Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-30T02:52:44.575Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 April 2016

Christiane Runyan
Affiliation:
The Johns Hopkins University
Paolo D'Odorico
Affiliation:
University of Virginia
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
Global Deforestation , pp. 195 - 248
Publisher: Cambridge University Press
Print publication year: 2016

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

Aber, J. D., Botkin, D. B., and Melillo, J. M. (1978). Predicting the effects of different harvesting regimes on forest floor dynamics in northern hardwoods. Canadian Journal of Forest Research, 8(3), 306–315.Google Scholar
Aber, J. D., and Melillo, J. M. (1991). Terrestrial Ecosystems. Saunders College Publishing, Orlando, FL.
Aber, J. D., Nadelhoffer, K. J., Steudler, P., and Melillo, J. M. (1989). Nitrogen saturation in northern forest ecosystems. BioScience, 39(6), 378–286.Google Scholar
Aber, J. D., Ollinger, S. V., Driscoll, C. T., Likens, G. E., Holmes, R. T., Freuder, R. J., and Goodale, C. L. (2002). Inorganic nitrogen losses from a forested ecosystem in response to physical, chemical, biotic, and climatic perturbations. Ecosystems, 5(7), 0648–0658.Google Scholar
Achard, F., Eva, H. D., Mayaux, P., Stibig, H. J., and Belward, A. (2004). Improved estimates of net carbon emissions from land cover change in the tropics for the 1990s. Global Biogeochemical Cycles, 18(2).Google Scholar
Aerts, R. (1997). Climate, leaf litter chemistry and leaf litter decomposition in terrestrial ecosystems: a triangular relationship, Oikos, 79, 439–449.Google Scholar
Aerts, R., Wallen, B., and Malmer, N. (1992). Growth-limiting nutrients in Sphagnum- dominated bogs subject to low and high atmospheric nitrogen supply. Journal of Ecology, 80, 131–140.Google Scholar
Agrawal, A., and Angelsen, A. (2009). Using community forest management to achieve REDD+ goals. In A. Angelsen (ed.) Realising REDD+: National Strategy and Policy Options, 201–212.
Agrawal, A., Cashore, B., Hardin, R., Shepherd, G., Benson, C., and Miller, D. (2013). Economic contributions of forests. In Background paper prepared for the United Nations Forum on Forests, Available at http://www. un.org/esa/forests/pdf/session_documents/unff10/EcoContrForests.pdf (accessed August 15, 2013).
Agrawal, A., Nepstad, D., and Chhatre, A. (2011). Reducing emissions from deforestation and forest degradation. Annual Review of Environment and Resources, 36, 373–396.Google Scholar
Agarwal, S. K. (2008). Fundamentals of Ecology. APH Publishing Corporation, New Delhi.
Agestam, E., Ekö, P. M., Nilsson, U., and Welander, N. T. (2003). The effects of shelterwood density and site preparation on natural regeneration of Fagus sylvatica in southern Sweden. Forest Ecology and Management, 176, 61–73.Google Scholar
Alcantara-Ayala, I., Esteban-Chavez, O., and Parrot, J. F. (2006). Landsliding related to land-cover change: A diachronic analysis of hillslope instability distribution in the Sierra Norte, Puebla, Mexico, Catena, 65(2), 152–165.
Aldrich, M., Billington, C., Edwards, M., and Laidlaw, R. (1997). Tropical montane cloud forests: An urgent priority for conservation (No. 2). Cambridge, UK, World Conservation Monitoring Centre.
Alexandratos, N., and Bruinsma, J. (2012). World agriculture towards 2030/2050: The 2012 revision. ESA Working paper No. 12-03. Rome, FAO.
Alkemade, R., van Oorschot, M., Miles, L., Nellemann, C., Bakkenes, M., and Ten Brink, B. (2009). GLOBIO3: A framework to investigate options for reducing global terrestrial biodiversity loss. Ecosystems, 12(3), 374–390.Google Scholar
Allison, G. B., Cook, P. G., Barnett, S. R., Walker, G. R., Jolly, I., and Hughes, M. (1990). Land clearance and river salinization in the Western Murray Basin, Australia. J. Hydrol., 119, 1–20.Google Scholar
Amiro, B. D. et al. (2006). Carbon, energy and water fluxes at mature and disturbed forest sites, Saskatchewan, Canada. Agricultural and Forest Meteorology, 136(3), 237–251.Google Scholar
Anderies, J. M. (2005). Minimal models and agroecological policy at the regional scale: An application to salinity problems in southeastern Australia, Reg. Environ.Change, 5, 1–17.Google Scholar
Anderies, J. M., Janssen, M. A., and Walker, B. H. (2002). Grazing management, resilience, and the dynamics of a fire-driven rangeland system. Ecosystems, 5, 23–44.Google Scholar
Andréassian, V. (2004). Waters and forests: From historical controversy to scientific debate. Journal of hydrology, 291(1), 1–27.Google Scholar
Angelsen, A. (1999). Agricultural expansion and deforestation: Modelling the impact of population, market forces and property rights. Journal of Development Economics, 58(1), 185–218.Google Scholar
Angelsen, A. (2010). Policies for reduced deforestation and their impact on agricultural production. Proceedings of the National Academy of Sciences, 107(46), 19639–19644.Google Scholar
Angelsen, A., and Kaimowitz, D. (1999). Rethinking the causes of deforestation: Lessons from economic models. The World Bank Research Observer, 14(1), 73–98.Google Scholar
Angelsen, A., and Kaimowitz, D. (2001). Introduction: The role of agricultural technologies in tropical deforestation. In Angelsen, A. and , D. Kaimowitz (eds.), Agricultural Technologies and Tropical Deforestation. CABI. Wallingford.
Angers, D. A., and Caron, J.. (1998). Plant-induced changes in soil structure: Processes and feedbacks. Biogeochemistry, 42(1/2), 55–72.Google Scholar
Anseeuw, W., Boche, M., Breu, T., Giger, M., Lay, J., Messerli, P., and Nolte, . (2012). Transnational land deals for agriculture in the global south K., available at http://www.landcoalition.org/fr/publications/transnationalland-deals-agriculture-global-south.
Aoki, M. (2001). Toward a Comparative Institutional Analysis. MIT Press, Cambridge, MA.
Armstrong, R. L. et al. (2001). State of the cryosphere: Response of the cryosphere to global warming. Glaciological Data, 18.
Ashraf, A., Maah, M. J., and Yusoff, I. (2011). Introduction to remote sensing of biomass, biomass and remote sensing of biomass, Dr. Islam Atazadeh (ed.), InTech, doi:10.5772/16462 available at: http://www.intechopen.com/books/biomass-and-remote-sensing-of-biomass/introduction-to-remote-sensing-of-biomass.
Aslam, M., and Prathapar, S. A. (2006). Strategies to mitigate secondary salinization in the Indus Basin of Pakistan: A selective review. Research report 97. International Water Management Institute (IWMI), Colombo, Sri Lanka.
Asner, G. P. (2001). Cloud cover in Landsat observations of the Brazilian Amazon. International Journal of Remote Sensing, 22(18), 3855–3862.Google Scholar
Attiwill, P. M., and Adams, M. A. (1993). Nutrient cycling in forests. New Phytologist, 124(4), 561–582.Google Scholar
Australian Bureau of Rural Sciences. (2010). Australia's Forests at a Glance 2010. Canberra, Australia, Australian Government Department of Agriculture, Fisheries and Forestry.
