holocene Thermal Maximum. Geografie, 116, 91110.Google Scholar
Ljungqvist, F. C., Krusic, P. J., Brattström, G., & Sundqvist, H. S. (2012). Northern Hemisphere temperature patterns in the last 12 centuries. Climate of the Past, 8, 227–49.Google Scholar
Ljungqvist, F. C., Krusic, P. J. & Sundqvist, H. S., Zorita, E., Brattström, G., & Frank, D. (2016). Northern Hemisphere hydroclimatic variability over the past twelve centuries. Nature, 532, 94–8.Google Scholar
Löwenborg, D. (2012). An Iron Age shock doctrine: did the AD 536–7 event trigger large-scale social changes in the Mälaren valley area? Journal of Archaeology and Ancient History, 4, 129.Google Scholar
Lucero, L. J. (2006). Water and Ritual: The Rise and Fall of Classic Maya Rulers. Austin: University of Texas Press.Google Scholar
Luterbacher, J. & Pfister, C. (2015). The year without a summer. Nature Geoscience, 8, 246–8.Google Scholar
Luterbacher, J., Werner, J. P. & Smerdon, J., Fernández-Donado, L., González-Rouco, F. J., Barriopedro, D., Ljungqvist, F. C, Büntgen, , Zorita, U., Wagner, E., Esper, S., McCarroll, J., Toreti, D., Frank, A., Jungclaus, D., Barriendos, J. H., Bertolin, M., Bothe, C., Brázdil, O., Camuffo, R., Dobrovolný, D., Gagen, P., García-Bustamante, M., Ge, E., Gómez-Navarro, Q., Guiot, J. J., Hao, J., Hegerl, Z., Holmgren, G. C., Klimenko, K., Martín-Chivelet, V. V., Pfister, J., Roberts, C., Schindler, N., Schurer, A., Solomina, A., von Gunten, O., Wahl, L., Wanner, E., Wetter, H., Xoplaki, O., Yuan, E., Zanchettin, N., Zhang, D., , H., & Zerefos, C. (2016). European summer temperatures since Roman times. Environmental Research Letters, 11, 024001. doi: 10.1088/1748–9326/11/1/024001.Google Scholar
Madsen, C. K. (2014) ‘Pastoral Settlement, Farming, and Hierarchy in Norse Vatnahverfi, South Greenland’. Copenhagen: University of Copenhagen, unpublished PhD thesis.Google Scholar
Mann, M. E., Zhang, Z. & Hughes, M. K., Bradley, R. S., Miller, S. K., Rutherford, S., & Ni, F. (2008). Proxy-based reconstructions of hemispheric and global surface temperature variations over the past two millennia. Proceedings of the National Academy of Sciences of the United States of America, 105, 13252–7.Google Scholar
Mann, M. E., Zhang, Z. & Rutherford, S., Bradley, R. S., Hughes, M. K., Shindell, D., Ammann, C., Faluvegi, G., & Ni, F. (2009). Global signatures and dynamical origins of the Little Ice Age and Medieval Climate Anomaly. Science, 326, 1256–60.Google Scholar
Matthews, J. A. & Briffa, K. R. (2005). The ‘Little Ice Age’: re-evaluation of an evolving concept. Geografiska Annaler, 87A, 1736.Google Scholar
Mayewski, P. A., Rohling, E. E. & Stager, J. C., Karlén, W., Maasch, , Meeker, K. A., Meyerson, L. D., Gasse, E. A., van Krevels, F., Holmgren, S., Lee-Thorp, K., Rosqvist, J., Rack, G., Staubvasser, F., Schneider, M., , R. R., & Steig, E. J. (2004). Holocene climate variability. Quaternary Research, 62, 243–55.Google Scholar
Mazepa, V. S. (2005). Stand density in the last millennium at the upper tree-line ecotone in the Polar Ural Mountains. Canadian Journal of Forest Research, 35, 2082–91.Google Scholar
McCormick, M., Büntgen, U. & Cane, , , M. A., Cook, , Harper, E. R., Huyers, K., Litt, P., Manning, T., Mayewski, S. W., More, P. A., Nicolussi, A. F. M., , K., & Tegel, W. (2012). Climate change during and after the Roman Empire: reconstructing the past from scientific and historical evidence. Journal of Interdisciplinary History, 4, 169220.Google Scholar
McMichael, A. J. (2012). Insights from past millennia into climatic impacts on human health and survival. Proceedings of the National Academy of Sciences, 109, 4730–7.Google Scholar
McNeill, J. R. (2016). Historians, superhistory, and climate change. In Jarrick, A., Myrdal, J., & Bondesson, M. Wallenberg, eds., Methods in World History: A Critical Approach. Lund: Nordic Academic Press, pp. 1943.Google Scholar
Meierding, E. (2013) Climate change and conflict: avoiding small talk about the weather. International Studies Review, 15, 185203.Google Scholar
Melander, E., Pettersson, T., & Themnér, L. (2016). Organized violence, 1989–2015. Journal of Peace Research, 53, 727–42.Google Scholar
Melvin, T. M. & Briffa, K. R. (2008). A ‘signal-free’ approach to dendroclimatic standardisation. Dendrochronologia, 26, 7186.Google Scholar
Moberg, A., Sonechkin, D. M., Holmgren, K., Datsenko, N. M., & Karlén, W. (2005). Highly variable Northern Hemisphere temperatures reconstructed from low- and high-resolution proxy data. Nature, 433, 613–17.Google Scholar
Morice, C. P., Kennedy, J. J., Rayner, N. A., & Jones, P. D. (2012). Quantifying uncertainties in global and regional temperature change using an ensemble of observational estimates: The HadCRUT4 data set. Journal of Geophysical Research: Atmospheres, 117, D08101. doi: 10.1029/2011JD017187.Google Scholar
Nelson, M. C., Ingram, S. E. & Dugmore, A. J., Streeter, R., Peeples, M. A., McGowern, T. H., Hegmin, M., Arneborg, J., Kintigh, K., Brewington, S., Spielmann, K. A., Simpson, I. A., Strawhacker, C., Comeau, L. E. L., Torvinen, A., Madsen, C. K., Hambrecht, G., & Smiarowski, K. (2016). Climate challenges, vulnerabilities, and food security. Proceedings of the National Academy of Sciences of the United States of America, 113, 298303.Google Scholar