Avissar, R., and Liu, Y. (1996). Three-dimensional numerical study of shallow convective clouds and precipitation induced by land surface forcing. Journal of Geophysical Research, 101(D3), 7499–518.Google Scholar
Avissar, R., Silva Dias, P. L., Silva Dias, M. A. F., and Nobre, C. (2002). The large-scale biosphere-atmosphere experiment in Amazonia (LBA): Insights and future research needs. Journal of Geophysical Research, 107(D20), 8086, doi:10.1029/2002JD002704.Google Scholar
Baath, E., and Soderstrom, B. (1979). Fungal biomass and fungal immobilization of plant nutrients in Swedish coniferous forest soils. Revue d'Ecologie et de Biologie du Sol 16, 477–489.Google Scholar
Bader, M., Rietkerk, M., and Bregt, A. (2007). Vegetation structure and temperature regimes of tropical alpine treelines. Arctic, Antarctic, and Alpine Research, 39, 353–364.Google Scholar
Baidya Roy, S. (2009). Mesoscale vegetation-atmosphere feedbacks in Amazonia. Journal of Geophysical Research, 114, D20111, doi:10.1029/2009JD012001.Google Scholar
Baidya Roy, S., and Avissar, R. (2002). Impact of land use/land cover change on regional hydrometeorology in Amazonia. J. Geophys. Res., 107, 8037, doi:10.1029/2000JD000266.Google Scholar
Balat, M. (2007). Global bio-fuel processing and production trends. Energy Explor Exploit, 25, 195–218.Google Scholar
Balat, M., and Balat, H. (2009). Recent trends in global production and utilization of bio-ethanol fuel. Applied Energy, 86(11), 2273–2282.Google Scholar
Balch, J. K., Nepstad, D. C., and Curran, L. M. (2009). Pattern and process: Fire-initiated grass invasion at Amazon transitional forest edges. In Fire Ecology of Tropical Ecosystems, (M. Cochrane, ed.), 481-502.
Balesdent, J., Manotti, A., and Guillet, B. (1987). Natural I3C abundance as a tracer for studies of soil organic matter dynamics. Soil Biol. Biochem. 19, 25–30.Google Scholar
Ball, M. C., Hodges, V. S., and Laughlin, G. P. (1991). Cold-induced photoinhibition limits regeneration of snow gum at tree-Line. Functional Ecology, 5, 663–668.Google Scholar
Balmford, A. et al. (2002). Economic reasons for conserving wild nature. Science, 297(5583), 950–953.Google Scholar
Balmford, A., Green, R. E., and Jenkins, M. (2003). Measuring the changing state of nature. Trends in Ecology and Evolution, 18(7), 326–330.Google Scholar
Balvanera, P., Kremen, C., and Martínez-Ramos, M. (2005). Applying community structure analysis to ecosystem function: Examples from pollination and carbon storage. Ecological Applications, 15(1), 360–375.Google Scholar
Barbier, E. B. (2004). Agricultural expansion, resource booms and growth in Latin America: Implications for long-run economic development. World Dev., 32, 137–57.Google Scholar
Barbier, E. B., and Burgess, J. C. (2001). Tropical deforestation, tenure insecurity, and unsustainability. Forest Science, 47(4), 497–509.Google Scholar
Barbier, E. B., Burgess, J. C., and Grainger, A. (2010). The forest transition: Towards a more comprehensive theoretical framework. Land Use Policy, 27(2), 98–107.Google Scholar
Barg, A. K., and Edmonds, R. L. (1999). Influence of partial cutting on site microclimate, soil nitrogen dynamics, and microbial biomass in Douglas-fir stands in western Washington. Canadian Journal of Forest Research, 29, 705–713.Google Scholar
Barnosky, A. D. et al. (2011). Has the Earth/'s sixth mass extinction already arrived?Nature, 471(7336), 51–57.Google Scholar
Barreto, P., Arima, E., and Brito, M. (2006). Cattle ranching and challenges for environmental conservation in the Amazon. State of the Amazon No. 5. Belem, Brazil: IMAZON.
Barron, A. R. (2007). Patterns and controls of nitrogen fixation in a lowland tropical forest, Panama, Ph.D. diss. Princeton, NJ, Princeton University.
Barroso, C. B., and Nahas, E. (2005). The status of soil phosphate fractions and the ability of fungi to dissolve hardly soluble phosphates. Applied Soil Ecol., 29, 73–83.Google Scholar
Barson, M., Randall, L., and Bordas, V. (2000). Land Cover Change in Australia. Results of the Collaborative Bureau of Rural Sciences – State Agencies’ Project on Remote Sensing of Land Cover Change. Kingston ACT, Australia: Bureau of Rural Sciences.
Barton, L., McLay, C. D. A., Schipper, L. A., and Smith, C. T. (1999). Annual denitrification rates in agricultural and forest soils: A review. Soil Research, 37(6), 1073–1094.Google Scholar
BassiriRad, H., Gutschick, V. P., and Lussenhop, J. (2001). Root system adjustments: Regulation of plant nutrient uptake and growth responses to elevated CO2. Oecologia, 126(3), 305–320.Google Scholar
Bauen, A., Chudziak, C., Vad, K., and Watson, P. (2010). A Casual Descriptive Approach to Modelling the GHG Emissions Associated with the Indirect Land Use Impacts of Biofuels. Final Report. A Study for the UK Department for Transport, E4tech, London.
Bauhus, J., and Barthel, R. (1995). Mechanisms for carbon and nutrient release and retention in beech forest gaps. II. The role of soil microbial biomass. Plant and Soil, 168–169, 585–592.Google Scholar
Beare, M. H., Hendrix, P. F., and Coleman, D. C. (1994). Water-stable aggregates and organic matter fractions in conventional and no-tillage soils. Soil Sci. Soc. Am. J., 58, 777–786.Google Scholar
Beguería, S. (2006). Changes in land cover and shallow landslide activity: A case study in the Spanish Pyrenees. Geomorphology, 74(1), 196–206.Google Scholar
Belcher, B., and Schreckenberg, K. (2007). Commercialisation of non-timber forest products: A reality check. Development Policy Review, 25(3), 355–377.Google Scholar
Bengtsson, J., and Wikström, F. (1993). Effects of whole-tree harvesting on the amount of soil carbon: model results. New Zealand Journal of Forest Science 23, 380–389.
Berg, A. S., and Joern, B. C. (2006). Sorption dynamics of organic and inorganic phosphorus compounds in soil. J. Environ. Qual., 35, 1855–1862.Google Scholar
Berglund, B. E. (2006). Agrarian landscape development in northwestern Europe since the neolithic: Cultural and climatic factors behind a regional/continental pattern. In Hornburg, A., andCrumley, C. L. (eds.), The World System and the Earth System: Global Socioenvironmental Change and Sustainability since the Neolithic. Left Coast Press, Walnut Creek, CA.
Berhe, A. A., Harte, J., Harden, J. W., and Torn, M. S. (2007). The significance of the erosion-induced terrestrial carbon sink. Bioscience, 57, 337–346.Google Scholar
Beringer, J., Chapin, F. S., Thompson, C. C., and McGuire, A. D. (2005). Surface energy exchanges along a tundra-forest transition and feedbacks to climate. Agricultural and Forest Meteorology, 131, 143–161.Google Scholar
Berndes, G., Bird, N., and Cowie, A. (2010). Bioenergy, Land Use Change and Climate Change Mitigation, IEA Bioenergy, Aadorf, Switzerland
Bernhardt, E. S., Likens, G. E., Buso, D. C., and Driscoll, C. T. (2003). In-stream uptake dampens effects of major forest disturbance on watershed nitrogen export. Proceedings of the National Academy of Sciences, 100(18), 10304–10308.Google Scholar
Berry, W. L. (1970). Characteristics of salts secreted by Tamarix aphylla. American Journal of Botany, 57, 1226–1230.Google Scholar
Bertness, M. D., and Leonard, G. H. (1997). The role of positive interactions in communities: Lessons from intertidal habitats. Ecology, 78(7), 1976–1989.Google Scholar
Best, A., L. et al. (2007). A critical review of paired catchment studies with reference to seasonal flows and climatic variability, Murray Darling Basin Comission Publication 11/03, Murray-Darling Basin Commission, Canberra, ACT, Australia.
Betts, A. K., and Ball, J. H. (1997). Albedo over the boreal forest. Journal of Geophysical Research, 102, 28,901–28,909.Google Scholar
Betts, R. A. (2000). Offset of the potential carbon sink from boreal forestation by decreases in surface albedo. Nature, 408, 187–190.Google Scholar
Bever, J. D., Schultz, P. A., Pringle, A., and Morton, J. B. (2001). Arbuscular mycorrhihzal fungi: More diverse than meets the eye and the ecological tale of why. BioScience, 51(11), 923–932.Google Scholar
Bhattarai, M., and Hammig, M. (2001). Institutions and the environmental Kuznets curve for deforestation: A crosscountry analysis for Latin America, Africa and Asia. World Development, 29(6), 995–1010.Google Scholar
Biggs, T. W., Dunne, T., and Martinelli, L. A. (2004). Natural controls and human impacts on stream nutrient concentrations in a deforested region of the Brazilian Amazon basin. Biogeochemistry, 68(2), 227–257.Google Scholar
Billington, C., Kapos, V., Edwards, M. S., Blyth, S., and Iremonger, S. (1996). Estimated Original Forest Cover Map – a First Attempt. WCMC, Cambridge, UK.
Bischetti, G. B., Chiaradia, E. A., Epis, T., and Morlotti, E. (2009). Root cohesion of forest species in the Italian Alps. Plant and Soil, 324(1–2), 71–89.Google Scholar
Bischetti, G. B., Chiaradia, E. A., Simonato, T., Speziali, B., Vitali, B., Vullo, P., and Zocco, A. (2005). Root strength and root area ratio of forest species in Lombardy (Northern Italy). Plant and Soil, 278, 11–22.Google Scholar
Bishop, D. M., and Stevens, M. E. (1964). Landslides on logged areas in southeast Alaska. USDA Forest Service, Northern Forest Experiment Station Research Paper NOR-1.
Blaike, P., and Brookfield, H. (1987). Land Degradation and Society. Methuen, London.
Bluffstone, R. A. (1995). The effect of labor market performance on deforestation in developing countries under open access: An example from rural Nepal. Journal of Environmental Economics and Management, 29(1), 42–63.Google Scholar
Boeken, B., and Orenstein, D. (2001). The effect of plant litter on ecosystem properties in a Mediterranean semi-arid shrubland. J. Veg. Sci., 12(6), 825–832.Google Scholar
Bohn, H., and Deacon, R. T. (2000). Ownership risk, investment, and the use of natural resources. American Economic Review, 526–549.Google Scholar
Bonan, G. (2008a). Ecological Climatology. Cambridge University Press, New York.
Bonan, G. B. (2008b). Forests and climate change: Forcings, feedbacks, and the climate benefits of forests. Science, 320, 1444–1449.Google Scholar
Bonan, G. B., Pollard, D., and Thompson, S. L. (1992). Effects of boreal forest vegetation on global climate. Nature, 359, 716–718.Google Scholar
Bonan, G. B., and Shugart, H. H. (1989). Environmental factors and ecological processes in boreal forests. Annu. Rev. Ecol. Syst., 20, 1–28.Google Scholar
Borg, H., Stoneman, G. L., and Ward, C. G. (1988). The effect of logging and regeneration on groundawater, streamflow and stream salinity in the southern forest of Western Australia. J. Hydrol., 99, 253–271.Google Scholar
Borlaug, N. (2007). Feeding a hungry world. Science 318(5849): 359–359.Google Scholar
Bormann, F. H., and Likens, G. E. (1979). Pattern and Process in a Forested Ecosystem. Springer-Verlag, New York.
Borneman, J., and Triplett, E. W. (1997). Molecular and microbial diversity in soils from Eastern Amazonia: Evidence for unusual microorganisms and microbial population shifts associated with deforestation. Applied and Environmental Microbiology, 63(7), 2647–2653.Google Scholar
Bosch, J. M., and Hewlett, J. A. (1982). A review of catchment experiments to determine the effect of vegetation changes on water yield and evapotranspiration. J. Hydrol., 55, 3–23.Google Scholar
Bossio, D. A. et al. (2005). Soil microbial community response to land use change in an agricultural landscape of western Kenya. Microbial Ecology, 49, 50–62.Google Scholar
Bouwman, A. F., Lee, D. S., Asman, W. A. H., Dentener, F. J., Van Der Hoek, K. W., and Olivier, J. G. J. (1997). A global high-resolution emission inventory for ammonia. Global biogeochemical cycles, 11(4), 561–587.Google Scholar
Bowman, D. M. et al. (2011). The human dimensions of fire regimes on Earth. J. Biogeogr., 1–14, doi:10.1111/j.1365-2699.2011.02595.x.Google Scholar
Brabb, E. E., and Harrod, B. L. (1989). Landslides: Extent and economic significance. In Proceedings of the International Geological Congress (No. 28).
Bradshaw, C. J. (2012). Little left to lose: Deforestation and forest degradation in Australia since European colonization. Journal of Plant Ecology, 5(1), 109–120.Google Scholar
Brady, N., and Weil, R. (2008). The Nature and Properties of Soils, 14th ed. Prentice Hall, Upper Saddle River, NJ.
Braithwaite, L. W. (1996). Conservation of arboreal herbivores: The Australian scene. Aust. J. Ecol., 21, 21–30.Google Scholar
Brannstrom, C. 2002. Rethinking the ‘Atlantic forest’ of Brazil: new evidence for land cover and land value in western Sao Paulo, 1900–1930.Journal of Historical Geography 28, 420–439.Google Scholar
Brannstrom, C., and Oliveira, A. M. S. (2000). Human modification of stream valleys in the western plateau of Sao Paulo, Brazil: Implications for environmental narratives and management. Land Degrad. Dev. 11, 535–548.Google Scholar
Brardinoni, F., Hassan, M. A., and Slaymaker, H. O. (2002). Complex mass wasting response of drainage basins to management in coastal British Columbia. Geomorphology, 49, 109–124.Google Scholar
Brauman, K. A., Daily, G. C., Duarte, T. K. E., and Mooney, H. A. (2007). The nature and value of ecosystem services: An overview highlighting hydrologic services. Annu. Rev. Environ. Resour., 32, 67–98.Google Scholar
Breman, H., and Kessler, J. J. (1995). Woody Plants in Agro-Ecosystems of Semi-Arid Regions. Springer, New York.
Briggs, J. M., and Knapp, A. K. (2001). Determinants of C3 forb growth and production in a C4 dominated grassland. Plant Ecology, 152, 93–100.Google Scholar
Brink, B. T. et al. (2007). Cross-roads of planet earth's life: Exploring means to meet the 2010 biodiversity target: Solution-oriented scenarios for Global Biodiversity Outlook 2. CBD technical series (31).