Newman, L. F., ed. (1990). Hunger in History: Food Shortage, Poverty, and Deprivation. Cambridge/Oxford: Blackwell.Google Scholar
O’Loughlin, J., Linke, A. M., & Witmer, F. D. (2014). Modeling and data choices sway conclusions about climate–conflict links. Proceedings of the National Academy of Sciences of the United States of America, 111, 2054–5.Google Scholar
PAGES 2k Consortium (2013). Continental-scale temperature variability during the past two millennia. Nature Geoscience, 6, 339–46.Google Scholar
Parker, G. (2013). Global Crisis: War, Climate and Catastrophe in the Seventeenth Century. New Haven, CT: Yale University Press.Google Scholar
Parry, M. L. (1978). Climatic Change, Agriculture and Settlement. Folkestone: William Dawson & Sons.Google Scholar
Parry, M. L. & Carter, T. R. (1983). Assessing Impacts of Climatic Change in Marginal Areas: The Search for an Appropriate Methodology. Laxenburg: International Institute for Applied Systems Analysis.Google Scholar
Parry, M. L., Carter, T. R., & Konijn, N. T., eds. (1988a). The Impact of Climatic Variations on Agriculture. Volume 1. Assessments in Cool Temperate and Cold Regions. Dordrecht: Kluwer.Google Scholar
Parry, M. L., Carter, T. R., & Konijn, N. T., eds. (1988b). The Impact of Climatic Variations on Agriculture. Volume 2. Assessments in Semi-arid Regions. Dordrecht: Kluwer.Google Scholar
Payette, S., Filion, L., Delwaide, A., & Begin, C. (1989). Reconstruction of tree-line vegetation response to long-term climate change. Nature, 341, 429–32.Google Scholar
Pettersson, T. & Wallensteen, P. (2015). Armed conflicts, 1946–2014. Journal of Peace Research, 52, 536–50.Google Scholar
Pfister, C. (2007). Climatic extremes, recurrent crises and witch hunts: strategies of European societies in coping with exogenous shocks in the late sixteenth and early seventeenth centuries. Medieval History Journal, 10, 3373.Google Scholar
Pierrehumbert, R. T. (2010). Principles of Planetary Climate. Cambridge: Cambridge University Press.Google Scholar
Post, J. (1985). Food Shortage, Climatic Variability, and Epidemic Disease in Preindustrial Europe: The Mortality Peak in the Early 1740s. Ithaca, NY: Cornell University Press.Google Scholar
Raleigh, C., Linke, A., & O’Loughlin, J. (2014). Extreme temperatures and violence. Nature Climate Change, 4, 76–7.Google Scholar
Raspopov, O. M., Dergachev, V. A. & Esper, J., Kozyreva, O., Frank, D., Ogurtsov, M. G., Kolström, T., & Shao, X. (2008). The influence of the de Vries (~200-year) solar cycle on climate variations: results from the central Asian mountains and their global link. Palaeogeography, Palaeoclimatology, Palaeoecology, 259, 616.Google Scholar
Raspopov, O. M., Dergachev, V. A., & Kolström, T. (2004). Periodicity of climate conditions and solar variability derived from dendrochronological and other palaeo-climatic data in high latitudes. Palaeogeography, Palaeoclimatology, Palaeoecology, 209, 127–39.Google Scholar
Renssen, H., Seppä, H., Heiri, , Roche, O., Goosse, D. M., , H., & Fichefet, T. (2009). The spatial and temporal complexity of Holocene thermal maximum. Nature Geoscience, 2, 411–14.Google Scholar
Robock, A. (2000). Volcanic eruptions and climate. Reviews of Geophysics, 38, 191219.Google Scholar
Rohde, R., Muller, R. A. & Jacobsen, R., Muller, R., Perlmutter, S., Rosefelt, A. Wurtele, J., Groom, D., & Wickham, C. (2013). A new estimate of the average earth surface land temperature spanning 1753 to 2011. Geoinformatics and Geostatistics: An Overview, 1, 1. doi: 10.4172/gigs.1000101.Google Scholar
Rosen, A. M. (2007). Civilizing Climate: Social Responses to Climate Change in the Ancient Near East. Lanham, MD: Altamira Press.Google Scholar
Rotberg, R. & Rabb, T., eds. (1985). Hunger in History: The Impacts of Changing Food Production and Consumption Patterns on Society. Cambridge: Cambridge University Press.Google Scholar
Sandweiss, D. H., Maasch, K. A., & Anderson, D. G. (1999). Transitions in the mid-Holocene. Science, 283, 499500.Google Scholar
Scheffer, M., Carpenter, S., Foley, J. A., Folke, C., & Walker, B. (2001). Catastrophic shifts in ecosystems. Nature, 413, 591–6.Google Scholar
Scheffran, J., Brzoska, M., Brauch, H. G., Link, P. M., & Schilling, J. eds. (2012). Climate Change, Human Security and Violent Conflict: Challenges for Societal Stability. Berlin: Springer.Google Scholar
Schneider, T., Bischoff, T., & Haug, G. H. (2014). Migrations and dynamics of the intertropical convergence zone. Nature, 513, 4553.Google Scholar
Schönwiese, C. D. (1995). Klimaänderungen – Daten, Analysen, Prognosen. Berlin: Springer Verlag [in German].Google Scholar
Sigl, M., Winstrup, M. & McConnell, J. R., Welten, K. C., Plunkett, G., Ludlow, F., Büntgen, U., Caffee, , Chellman, M., Dahl-Jensen, N., Fischer, D., Kipfstuhl, H., Kostick, S., Maselli, C., Mekhaldi, O. J., Mulvaney, F., Muscheler, R., Pasteris, R., Pilcher, D. R., Salzer, J. R., Schüpbach, M., Steffensen, S., Vinther, J. P., , B. M., & Woodruff, T. E. (2015). Timing and climate forcing of volcanic eruptions for the past 2,500 years. Nature, 523, 543–9.Google Scholar