Brock, W. A., and Carpenter, S. R. (2006). Variance as a leading indicator of regime shift in ecosystem services. Ecology and Society, 11(2), 9.Google Scholar
Broich, M., Hansen, M. C., Potapov, P., Adusei, B., Lindquist, E., and Stehman, S. V. (2011). Time-series analysis of multi-resolution optical imagery for quantifying forest cover loss in Sumatra and Kalimantan, Indonesia. International Journal of Applied Earth Observation and Geoinformation, 13(2), 277–291.Google Scholar
Brolsma, R. J., and Bierkens, M. P. F. (2007). Groundwater – soil water – vegetation dynamics in a temperate forest ecosystem along a slope. Water Resour. Res., 43, W01414, doi:10.1029/2005WR004696.Google Scholar
Brook, B. W., Sodhi, N. S., and Bradshaw, C. J. (2008). Synergies among extinction drivers under global change. Trends in Ecology and Evolution, 23(8), 453–460.Google Scholar
Brooks, T. M., Pimm, S. L., and Oyugi, J. O. (1999). Time lag between deforestation and bird extinction in tropical forest fragments. Conservation Biology, 13(5), 1140–1150.Google Scholar
Brovkin, V. et al. (1998). On the stability of the atmosphere – vegetation system in the Sahara/Sahel region. Journal of Geophysical Research – Atmospheres, 103, 31613–31624.Google Scholar
Brown, A. E., Zhang, L., McMahon, T. A., Western, A. W., and Vertessy, R. A. (2005). A review of paired catchment studies for determining changes in water yield resulting from alterations in vegetation. Journal of Hydrology, 310(1), 28–61.Google Scholar
Brown, J. C., Jepson, W., and Price, K. P. (2004). Expansion of mechanized agriculture and land-cover change in Southern Rondônia, Brazil. J. Latin Am. Geogr., 3, 96–102.Google Scholar
Brown, S., and Lugo, A. E. (1990). Tropical secondary forests. Journal of Tropical Ecology, 6(01), 1–32.Google Scholar
Bruce, J. P., Frome, M., Haites, E., Janzen, H., Lal, R., and Paustian, K. (1999). Carbon sequestration in soils. Journal of Soil and Water Conservation, 54(1), 382–389.Google Scholar
Bruce, J., Wendland, K., Naughton-Treves, L., 2010. Whom to Pay? Key Concepts and Terms regarding Tenure and Property Rights in Payment-based Forest Ecosystem Conservation, Land Tenure Center Policy Brief. Land Tenure Center, Madison, WI.
Bruijnzeel, L. A. (2002). Hydrology and water management in the humid tropics, Proceedings of the Second International Colloquium, 22–26 March 1999, International Hydrology Programme, Technical Documents in Hydrology: No. 52 UNESCO, Paris.
Bruijnzeel, L. A., and Veneklaas, E. J. (1998). Climatic conditions and tropical montane forest productivity: The fog has not lifted yet. Ecology 79(1), 3–9.Google Scholar
Bruschi, V. M. et al. (2013). Land management versus natural factors in land instability: Some examples in northern Spain. Environmental Management, 52(2), 398–416.Google Scholar
Brussaard, L., De Ruiter, P. C., and Brown, G. G. (2007). Soil biodiversity for agricultural sustainability. Agriculture, Ecosystems and Environment, 121(3), 233–244.Google Scholar
Bryson, R. A., Irving, W. N., and Larsen, J. A. (1965). Radiocarbon and soil evidence of former forest in the southern Canadian tundra. Science, 147, 46–48.Google Scholar
Bubb, P., May, I., Miles, L., and Sayer, J. (2004). Cloud Forest Agenda. UNEP-WCMC, Cambridge, UK, available at: http://www.unep-wcmc.org/resources/publications/UNEP_WCMC_bio_series/20.htm.
Buchner, O., and Neuner, G. (2011). Winter frost resistance of Pinus cembra measured in situ at the alpine timberline as affected by temperature conditions. Tree Physiology, 31, 1217–1227.Google Scholar
Bullock, D. G. (1992). Crop rotation. Critical Reviews in Plant Sciences, 11(4), 309–332.Google Scholar
Bullock, J. M., Pywell, R. F., Burke, M. J., and Walker, K. J. (2001). Restoration of biodiversity enhances agricultural production. Ecology Letters, 4(3), 185–189.Google Scholar
Bulson, H. A. J., Snaydon, R. W., and Stopes, C. E. (1997). Effects of plant density on intercropped wheat and field beans in an organic farming system. Journal of Agricultural Science, 128(1), 59–71.Google Scholar
Burdon, J. J. (1993). The structure of pathogen populations in natural plant communities. Annual Review of Phytopathology, 31(1), 305–323.Google Scholar
Burgess, J. C. (1993). Timber production, timber trade and tropical deforestation. Ambio, 136–143.Google Scholar
Burgess, S. S., Adams, M. A., Turner, N. C., and Ong, C. K. (1998). The redistribution of soil water by tree root systems. Oecologia, 115(3), 306–311.Google Scholar
Burney, J. A., Davis, S. J., and Lobell, D. B. (2010). Greenhouse gas mitigation by agricultural intensification. Proceedings of the national Academy of Sciences, 107(26), 12052–12057.Google Scholar
Burton, P. J., Messier, C., Weetman, G. F., Prepas, E. E., Adamowicz, W. L., and Tittler, R. (2003). The current state of boreal forestry and the drive for change. In Towards Sustainable Management of the Boreal Forest. In Burton, P. J., Messier, C., Smith, D. W. and Adamowicz, W. L.. NRC Research Press, Ottawa, pp. 1–40.
Butt, N., de Oliveira, P. A., and Costa, M. H. (2011). Evidence that deforestation affects the onset of the rainy season in Rondonia, Brazil. J. Geophys. Res.116,D11120.Google Scholar
Byerlee, D., and Rueda, X. (2015). From public to private standards for tropical commodities: A century of global discourse on land governance on the forest frontier. Forests, 6(4), 1301–1324.Google Scholar
Cai, X., McKinney, D. C., and Rosegrant, M. W. (2003). Sustainability analysis for irrigation water management in the Aral Sea region. Agric. Syst., 76, 1043–1066.Google Scholar
Caldwell, B. A., Griffiths, R. P., and Sollins, P. (1999). Soil enzyme response to vegetation disturbance in two lowland Costa Rican soils. Soil Biology and Biogeochemistry, 31, 1603–1608.Google Scholar
Camacho Villa, T. C., Maxted, N., Scholten, M. A., and Ford-Lloyd, B. V. (2005). Defining and identifying crop landraces. Plant Genetic Resources: Characterization and Utilization, 3(3), 373–384.Google Scholar
Camargo, A. M. M. P. (2008). Dynamics and tendency of sugarcane expansion over other agricultural activities. State of São Paulo 2001–2006. Informações Econômicas, 38(3), 47–66.
Campbell, G. S., and Norman, J. M. (1998). An Introduction to Environmental Biophysics, 2nd ed. Springer, New York, doi:10.1007/978-1-4612-1626-1.
Campbell, J. E., Lobell, D. B., Genova, R. C., and Field, C. B. (2008). The global potential of bioenergy on abandoned agriculture lands. Environmental Science and Technology, 42(15), 5791–5794.Google Scholar
Canadell, J. G. et al. (2007). Saturation of the terrestrial carbon sink. In Terrestrial Ecosystems in a Changing World. Springer, Berlin and Heidelberg, pp. 59–78.
Canadian Council of Forest Ministers (CCFM). (2002). Criteria and indicators of sustainable forest management in Canada: National status 2005. Ottawa, Natural Resources Canada.
Carle, J., and Holmgren, P. (2008). Wood from planted forests: A global outlook 2005–2030. For. Prod. J., 58, 6–18.Google Scholar
Carlson, K. M. et al. (2012). Expanding oil palm plantations in West Kalimantan, Indonesia: Impacts on land cover change and carbon emissions. Proceedings of the National Academy of Sciences, doi/10.1073/pnas.1200452109.Google Scholar
Carlyle-Moses, D. E. (2004). Throughfall, stemflow, and canopy interception loss fluxes in a semi-arid Sierra Madre Oriental matorral community. Journal of Arid Environments, 58(2), 181–202.Google Scholar
Carnicer, J., Barbeta, A., Sperlich, D., Coll, M., and Peñuelas, J. (2013). Contrasting trait syndromes in angiosperms and conifers are associated with different responses of tree growth to temperature on a large scale. Front. Plant Sci. 4:409, doi:10.3389/fpls.2013.00409.Google Scholar
Carpenter, S., and Brock, W. A. (2006). Rising variance: A leading indicator of ecological transition. Ecol. Lett., 9, 308–315.Google Scholar
Carpenter, S. R. et al. (2011). Early warnings of regime shifts: A whole ecosystem experiment. Science, 10.1126/science.1203672.Google Scholar
Carr, D. (2009). Population and deforestation: Why rural migration matters. Progress in Human Geography, 33(3), 355–378.Google Scholar
Carter, C., Finley, W., Fry, J., Jackson, D., and Willis, L. (2007). Palm oil markets and future supply. European Journal of Lipid Science and Technology, 109(4), 307–314.Google Scholar
Cassman, K. G. (1999). Ecological intensification of cereal production systems: Yield potential, soil quality, and precision agriculture. Proceedings of the National of Sciences, 96(11), 5952–5959.Google Scholar
Cassman, K. G., Kropff, M. J., Gaunt, J., and Peng, S. (1993). Nitrogen use efficiency of rice reconsidered: What are the key constraints?Plant and Soil, 155(1), 359–362.Google Scholar
Cavelier, J., and Goldstein, G. (1989). Mist and fog interception in elfin cloud forest in Colombia and Venezuela. Journal of Tropical Ecology, 5, 309–322.Google Scholar
Cavelier, J., Solis, D., and Jaramillo, M. A. (1996). Fog interception in montane forests across the Central Cordillera of Panama. Jounral of Tropical Ecology, 12, 357–369.Google Scholar
Cavender-Bares, J., Cortes, P., Rambal, S., Joffre, R., Miles, B., and Rocheteau, A. (2005). Summer and winter sensitivity of leaves and xylem to minimum freezing temperatures: A comparison of co-occurring Mediterranean oaks that differ in leaf lifespan. New Phytologist, 168, 597–612.Google Scholar
Cernusak, L. A. et al. (2013). Tropical forest responses to increasing atmospheric CO2: Current knowledge and opportunities for future research. Functional Plant Biology, 40(6), 531–551.Google Scholar
Cerri, C. C., Volkoff, B., and Andreaux, F. (1991). Nature and behaviour of organic matter in soils under natural forest, and after deforestation, burning and cultivation, near Manaus. Forest Ecology and Management, 38(3), 247–257.Google Scholar
Cerri, C. C., Volkoff, B., and Eduardo, B. P. (1985). Efeito do desmatament sobre a biomassa microbiana em latossolo amarelo da Amazoˆnia. Revista Brasileira Cieˆn. Solo, 9(1), 1–4.Google Scholar
Chaer, G., Fernandes, M., Myrold, D., and Bottomley, P. (2009). Comparative resistance and resilience of soil communities and enzyme activities in adjacent native forest and agricultural soils. Microb. Ecol., 58, 414–424.Google Scholar
Chagnon, F. J. F., and Bras, R. L. (2005). Contemporary climate in the Amazon.Geophys. Res. Lett., 32, L13703.Google Scholar
Chagnon, F. J. F., Bras, R. L., and Wang, J. (2004). Climatic shift in patterns of shallow cumulus clouds over the Amazon. Geophys. Res. Lett., 31, L24212, doi:10.1029/2004GL021188.Google Scholar
Chamberlain, A. C., and Little, P. (1981). Transport and capture of particles by vegetation. In Grace, J., E. D. Ford, , and Jarvis, P. G. (eds.), Plants and Their Atmospheric Environment. Blackwell, Oxford, UK.
Chambers, J. C., and Linnerooth, A. R. (2001). Restoring riparian meadows currently dominated by Artemisia using alternative state concepts – the establishment component. Appl. Veg. Sci., 4, 157–166.Google Scholar
Chang, M., (2002). Forest Hydrology: An Introduction to Water and Forests. CRC Press, Boca Raton, FL.
Chapin, F. S., Eugster, W., McFadden, J. P., Lynch, A. H., and Walker, D. A. (2000a). Summer differences among Arctic ecosystems in regional climate forcing. Journal of Climate, 13, 2002–2010.Google Scholar
Chapin, F. S., Matson, P. A., and Mooney, H. A.(2002). Principles of Terrestrial Ecosystem Ecology. Springer-Verlag, New York.
Chapin, F. S., and Shaver, G. R. (1985). Individualistic growth response of tundra plant species to environmental manipulations in the field. Ecology, 66(2), 564–576.Google Scholar
Chapin, F. S., van Cleve, K., and Chapin, M. C. (1979). Soil temperature and nutrient cycling in the tussock growth form of Eriophorum vaginatum. Journal of Ecology, 67(1), 169–189.Google Scholar
Chapin, F. S. et al. (2000b). Arctic and boreal ecosystems of western North America as components of the climate system. Global Change Biology, 6, 211–223.Google Scholar
Chapin, F. S. et al. (2000c). Consequences of changing biodiversity. Nature, 405(6783), 234–242.Google Scholar
Chapin, F. S. et al. (2005). Role of land-surface changes in Arctic summer warming. Science, 310, 657–660.Google Scholar
Chapin, F. S. et al. (2006). Reconciling carbon-cycle concepts, terminology, and methods. Ecosystems, 9(7), 1041–1050.Google Scholar
Chapin, F. S. et al. (2010). Ecosystem stewardship: Sustainability strategies for a rapidly changing planet. Trends in Ecology and Evolution, 25, 241–249.Google Scholar
Chapman, V. (1975). Mangrove biogeography. Proceedings of the International Symposium on Biology and Management of Mangroves,ed. by Walsh, G., S. Snedaker and H. Teas, pp. 3–22.