Sinha, A., Berkelhammer, M., Stott, L., Mudelsee, M., Cheng, H., & Biswas, J. (2011). The leading mode of Indian Summer Monsoon precipitation variability during the last millennium. Geophysical Research Letters, 38, L15703. doi: 10.1029/2011GL047713.Google Scholar
Slavin, P. (2016). Climate and famines: a historical reassessment. WIREs Climate Change, 7, 433–47.Google Scholar
Solomina, O. N., Bradley, R. S. & Jomelli, V., Geirsdottir, A., Kaufman, D. S., Koch, J., McKay, N. P., Masiokas, M., Miller, G., Nesje, A., Nicolussi, K., Owen, L. A., Putnam, A. E., Wanner, H., Wiles, G., & Yang, B. (2016). Glacier fluctuations during the past 2000 years. Quaternary Science Reviews, 149, 6190.Google Scholar
Sorokin, P. (1975). Hunger as a Factor in Human Affairs Gainesville: University Presses of Florida.Google Scholar
St. George, S. (2014). An overview of tree-ring width records across the Northern Hemisphere. Quaternary Science Reviews, 95, 132–50.Google Scholar
Steinhilber, F., Abreua, J. A. & Beer, J., Brunner, I., Christi, M., Fisher, H., Heikkilä, U., Kubik, , Mann, O. W., McCracken, M., Miller, K., Miyahara, H., Oerter, H., , H., & Wilhelms, F. (2012). 9,400years of cosmic radiation and solar activity from ice cores and tree rings. Proceedings of the National Academy of Sciences of the United States of America, 109, 5967–71.Google Scholar
Stothers, R. B. (1999). Volcanic dry fogs, climate cooling, and plague pandemics in Europe and the Middle East. Climatic Change, 42, 713–23.Google Scholar
Su, Y., Liu, L., Fang, X. Q., & Ma, Y. N. (2016). The relationship between climate change and wars waged between nomadic and farming groups from the Western Han Dynasty to the Tang Dynasty period. Climate of the Past, 12, 137–50.Google Scholar
Taricco, C., Mancuso, S., Ljungqvist, F. C., Alessio, S., & Ghil, M. (2015). Multispectral analysis of Northern Hemisphere temperature records over the last five millennia. Climate Dynamics, 45, 83104.Google Scholar
Tierney, J. E., Pausata, F. S. R., & deMenocal, P. B. (2017). Rainfall regimes of the Green Sahara. Science Advances, 3, e1601503. doi: 10.1126/sciadv.1601503.Google Scholar
Toohey, M., Krüger, K. & Sigl, , , M., Stordal, , , F., & Svensen, H. (2016). Climatic and societal impacts of a volcanic double event at the dawn of the Middle Ages. Climatic Change, 136, 401–12.Google Scholar
Tvauri, A. (2014). The impact of the climate catastrophe of 536–537 AD in Estonia and neighbouring areas. Estonian Journal of Archaeology, 18, 3056.Google Scholar
Wagner, G., Beer, J. & Masarik, J., Muscheler, R., Kubik, P. W., Mende, W., Laj, C., Raisbeck, G. M., & Yiou, F. (2001). Presence of the solar de Vries cycle (~205 years) during the last ice age. Geophysical Research Letters, 28, 303–6.Google Scholar
Walter, J. & Schofield, R., eds. (1989). Famine, Disease and the Social Order in Early Modern Society. Cambridge: Cambridge University Press.Google Scholar
Wanner, H., Beer, J. & Bütikofer, J., Crowley, T. J., Cubasch, U., Flückinger, J., Goosse, H., Grosjean, M., Joos, F., Kaplan, J. O., Küttel, M., Müller, S. A., Prentice, I. C., Solomina, O., Stocker, T. F., Tarasov, P., Wagner, M. & Widmann, M. (2008). Mid- to Late Holocene climate change: an overview. Quaternary Science Reviews, 27, 17911828.Google Scholar
Wanner, H., Solomina, O., Grosjean, M., Ritz, S., & Jetel, M. (2011). Structure and origin of Holocene cold events. Quaternary Science Reviews, 30, 3109–23.Google Scholar
Widgren, M. (2012). Climate and causation in the Swedish Iron Age: learning from the present to understand the past. Geografisk Tidskrift, 112, 126–34.Google Scholar
Willard, D. A., Bernhardt, C. E., Korejwo, D. A., & Meyers, S. R. (2005). Impact of millennial-scale Holocene climate variability on eastern North American terrestrial ecosystems: pollen-based climatic reconstruction. Global and Planetary Change, 47, 1735.Google Scholar
Wilson, R., Anchukaitis, K. & Briffa, K. R., Büntgen, U., Cook, , D’Arrigo, E., Davi, R., Esper, N., Frank, J., Gunnarsson, D., Hegerl, B., Helama, G., Klesse, S., Krusic, S., Linderholm, P. J., Myglan, H. W., Osborn, V., Rydval, T. J., Schneider, M., Schurer, L., Wiles, A., Zhang, G., , P., & Zorita, E. (2016). Last millennium northern hemisphere summer temperatures from tree rings: Part I: The long term context. Quaternary Science Reviews, 134, 118.Google Scholar
Yang, B., Kang, S., Ljungqvist, F. C., Zhao, Y., He, M., & Qin, C. (2014). Drought variability at the northern fringe of the Asian summer monsoon region over the past millennia. Climate Dynamics, 43, 845–59.Google Scholar
Zhang, D., Lee, H. F., Wang, C., Lie, B., Pei, Q., Zhang, J., & An, Y. (2011). The causality analysis of climate change and large-scale human crisis. Proceedings of the National Academy of Sciences of the United States of America, 108, 17296–301.Google Scholar
Zhang, D. D., Pei, Q. & Lee, H. F., Zhang, J., Chan, C. Q., Li, B., Li, J., & Zhang, X. (2015). The pulse of imperial China: a quantitative analysis of long-term geopolitical and climatic cycles. Global Ecology and Biogeography, 24, 187–96.Google Scholar
Zhang, D. D., Zhang, J., Lee, H. F., & He, Y. (2007). Climate change and war frequency in eastern China over the last millennium. Human Ecology, 35, 403–14.Google Scholar