Charles, D. (2001). Seeds of discontent. Science, 294(5543), 772–775.Google Scholar
Charley, J. L., and West, N. E. (1975). Plant-induced soil chemical patterns in some shrub dominated semi-desert ecosystems of Utah. Journal of Ecology, 63, 945–963.Google Scholar
Chen, H., Qualls, R. G., and Miller, G. C. (2002). Adaptive responses of Lepidium latifolium to soil flooding: Biomass allocation, adventitious rooting, aerenchyma formation and ethylene production. Environmental and Experimental Botany, 48, 119–128.Google Scholar
Chen, J., Franklin, J. F., and Spies, T. A. (1993). Contrasting microclimates among clearcut, edge, and interior of old-growth Douglas fir forest. Agricultural and Forest Meteorology, 63, 219–237.Google Scholar
Chen, T. C., Yoon, J., St. Croix, K. J., and Takle, E. (2001). Suppressing impacts of the Amazonian deforestation by global circulation change. Bull. Am. Meteorol. Soc., 82, 2209–2216.Google Scholar
Chengrui, M., and Dregne, H. E. (2001). Review article: Silt and the future development of China's Yellow River. The Geographical Journal, 167(1), 7–22.Google Scholar
Chibnik, M. (1994). Risky rivers: The economics and politics of floodplain farming in Amazonia. University of Arizona Press, Tucson.
Chokkalingam, U, and De Jong, W. (2001). Secondary forest: A working definition and typology. Int. For. Rev., 3, 19–26.Google Scholar
Chomitz, K. M., Alger, K., Thomas, T. S., Orlando, H., and Nova, P. V. (2005). Opportunity costs of conservation in a biodiversity hotspot: The case of southern Bahia. Environment and Development Economics, 10(03), 293–312.Google Scholar
Chomitz, K., Buys, P., Luca, G. De, Thomas, T. S., and Wertz-Kanounnikoff, S. (2006). At loggerheads? Agricultural expansion, poverty reduction and environment in the tropical forests. World Bank Publications.
Choné, T., Andreux, F., Correa, J. C., Volkoff, B., and Cerri, C. C. (1991). Changes in organic matter in an oxisol from the Central Amazonian forest during eight years as pasture, determined by 13C isotopic composition. In Berthelin, J. (ed.), Diversity of Environmental Biogeochemistry. Elsevier, Amsterdam, pp. 397–405.
Chopra, K. (1993). The value of non-timber forest products: An estimation for tropical deciduous forests in India. Economic Botany, 47(3), 251–257.Google Scholar
Chu, P. S., Yu, Z.-P., and Hastenrath, S. (1994). Detecting climate change concurrent with deforestation in the Amazon Basin: Which way has it gone?Bull. Am. Meteorol. Soc., 75, 579–582.Google Scholar
Clarke, C. J., George, R. J., Bell, R. W., and Hatton, T. J. (2002). Dryland salinity in south-western Australia: Its origins, remedies and furute research directions. Aust J Soil Res 2002; 40, 93–113.Google Scholar
Claussen, M. (1997). Modeling bio-geophysical feedback in the African and Indian monsoon región. Climate Dynamics, 13(4), 247–257.Google Scholar
Claussen, M., Brovkin, V., Ganopolski, A., Kubatzki, C., and Petoukhov, V. (2003). Climate change in Northern Africa: The past is not the future. Climatic Change, 57, 99–118.Google Scholar
Clay, J. W., and Clement, C. R. (1993). Selected species and strategies to enhance income generation from Amazonian forests. FO:Misc/93/6 Working Paper. FAO, Rome.
Cleveland, C. C. et al. (1999). Global patterns of terrestrial biological nitrogen (N2) fixation in natural ecosystems. Global Biogeochem. Cycles, 13(2), 623–645, doi:10.1029/1999GB900014.Google Scholar
Cleveland, C. C., and Liptzin, D. (2007). C: N: P stoichiometry in soil: Is there a “Redfield ratio” for the microbial biomass?Biogeochemistry, 85(3), 235–252.Google Scholar
Cleveland, C. C., Townsend, A. R., and Schmidt, S. K. (2002). Phosphorus limitation of microbial processes in moist tropical forests. Ecosystems, 5, 680–691.Google Scholar
Cochrane, M. A. (2003). Fire science for rainforests. Nature, 421, 913–919.Google Scholar
Cochrane, M. A., and Barber, C. P. (2009). Climate change, human land use and future fires in the Amazon. Global Chnage Biology, 15, 601–612.Google Scholar
Cochrane, M. A. et al. (1999). Positive feedbacks in the fire dynamic of closed canopy tropical forests. Science, 284, 1832–1835.Google Scholar
Cohuet, A. et al. (2004). High malaria transmission intensity due to Anopheles funestus (Diptera: Culicidae) in a village of savannah-forest transition area in Cameroon. J Med Entomol 41(5), 901–905.Google Scholar
Coimbra, C. E. A. (1991). Environmental change and human disease: A view from Amazonia. Journal of Human Ecology, 2, 15–21.Google Scholar
Collatz, G. J. et al. (1991). Physiological and environmental regulations of stomatal conductance, photosynthesis and transpiration: A model that includes a laminar boundary-layer. Agric. For. Meteorol., 54(2–4), 107–136, doi:10.1016/0168-1923(91) 90002–8.Google Scholar
Colpaert, J. V., van Laere, A., van Tichelen, K. K., and van Assche, J. A. (1997). The use of inositol hexaphosphate as a phosphorus source by mycorrhizal and non-mycorrhizal Scots pine (Pinus sylvestris). Funct. Ecol., 11, 407–415.Google Scholar
Coluzzi, M. (1994). Malaria and the Afrotropical ecosystems: Impact of man-made environmental changes. Parassitologia, 36(1–2), 223–227.Google Scholar
Commonwealth of Australia. (1999). International forest conservation: Protected areas and beyond. A discussion paper for the intergovernmental forum on forests. Commonwealth of Australia, Canberra, Australia.
Conlin, T. S. S., and Lieffers, V. J. (1992). Root respiration of boreal forest conifers under anoxia and cool conditions. Can. J. For. Res., 23, 767–771.Google Scholar
Convention on Biological Diversity. (2001). Report of the ad hoc technical expert group on forest biological diversity, available at: https://www.cbd.int/forest/definitions.shtml (accessed on June 25, 2015).
Cook, B. I., Anchuk, K. A., Kaplan, J., Puma, M. J., Kelley, M., and Gueyffier, D. (2012). Pre-Colombian deforestation as an amplifier of drought in Mesoamerica. Geophysical Research Letters, doi:10.1029/2012GL052565.Google Scholar
Cordell, D., Drangert, J. O., and White, S. (2009). The story of phosphorus: Global food security and food for thought. Global Environmental Change, doi:10.1016/j.gloenvcha.2008.10.09.Google Scholar
Corman, A., Crozat, Y., and Cleyet-Marel, J. C. (1987). Modeling of survival kinetics of some Bradyrhizobium japonicum strains in soils. Biol. Fertil. Soils, 4, 79–84.Google Scholar
Costa, M. H., and Foley, J. A. (1999). Trends in the hydrological cycle of the Amazon basin. Journal of Geophysical Research – Atmospheres, 104, 14189–14198.Google Scholar
Costa, M. H., and Foley, J. A. (2000). Combined effects of deforestation and doubled atmospheric CO2 concentrations on the climate of Amazonia. J. Clim., 13, 18–34.Google Scholar
Costa, M. H. and Foley, J. H. (1997). Water balance of the Amazon Basin: dependence on vegetation cover and canopy conductance.J. Geophys. Res., 102, 23973–23989.Google Scholar
Costanza, R., and Folke, C. (1997). Valuing ecosystem services with efficiency, fairness, and sustainability as goals. In Nature's Services: Societal Dependence on Natural Ecosystems. Island Press, Washington, DC, pp. 49–70.
Costello, M. J., May, R. M., and Stork, N. E. (2013). Can we name Earth's species before they go extinct?Science, 339(6118), 413–416.Google Scholar
Costello, M. J., Wilson, S., and Houlding, B. (2012). Predicting total global species richness using rates of species description and estimates of taxonomic effort. Systematic Biology, 61(5), 871–883.Google Scholar
Cotula, L. 2011. Land Deals in Africa: What is in the Contracts?London: IIED. http://pubs.iied.org/pdfs/12568IIED.pdf
Covington, W. W., and Sackett, S. S. (1992). Soil mineral nitrogen changes following prescribed burning in ponderosa pine. Forest Ecology and Management, 54(1), 175–191.Google Scholar
Craft, A. B., and Simpson, R. D. (2001). The value of biodiversity in pharmaceutical research with differentiated products. Environ. Resource Econ. 18, 1–17.Google Scholar
Cramer, V. A., and Hobbs, R. J. (2002). Ecological consequences of altered hydrological regimes in fragmented ecosystems in southern Australia: Impacts and possible management responses. Austral Ecol., 27, 546–564.Google Scholar
Crews, T. E. (1999). The presence of nitrogen fixing legumes in terrestrial communities: Evolutionary vs ecological considerations. Biogeochemistry, 46(1), 233–246.Google Scholar
Crews, T. E., Farrington, H., and Vitousek, P. M. (2000). Changes in asymbiotic, heterotrophic nitrogen fixation on leaf litter of metrosideros polymorpha with long-term ecosystem development in Hawaii. Ecosystems, 3(4), 386–395, doi:10.1007/s100210000034.Google Scholar
Cromack, K., Todd, R. L., and Monk, C. D. (1975). Patterns of basidiomycete nutrient accumulation in conifer and deciduous forest litter. Soil Biology and Biochemistry 7, 265–268.Google Scholar
Cropper, M., and Griffiths, C. (1994). The interaction of population growth and environmental quality. American Economic Review, 250–254.Google Scholar
Crozier, M. J. (2005). Multiple-occurrence regional landslide events in New Zealand: Hazard management issues. Landslides, 2(4), 247–256.Google Scholar
Culas, R. J. (2007). Deforestation and the environmental Kuznets curve: An institutional perspective. Ecological Economics, 61(2), 429–437.Google Scholar
Culf, A. D., Esteves, J. L., Marques Filho, A. O., and Rocha, H. R. (1996). Radiation, temperature and humidity over forest and pasture in Amazonia. In Gash, J. H. C., Nobre, C. A., Roberts, J., and Victoria, R. L. (eds.), Amazonian Deforestation and Climate. Chichester, UK, John Wiley, pp. 175–191.
Da Silveira, L. and Sternberg, L. (2001). Savanna-forest hysteresis in the tropics. Global Ecology and Biogeography, 10(4), 369–378.Google Scholar
Daily, G. C., Ehrlich, P. R., and Sanchez-Azofeifa, G. A. (2001). Countryside biogeography: Use of human-dominated habitats by the avifauna of southern Costa Rica. Ecological Applications, 11(1), 1–13.Google Scholar
Dakora, F. D., and Phillips, D. A. (2002). Root exudates as mediators of mineral acquisition in low-nutrient. Plant and Soil, 245, 35–47.Google Scholar
Dakos, V., Scheffer, M,. van New, E. H., Brovkin, V., Petoukhov, V., and Held, H. (2008). Slowing down as an early warning signal for abrupt climate change. PNAS, 105(38), 14308–14312.Google Scholar
Dames, Moore. (1980). Report on the Rehabilitation of Rio Blanco Hydroelectric Project for Puerto Rico Electric Power Authority, Houston, TX, Job No. 11905-002-14.
D'Antonio, C. M., and Vitousek, P. M. (1992). Biological invasions by exotic grasses, the grass/fire cycle, and global change. Annual Review of Ecology and Systematics, 23, 63–87.Google Scholar
Darby, H. C. (1956). The clearing of the woodland in Europe. In Thomas, W. L. Jr. (ed.), Man's Role in Changing the Face of the Earth. University of Chicago Press, Chicago.
Das, R., Lawrence, D., D'Odorico, P., and DeLonge, M. (2011). Impact of land use change on atmospheric P inputs in a tropical dry forest. J. Geophys. Res., doi:10.1029/2010JG001403.Google Scholar
Dasgupta, P. (2013). The nature of economic development and the economic development of nature. Economic and Political Weekly, XLVIII, 38–51.Google Scholar
Daszak, P., Cunningham, A. A., and Hyatt, A. D. (2001). Anthropogenic environmental change and the emergence of infectious diseases in wildlife. Acta Tropica, 78(2), 103–116.Google Scholar
Daubenmire, R. (1954) Alpine timberlines in the Americas and their interpretation. Butler University Botanical Studies, 11, 119–136.Google Scholar
Davidson, E. A. et al. (2004). Nitrogen and phosphorus limitation of biomass growth in a tropical secondary forest. Ecological Applications, 14(4), Supplement, S150–S163.Google Scholar
Davidson, E. A. et al. (2012). The Amazon basin in transition. Nature, 481, 321–328.Google Scholar
Davies, K. F., Margules, C. R., and Lawrence, J. F. (2004). A synergistic effect puts rare, specialized species at greater risk of extinction. Ecology, 85(1), 265–271.Google Scholar
Davies-Colley, R. J., Payne, G. W., and van Elswijk, M. (2000). Microclimate gradients across a forest edge. New Zealand Journal of Ecology, 24(2), 111–121.Google Scholar
Davis, K. F., D'Odorico, P., and Rulli, M. C. (2014). Moderating diets to feed the future. Earth's Future, 2, doi:10.1002/2014EF000254.Google Scholar
Davis, K. F., Yu, K., Rulli, M. C., Pichdara, L., and D'Odorico, P. (2015). Accelerated deforestation by large-scale land acquisitions in Cambodia, Nature Geoscience, 8, 772-775, DOI: 10.1038/NGEO2540.
Davison, E. M., and Tay, F. C. S. (1985). The effect of waterlogging on seedlings of Eucalyptus marginata. New Phytologist, 101(4), 743–753.Google Scholar
Dawson, T. E. (1998). Fog in the California redwood forest: Ecosystem inputs and use by plants. Oecologia, 117, 4, 476–485.Google Scholar
de Almeida, A. L. O., and Campari, J. S. (1995). Sustainable settlement in the Brazilian Amazon. Oxford University Press, Oxford.
DeBano, L. F. (1966). Formation of non-wettable soils involves heat transfer mechanism, USDA Forest Service Research Note PSW-132.
DeBano, L. F. (2000). The role of fire and soil heating on water repellency in wildland environments: A review. J. Hydrol., 231, 195–206.Google Scholar
DeBano, L. F., Savage, S. M., and Hamilton, D. A. (1976). The transfer of heat and hydrophobic substances during burning. Soil Science Society America Journal, 40, 779–782.Google Scholar
de Castro, M. C., Monte-Mór, R. L., Sawyer, D. O., Singer, B. H. (2006). Malaria risk on the Amazon frontier. Proc. Natl. Acad. Sci. USA, 103(7), 2452–2457.Google Scholar
de Chantal, M., Holt Hanssen, K., Granhus, A., Bergsten, U., Ottosson Löfvenius, M., and Grip, H. (2007). Frost-heaving damage to one-year-old Picea abies seedlings increases with soil horizon depth and canopy gap size. Canadian Journal of Forest Research, 37, 1236–1243.Google Scholar
de Chantal, M., Rita, H., Bergsten, U., and Grip, H. (2009). Frost heaving of Picea abies seedlings as influenced by soil preparation, planting technique, and location along gap-shelterwood gradients. Silva Fennica, 43, 39–50.Google Scholar
de Moraes, J. F., Volkoff, B. C. C. C., Cerri, C. C., and Bernoux, M. (1996). Soil properties under Amazon forest and changes due to pasture installation in Rondônia, Brazil. Geoderma, 70(1), 63–81.