Zhang, P., Cheng, H. & Edwards, E., Chen, F., Wang, Y., Yang, X., Liu, J., Tan, M., Wang, X., Liu, J., An, C., Dai, Z., Zhou, J., Zhang, D., Jia, J., Jin, L., & Johnson, K. R. (2008). A test of climate, sun, and culture. Science, 322, 940–2.Google Scholar
Zhang, Z., Tian, H. & Cazelles, B., Kausrud, K. L., Bräuning, A., Guo, , , F., & Stenseth, C. (2010). Periodic climate cooling enhanced natural disasters and wars in China during AD 10–1900. Proceedings of the Royal Society B: Biological Sciences, 277, 3745–53.Google Scholar

References

Agrawal, A. & Gibson, C. C. (1999). Enchantment and disenchantment: the role of community in natural resource conservation. World Development, 27 (4), 629–49.Google Scholar
Babai, D. & Molnár, Z. (2014). Small-scale traditional management of highly species-rich grasslands in the Carpathians. Agriculture, Ecosystem and Environment, 182, 123–30.Google Scholar
Berkes, F., Colding, J., & Folke, C. (2000). Rediscovery of traditional ecological knowledge as adaptive management. Ecological Applications, 10 (5), 1251–62.Google Scholar
Berkes, F. & Folke, C. (1998). Linking social and ecological systems for resilience and sustainability. In Berkes, F. & Folke, C., eds. Linking Social and Ecological Systems. Cambridge: Cambridge University Press, pp. 125.Google Scholar
Biró, E., Babai, D., Bódis, J., & Molnár, Z. (2014). Lack of knowledge or loss of knowledge? Traditional ecological knowledge of population dynamics of threatened plant species in East-Central Europe. Journal for Nature Conservation, 22, 318–25.Google Scholar
Dahlström, A., Iuga, , , A., & Lennartsson, T. (2013). Managing biodiversity rich hay meadows in the EU: a comparison of Swedish and Romanian grasslands. Environmental Conservation, 40 (2), 194205.Google Scholar
Dunn, E. C. (2003). Trojan pig: paradoxes of food safety regulation. Environment and Planning, 35 (8), 14931511.Google Scholar
Einarsson, P. (2015). Traditionell kunskap i modernt lantbruk. CBM:s skriftserie 91. Uppsala: Centrum för Biologisk Mångfald.Google Scholar
Eriksson, O., Bolmgren, K., Westin, A., & Lennartsson, T. (2015). Historic hay cutting dates from Sweden 1873–1951 and their implications for conservation management of species-rich meadows. Biological Conservation, 184, 100–7.Google Scholar
Gaignebet, C. (1990). La folle journée de Figaro: entre poils et jardinets. In Mesnil, M., ed. Les Plantes et les Saisons. Calendriers et Représentations. Collection Ethnologies d’Europe. Bruxelles: Institut de Sociologie.Google Scholar
Gherman, T. (2002). Meteorologia populară. Observări, credințe și obiceiuri. Bucarest: Paideia.Google Scholar
Glassie, H. (1995). Tradition. The Journal of American Folklore, 108 (430), 395412.Google Scholar
Gómez-Baggethun, E., Reyes-Garcia, V., Olsson, P., & Montes, C. (2012). Traditional ecological knowledge and community resilience to environmental extremes: a case study in Doñana, SW Spain. Global Environmental Change, 22, 640–50.Google Scholar
Grigg, D. B. (1974). The Agricultural Systems of the World: An Evolutionary Approach. Cambridge: Cambridge University Press.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. InTech.Google Scholar
Handler, R. & Linnekin, J. (1984). Tradition, genuine or spurious. The Journal of American Folklore, 97 (358), 273–90.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: Presa Universitara Clujeana, pp. 367–70.Google Scholar
Huband, S. & McCracken, D. (2011). Understanding high nature value agriculture in the Romanian Carpathians: a case study. In Knowles, B., ed. Mountain Hay Meadows: Hotspots of Biodiversity and Traditional Culture. London: Society of Biology.Google Scholar
Iancu, B. & Stroe, M. (2016). In search of eligibility: Common Agricultural Policy and the reconfiguration of hay-meadows management in the Romanian highlands. MARTOR: The Museum of the Romanian Peasant Anthropology Review, 21, 129–44.Google Scholar
Iuga, A. (2016). Intangible hay heritage in Șurdești. MARTOR: The Museum of the Romanian Peasant Anthropology Review, 21, 6784.Google Scholar
Kideckel, D. (2001). Labor and Society in the Jiu Valley and Fagaras Regions of Romania. Washington, DC: National Council for Eurasian and East European Research.Google Scholar
Kideckel, D. (2006). Colectivism și singurătate în satele românești: Țara Oltului în perioada comunistăși în primii ani după Revoluție. Iași: Polirom.Google Scholar
Kligman, G. & Verdery, K. 2011. Peasants under Siege: The Collectivization of Romanian Agriculture, 1949–1962. Princeton: Princeton University Press.Google Scholar
Larsson, J. (2009). Fäbodväsendet 1550–1920: ett centralt element i Nordsveriges jordbrukssystem. Uppsala: Swedish University of Agricultural Sciences.Google Scholar
Lave, J. & Wenger, E. (1991). Situated Learning. Legitimate Peripheral Participation. Cambridge: Cambridge University Press.Google Scholar
Lennartsson, T. & Helldin, J.-O. (2007). Agricultural landscapes in Eastern Europe as reference areas for Swedish land management. In Hasund, K. P. & Helldin, J.-O., eds. Valuable Agricultural Landscapes – the Importance of Romania and Scandinavia for Europe, KSLA 2007: 5. Stockholm: Kungl. Skogs- och Lantbruksakademien.Google Scholar
Magnusson, L. & Ottosson, L. eds. (2009). The Evolution of Path Dependence. Cheltenham: Edward Elgar.Google Scholar
Manoliu, V. (1999). Mic dicționar de astrologie și meteorology țărănească. Bucarest: Mentor.Google Scholar
Mesnil, M. (1997). Etnologul între șarpe și balaur; Mesnil, M. and Popova, A. Eseuri de mitologie balcanică. Bucharest: Paideia.Google Scholar
Mihăilescu, V. (2000). La maisnie diffuse, du communisme au capitalisme: Questions et hypothèses, Balkanologie, 4:2. www.balkanologie.revues.org/334Google Scholar
Mihăilescu, V., Nicolau, , Gheorghiu, V., , M., & Dirnovan, G. (1993). Snagov – trei proiecții asupra sistematizării. Sociologie Românească, 4 (1), 1832.Google Scholar
Myrdal, J. (1999). Jordbruket under feodalismen. 1000–1700. Stockholm: Natur och kultur/ Lts förlag.Google Scholar
Myrdal, J. (2012). Boskapsskötseln under medeltiden. En källpluralistisk studie. Stockholm: Nordiska Museets förlag.Google Scholar
Myrdal, J. & Morell, M., eds. (2011). The Agrarian History of Sweden. From 4000 BC to AD 2000. Lund: Nordic Academic Press.Google Scholar
Nicolau, I. (1998). Ghidul sărbătorilor românești. Bucharest: Humanitas.Google Scholar
O’Rourke, E. (1999). Changing identities, changing landscapes: human-land relations in transition in the Aspre, Roussillon. Cultural Geographies, 6 (1), 29.Google Scholar
Österling, P.-A. (2010). Dialekt och folkminnesarkivens material – etnologi på Institutet för språk och folkminnen (SOFI) exemplet ULMA. In Tunón, H. and Dahlström, A., eds., Nycklar till kunskap. Om människans bruk av naturen. Uppsala: Centrum för Biologisk Mångfald and Stockholm: Kungl. skogs- och lantbruksakademien, pp. 5966.Google Scholar
Oteros-Rozas, E., Ontillera-Sánchez, R., Sanosa, P., Gómez-Baggethun, E., Reyes-García, V., & González, J. A. (2013). Traditional ecological knowledge among transhumant pastoralists in Mediterranean Spain. Ecology and Society, 18 (3), 33.Google Scholar
Persson, J. & Nilsson, N. Ö. (1996). Lien och dess marker. Stockholm: Lts förlag.Google Scholar
Ruiz-Mallén, I. & Corbera, E. (2013). Community-based conservation and traditional ecological knowledge: Implications for social-ecological resilience. Ecology and Society, 18(4), 12. http://dx.doi.org/10.5751/ES-05867-180412Google Scholar
Schön, E. (2005). Folktrons ABC. Stockholm: Carlsson.Google Scholar
Stahl, H. H. (1946). Sociologia satului devălmaș românesc. Organizarea economică și juridică a trupurilor de moșie. Bucarest: Fundația Regele Mihai I.Google Scholar
Stroe, M. (2015). Rigid norms, flexible subjects: socio-ecological resilience in contested hay landscapes. In Troc, G. & Iancu, B., eds., Modes of Appropriation and Social Resistance, Bucharest: Tritonic, pp. 103–20.Google Scholar
Szelenyi, I., ed. (1998). Privatizing the Land. Rural Political Economy in Post-Communist and Socialist Societies. London: Routledge.Google Scholar
Torsello, D. (2003). Trust, Property and Social Change in a Southern Slovakian Village. Munster: Lit. Verlag.Google Scholar
Tunón, H. (2015). Studiet av bruket av naturen. Ur etnobiologins historia. In Tunón, H., ed., Etnobotanik. Planter i skikog brug, i historien og i folkemedicinen. Vagn J. Brøndegaards biografi, bibliografi og artikler på dansk i udvalg. Vol. 1. Stockholm: Kungl. Skogs- och Lantbruksakademien and Uppsala: Centrum för Biologisk Mångfald, pp. 3346.Google Scholar
Tunón, H., Kvarnström, M., Axelsson Linkowski, W., & Westin, A. (2014). Hur bör Sverige genomföra artiklarna 8j and 10c i syfte att uppnå Aichi-mål 18 i FN:s Konvention om biologisk mångfald?. Uppsala: Centrum för Biologisk Mångfald.Google Scholar
Verdery, K. (2003). The Vanishing Hectare: Property and Value in Postsocialist Transylvania. London: Cornell University PressGoogle Scholar
Wästfelt, A., Saltzman, K., Gräslund Berg, E., & Dahlberg, A. (2012). Landscape care paradoxes: Swedish landscape care arrangements in a European context. Geoforum, 43, 1171–81.Google Scholar
Westin, A., Isacson, M., & Lennartsson, T. (2017). Land and labour as resources of an integrated peasant economy in a Swedish mining district during the 1860s great famine. In Panjek, A. Larsson, J. and Mocarelli. L, eds., Integrated Peasant Economy in a Comparative Perspective – Alps, Scandinavia and Beyond, Koper: University of Primorska Press, pp. 309–32.Google Scholar

References

Al-Abdulrazzak, D., Naidoo, R., Palomares, M. L. D., & Pauly, D. (2012). Gaining perspective on what we’ve lost: the reliability of encoded anecdotes in historical ecology. PLoS ONE 7(8), e43386.Google Scholar
Atran, S., Medin, D., & Noss, N. (2004). Evolution and devolution of knowledge: a tale of two biologies. Journal of the Royal Anthropological Institute, 10(2), 395420.Google Scholar
Balint, P. J., Stewart, R. E., Desai, A., & Walters, L. C. (2011). Wicked Environmental Problems. Managing Uncertainty and Conflict. Washington, DC: Island Press.Google Scholar
Barthel, S., Folke, C., & Colding, J. (2010). Social-ecological memory in urban gardens – retaining the capacity for management of ecosystem services. Global Environmental Change, 20 (2010), 255–65.Google Scholar
Belvedresi, R. E. (2014). Book review symposium. Review of the book The Collective Memory Reader. Memory Studies, 7(1), 108–31.Google Scholar
Berger, P. & Luckmann, T. (1966). The Social Construction of Reality. New York: Penguin Books.Google Scholar
Bishop, K. H., Beven, K., Destouni, G., Abrahamsson, K., Andersson, L., Johnson, R. K., Rohde, J., & Hjerdt, N. (2009). Nature as the ‘natural’ goal for water management: a conversation. Ambio, 38(4), 209–14.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
Brinck, E. & Frost, C. (2009). Evaluation of amendments used to prevent sodification of irrigated fields. Applied Geochemistry, 24(11), 2113–22.Google Scholar
Brown, C. J. & Trebilco, R. (2014). Unintended cultivation, shifting baselines, and conflict between objectives for fisheries and conservation. Conservation Biology, 28(3), 677–88.Google Scholar
Bull, J. W., Gordon, A., Law, E. A., Suttle, K. B., & Milner-Gulland, E. J. (2014). Importance of baseline specification in evaluating conservation interventions and achieving no net loss of biodiversity. Conservation Biology, 28(3), 799809.Google Scholar
Bull, J. W., Gordon, A., Watson, J. E. M., & Maron, M. (2016). Seeking convergence on the key concepts in ‘no net loss’ policy. Journal of Applied Ecology, 53(6), 1686–93.Google Scholar
Burr, V. (1995). An Introduction to Social Constructionism. London: Routledge.Google Scholar
Campbell, L. M., Gray, N. J., Hazen, E. L., & Shackeroff, J. M. (2009). Beyond baselines: rethinking priorities for ocean conservation. Ecology and Society, 14(1), 14.Google Scholar
Cardinale, B. J., Duffy, J. E., Gonzalez, A., Hooper, D. U., Perrings, C., Venail, P., Narwani, A., Mace, G. M., Tilman, D., Wardle, D. A., Kinzig, A. P., Daily, G. C., Loreau, M., Grace, J. B., Larigauderie, A., Srivastava, D. S., & Naeem, S. (2012). Biodiversity loss and its impact on humanity. Nature, 486, 5967.Google Scholar
CBD. (2007). CBD Guidelines on Biodiversity and Tourism Development. Secretariate of the Convention on Biological Diversity, February (2007). CBD, UNEP.Google Scholar
Clavero, M. (2014). Shifting baselines and the conservation of non-native species. Conservation Biology, 28(5), 1434–6.Google Scholar
Cook, B. & Kothari, U. (2001). Participation: The New Tyranny? London: Zed Books.Google Scholar
Comer, P. J. (1997). Letter. Conservation Biology, 11(2), 301–3.Google Scholar
Corlett, R. T. (2015). The Anthropocene concept in ecology and conservation. Trends in Ecology and Evolution, 30(1), 3641.Google Scholar
Costa, P. M., Stuart, M., Pinard, M., & Phillips, G. (2000). Elements of a certification system for forestry-based carbon offset projects. Mitigation and Adaptation Strategies for Global Change, 5(1), 3950.Google Scholar
Dallimer, M., Tinch, D., Acs, S., Hanley, N., Southall, H. R., Gaston, K. J., & Armsworth, P. R. (2009). 100 years of change: examining agricultural trends, habitat change and stakeholder perceptions through the 20th century. Journal of Applied Ecology, 46(2), 334–43.Google Scholar
Daw, T. M. (2010). Shifting baselines and memory illusions: what should we worry about when inferring trends from resource user interviews? Animal Conservation, 13(6), 534–5.Google Scholar
Denevan, W. M. (1992). The pristine myth: the landscape of the Americas in 1492. Annals of the Association of American Geographers, 82(3), 369–85.Google Scholar
De Vries, D. (2005). Choosing your baseline carefully: integrating historical and political ecology in the evaluation of environmental intervention projects. Journal of Ecological Anthropology, 9(1), 3550.Google Scholar
Downs, P., Singer, M., Orr, B., Diggory, Z., & Church, T. (2011). Restoring ecological integrity in highly regulated rivers: the role of baseline data and analytical references. Environmental Management, 48(4), 847–64.Google Scholar
Drew, J., Philipp, C., & Westneat, M. W. (2013). Shark tooth weapons from the 19th century reflect shifting baselines in the Central Pacific predator assemblies. PLoS ONE, 8(4), e59855.Google Scholar
EU Directive (2000). Directive 2000/60/EC of the European Parliament and of the Council of 23 October 2000 Establishing a Framework for Community Action in the Field of Water Policy. Official Journal of the European Communities. L 327, 173.Google Scholar
Ellis, E. C. (2011). Anthropogenic transformation of the terrestrial biosphere. Philosophical Transactions of the Royal Society A, 369(1938), 1010–35.Google Scholar
Escobar, A. (1999). After nature: steps to an antiessentialist political ecology. Current Anthropology, 40(1), 130.Google Scholar
Ferraro, P. J. & Pattanayak, S. K. (2006). Money for nothing? A call for empirical evaluation of biodiversity conservation investments. PLoS Biology, 4(4), 482–8.Google Scholar
Foster, N. L., Foggo, A., & Howell, K. L. (2013). Using species-area relationships to inform baseline conservation targets for the deep north east Atlantic. PLoS ONE, 8(3), e58941.Google Scholar
Gavriely-Nuri, D. (2014). Collective memory as a metaphor: the case of speeches by Israeli prime ministers 2001–2009. Memory Studies, 7(1), 4660.Google Scholar
Gergen, K. J. (2001), Social Construction in Context. Sage Publications.Google Scholar
Golding, L. A., Timperley, M. H., & Evans, C. W. (1997). Non-lethal responses of the freshwater snail Potamopyrgus antipodarum to dissolved arsenic. Environmental Monitoring and Assessment, 47(3), 239–54.Google Scholar
Gómez-Pompa, A. & Kaus, A. (1992). Taming the wilderness myth. BioScience, 42(4), 271379.Google Scholar
Gray, T. N. E., Phan, C., Pin, C., & Prum, S. (2012). Establishing a monitoring baseline for threatened large ungulates in eastern Cambodia. Wildlife Biology, 18(4), 406–13.Google Scholar
Haila, Y. (1997). A ‘natural’ benchmark for ecosystem function. Conservation Biology, 11(2), 300–1.Google Scholar
Haila, Y. (2000). Beyond the nature-culture dualism. Biology and Philosophy, 15(2), 155–75.Google Scholar
Halbwachs, M. (1992). On Collective Memory. Transl./ed. Coser, L. A.. Chicago: University of Chicago Press.Google Scholar
Healy, C. & Tumarkin, M. (2011). Social memory and historical justice. Introduction Memory Studies, 4(1), 312.Google Scholar
Hoffmann, B. D., Auina, S., & Stanley, M. C. (2014). Targeted research to improve species management: Yellow Crazy ant Anoplolepis gracilipes in Samoa. PLoS ONE, 9(4), e95301.Google Scholar
Hooper, D. U., Carol Adair, E., Cardinale, B. J., Byrnes, J. E. K., Hungate, B. A., Matulich, K. L., Gonzalez, A., Duffy, J. E., Gamfeldt, L., & O’Connor, M. I. (2012). A global synthesis reveals biodiversity loss as a major driver of ecosystem change. Nature, 486, 105–9.Google Scholar
Hornborg, A. & Pálsson, G., eds. (2000). Negotiating Nature: Culture, Power, and Environmental Argument. Lund, Sweden: Lund Studies in Human Ecology, Lund University Press.Google Scholar
Hunter, M. (1996). Benchmarks for managing ecosystems: are human activities natural? Conservation Biology, 10(3), 695–7.Google Scholar
IAEA. (1999). Maintenance of Records for Radioactive Waste Disposal. Report IAEA-TECDOC-1097. International Atomic Energy Agency.Google Scholar
IAEA. (2001). Waste Inventory Record Keeping Systems (WIRKS) for the Management and Disposal of Radioactive Waste. Report IAEA-TECDOC-1222. International Atomic Energy Agency.Google Scholar
Jackson, J. B. C. & Alexander, K. E. (2011). Introduction: the importance of shifting baselines. In Jackson, J. B. C., Alexander, K. E., & Sala, E., eds., Shifting Baselines. The Past and the Future of Ocean Fisheries. Washington, DC: Island Press, pp. 17.Google Scholar
Jackson, J. B. C., Alexander, K. E., & Sala, E. eds. (2011). Shifting Baselines. The Past and the Future of Ocean Fisheries. Washington, DC: Island Press.Google Scholar
Josefsson, T. (2009). Pristine Forest Landscapes as Ecological References. Human Land Use and Ecosystem Change in Boreal Fennoscandia. Uppsala: Swedish University of Agricultural Sciences.Google Scholar
Josefsson, T., Bergman, I., & Östlund, L. (2010). Quantifying Sami settlement and movement patterns in northern Sweden 1700–1900. Arctic, 63(2), 141–54.Google Scholar
Josefsson, T., Hörnberg, G., & Östlund, L. (2009). Long-term human impact and vegetation changes in a boreal forest reserve: implications for the use of protected areas as ecological references. Ecosystems, 12(6), 1017–36.Google Scholar
Kai, Z., Woan, T. S., Jie, L., Goodale, E., Kitajima, K., Bagchi, R., & Harrison, R. D. (2014). Shifting baselines on a tropical forest frontier: extirpations drive declines in local ecological knowledge. PLoS ONE, 9(1), e86598.Google Scholar
Kenny, M. G. (1999). A place for memory: the interface between individual and collective history. Comparative Studies in Society and History, 41(3), 420–37.Google Scholar
Knowlton, N. & Jackson, J. B. C. (2008). Shifting baselines, local impacts, and global change on coral reefs. PLoS Biology, 6(2), e54.Google Scholar
Latour, B. (2004). Politics of Nature. How to Bring the Sciences into Democracy. Cambridge, MA: Harvard University Press.Google Scholar
Lélé, S. & Norgaard, R. B. (1996). Sustainability and the scientist’s burden. Conservation Biology, 10(2), 354–65.Google Scholar
Lewis, S. L. & Maslin, M. A. (2015). Defining the Anthropocene. Nature, 519, 171–80.Google Scholar
Linde, C. (1997). Narrative: experience, memory, folklore. Journal of Narrative and Life History, 7(1), 281–90.Google Scholar
Linde, C. (2000). The acquisition of a speaker by a story: how history becomes memory and identity. Ethnos, 28(4), 608–32.Google Scholar
Lorimer, J., Sandom, C., Jepson, P., Doughty, C., Barua, M., & Kirby, K. J. (2015). Rewilding: science, practice, and politics. Annual Review of Environment and Resources, 40(1), 3962.Google Scholar
Maron, M., Bull, J. W., Evans, M. C., & Gordon, A. (2015). Locking in loss: baselines of decline in Australian biodiversity offset policies. Biological Conservation, 192, 504–12.Google Scholar
Maron, M., Hobbs, R. J., Moilanen, A., Matthews, J. W., Christie, K., Gardner, T. A., Keith, D. A., Lindenmayer, D. B., & McAlpine, C. A. (2012). Faustian bargains? Restoration realities in the context of biodiversity offset policies. Biological Conservation, 155 (2012), 141–8.Google Scholar
McClenachan, L., Feretti, F., & Baum, J.K. (2012). From archives to conservation: why historical data are needed to set baselines for marine animals and ecosystems. Conservation Letters, 5(5), 349–59.Google Scholar
McDonald-Madden, E., Gordon, A., Wintle, B. A., Walker, S., Grantham, H., Carvalho, S., Bottrill, M., Joseph, L., Ponce, R., Stewart, R., & Possingham, H. P. (2009). ‘True’ conservation progress. Science, 323(5910), 43–4.Google Scholar
Michelutti, N., McCleary, K. M., Antoniades, D., Sutherland, P., Blais, J. M., Douglas, M. S. V., & Smol, J. P. (2013). Using paleolimnology to track the impacts of early Arctic peoples on freshwater ecosystems from southern Baffin Island, Nunavut. Quaternary Science Reviews, 76, 8295.Google Scholar
Natlandsmyr, B. & Hjelle, K. L. (2016). Long-term vegetation dynamics and land-use history: providing a baseline for conservation strategies in protected Alnus glutinosa swamp woodlands. Forest Ecology and Management, 372, 7892.Google Scholar
Nora, P. (1989). Between memory and history: les lieux de memoire. Representations, 26, 724.Google Scholar
Nuttall, M. (1992). Arctic Homeland: Kinship, Community and Development in Northwest Greenland. London: Belhaven.Google Scholar
Nykvist, B. & von Heland, J. (2014). Social-ecological memory as a source of general and specified resilience. Ecology and Society, 19(2), 47.Google Scholar
Olick, J. K. & Robbins, J. (1998). Social memory studies: from ‘collective memory’ to the historical sociology of mnemonic practices. Annual Review of Sociology, 24, 105–40.Google Scholar
Olick, J. K., Vinitzky-Seroussi, V., & Levy, D. (2014). Response to our critics. Memory Studies, 7(1), 131–38.Google Scholar
Ortmann, G. (2010). On drifting rules and standards. Scandinavian Journal of Management, 26(2), 204–14.Google Scholar
Papworth, S. K., Rist, J., Coad, L., & Milner-Gulland, E. J. (2009). Evidence for shifting baseline syndrome in conservation. Conservation Letters, 2(2), 93100.Google Scholar
Pauly, D. (1995). Anecdotes and the shifting baseline syndrome of fisheries. TREE, 10(10), 430.Google Scholar
Rautio, A.-M., Josefsson, T., & Östlund, L. (2014). Sami resource utilization and site selection: historical harvesting of inner bark in northern Sweden. Human Ecology, 42(1), 137–46.Google Scholar
Reed, M. S. (2008). Stakeholder participation for environmental management: a literature review. Biological Conservation, 141(10), 2417–31.Google Scholar
Rice, J. (2013). Further beyond the Durkheimian problematic: environmental sociology and the co-construction of the social and the natural. Sociological Forum, 28(2), 236–60.Google Scholar
Roediger, H. L., Marsh, E. J., & Lee, S. C. (2002) Varieties of memory. In Medin, D. L & Pashler, H., eds., Stevens Handbook of Experimental Psychology: Memory and Cognitive Processes, vol. 2, 3rd ed. New York: John Wiley and Sons, pp. 141.Google Scholar
Rothberg, M. (2009). Multidirectional Memory: Remembering the Holocaust in the Age of Decolonisation. Stanford, CA: Stanford University Press.Google Scholar
Ryan, J. C. (2013). Botanical memory: exploring emotional recollections of native flora in the southwest of western Australia. Emotion, Space and Society, 8, 2738.Google Scholar
Sáenz-Arroyo, A., Robert, , Torre, C. M., Carino-Olvera, J., , M., & Enriquez-Andrade, R. R. (2005). Rapidly shifting environmental baselines among fishers of the Gulf of California. Proceedings of the Royal Society B, 272(1575), 1957–62.Google Scholar
Setten, G. & Austrheim, G. (2012). Changes in land use and landscape dynamics in mountains of northern Europe: challenges for science, management and conservation. International Journal of Biodiversity Science, Ecosystem Services and Management, 8(4), 287–91.Google Scholar
Setten, G., Stenseke, M., & Moen, J. (2012). Ecosystem services and landscape management: three challenges and one plea. International Journal of Biodiversity Science, Ecosystem Services and Management, 8(4), 305–12.Google Scholar
Sörlin, S. (1991). Humanekologi: vägar till överlevnad. Umeå: Forum för tvärvetenskap, Umeå universitet.Google Scholar
SOU 2007: 38. Kunskapsläget på kärnavfallsområdet 2007. Nu levandes ansvar, framtida generationers frihet. Stockholm: Statens råd för kärnavfallsfrågorGoogle Scholar
SOU 2015: 11. Kunskapsläget på kärnavfallsområdet 2015. Kontroll, dokumentation och finansiering för ökad säkerhet. Stockholm: Kärnavfallsrådet.Google Scholar
Terry, J. (2013). ‘When the sea of living memory has receded’: cultural memory and literary narratives of the Middle Passage. Memory Studies, 6(4), 474–88.Google Scholar
Ullberg, S. (2013). La Inundación – katastrofer och minnets politik i Argentina. In Viktorin, M. & Widmarrd, C., eds., Antropologi och tid. Stockholm: Svenska sällskapet för Antropologi och Geografi, pp. 175–92.Google Scholar
Valinia, S., Hansen, H.-P., Futter, M. N., Sriskandarajah, N., & Fölster, J. (2012). Problems with the reconciliation of good ecological status and public participation in the Water Framework Directive. Science of the Total Environment, 433(2012), 482–90.Google Scholar
Wilson, R. A. (2005). Collective memory, group minds, and the extended mind thesis. Cognitive Processing, 6, 227–36.Google Scholar
Zelizer, B. (1995). Reading the past against the grain: the shape of memory studies. Critical Studies in Mass Communication, 12(2), 2014–239.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
×