DeFries, R. S., Foley, J. A., and Asner, G. P. (2004). Land-use choices: Balancing human needs and ecosystem function. Frontiers in Ecology and the Environment, 2(5), 249–257.Google Scholar
DeFries, R. S., Houghton, R. A., Hansen, M. C., Field, C. B., Skole, D., & Townshend, J. (2002). Carbon emissions from tropical deforestation and regrowth based on satellite observations for the 1980s and 1990s.Proceedings of the National Academy of Sciences, 99(22), 14256–14261.Google Scholar
DeFries, R., and Rosenzweig, C. (2010). Toward a whole-landscape approach for sustainable land use in the tropics. Proceedings of the National Academy of Sciences, 107(46), 19627–19632.Google Scholar
DeFries, R. S., Rudel, T., Uriarte, M., and Hansen, M. (2010). Deforestation driven by urban population growth and agricultural trade in the twenty-first century. Nature Geoscience, 3(3), 178–181.Google Scholar
DeGraff, J. V. (1979). Initiation of shallow mass movement by vegetative-type conversion. Geology, 7, 426–429.Google Scholar
DeGraff, V. J., Bryce, R., Jibson, R. W., Mora, S., and Rogers, C. T. (1989). Landslides: Their Extent and Significance in the Caribbean. In Brabb, E. E., and Harrold, B. L. (eds.), Landslides: Extent and Ecological Significance. Balkema, Rotterdam.
DeLonge, M. S. et al. (2006). Rain forest islands in the Chilean semiarid region: Fog-dependency. Ecosystem p4, 154–160.Google Scholar
DeLonge, M., D'Odorico, P., and Lawrence, D. (2008). Feedbacks between phosphorus deposition and canopy cover: The emergence of multiple stable states in tropical dry forests. Global Change Biology, 14(1), 154–160.Google Scholar
Del Pozo, A., Pérez, P., Gutiérrez, D., Alonso, A., Morcuende, R., and Martínez-Carrasco, R. (2007). Gas exchange acclimation to elevated CO2 in upper-sunlit and lower-shaded canopy leaves in relation to nitrogen acquisition and partitioning in wheat grown in field chambers. Environmental and Experimental Botany, 59(3), 371–380.Google Scholar
DeLucia, E. V., Drake, J. E., Thomas, R. B., and Gonzalez-Meler, M. (2007). Forest carbon use efficiency: Is respiration a constant fraction of gross primary production?Global Change Biology, 13(6), 1157–1167.Google Scholar
del-Val, E. et al. (2006). Rain forest islands in the Chilean semiarid region: Fog-dependency, ecosystem persistence and tree regeneration. Ecosystems, 9(4), 598–608.Google Scholar
Dentener, F. et al. (2006). Nitrogen and sulfur deposition on regional and global scales: A multimodel evaluation. Global Biogeochemical Cycles, 20(4).Google Scholar
Deo, R. C. (2011). Links between native forest and climate in Australia. Weather, 66, 64–69.Google Scholar
DeVries, J., and Toenniessen, G. H. (2001). Securing the harvest: Biotechnology, breeding, and seed systems for African crops. CABI.
Diamond, J. (2005). Collapse: How societies choose to fail or succeed. Penguin, New York.
Diaz, R. J., and Rosenberg, R. (2008). Spreading dead zones and consequences for marine ecosystems. Science, 321(5891), 926–929.Google Scholar
Diaz-Ravina, M., Acea, M. J., and Carballas, T. (1995). Seasonal changes in microbial biomass and nutrient flush in forest soils. Biology and Fertility of Soils, 19, 220–226.Google Scholar
Dickinson, R. E., and Henderson-Sellers, A. (1988). Modelling tropical deforestation: A study of GCM land-surface parameterizations. Q J Roy. Meteor. Soc., 114, 439–462.Google Scholar
Didia, D. O. (1997). Democracy, political instability and tropical deforestation. Global Environmental Change, 7(1), 63–76.Google Scholar
Dietrich, W. E., Reiss, R., Hsu, M. L., and Montgomery, D. (1995). A process-based model for colluvial soil depth and shallow landslides using digital elevation data. Hydrol. Processes, 9, 383–400.Google Scholar
Dirmeyer, P. A., and Brubaker, K. L. (2007). Characterization of the global hydrologic cycle from a back-trajectory analysis of atmospheric water vapor. J. Hydrometeorol., 8(1), 20–37, doi:10.1175/JHM557.1.Google Scholar
Dirzo, R., and Raven, P. H. (2003). Global state of biodiversity and loss. Annual Review of Environment and Resources, 28(1), 137–167.Google Scholar
Dixon, R. K., Solomon, A. M., Brown, S., Houghton, R. A., Trexier, M. C., and Wisniewski, J. (1994). Carbon pools and flux of global forest ecosystems. Science, 263(5144), 185–190.Google Scholar
Doak, D. F., Bigger, D., Harding, E. K., Marvier, M. A., O'malley, R. E., and Thomson, D. (1998). The statistical inevitability of stability-diversity relationships in community ecology. The American Naturalist, 151(3), 264–276.Google Scholar
D'Odorico, P., Carr, J. A., Laio, F., Ridolfi, L., and Vandoni, S. (2014). Feeding humanity through global food trade. Earth's Future, 2, doi:10.1002/2014EF000250.Google Scholar
D'Odorico, P., Caylor, K., Okin, G. S., and Scanlon, T. M. (2007). On soil moisture–vegetation feedbacks and their possible effects on the dynamics of dryland ecosystems. J. Geophys. Res., 112, G04010, doi:10.1029/2006JG000379.Google Scholar
D'Odorico, P., Engel, V., Carr, J. A., Oberbauer, S. F., Ross, M. S., and Sah, J. P. (2011). Tree-grass coexistence in the Everglades freshwater system. Ecosystems, 14, 298–310.Google Scholar
D'Odorico, P., and Fagherazzi, S. (2003). A probabilistic model of rainfall-triggered shallow landslides in hollows: A long-term analysis. Water Resour. Res., 39(9), 1262.Google Scholar
D'Odorico, P., Fuentes, J. D., Pockman, W. T., Collins, S. L., He, Y., Medeiros, J. S., De Wekker, S., and Litvak, M. E. (2010a) Positive feedback between microclimate and shrub encroachment in the northern Chihuahuan desert. Ecosphere, 1, art17, doi:10.1890/ES10-00073.1.Google Scholar
D'Odorico, P., He, Y., Collins, S., De Wekker, S. F., Engel, V., and Fuentes, J. D. (2013). Vegetation–microclimate feedbacks in woodland-grassland ecotones. Global Ecology and Biogeography, 22(4), 364–379.Google Scholar
D'Odorico, P., Laio, F., Porporato, A., Ridolfi, L., Rinaldo, A., and Rodriguez-Iturbe, I. (2010b). Ecohydrology of terrestrial ecosystems. Bioscience, 60(11), 898–907.Google Scholar
D'Odorico, P., Laio, F., and Ridolfi, L. (2006). A probabilistic analysis of fire-induced tree-grass coexistence in savannas. American Naturalist, 167(3), E79–E87.Google Scholar
D'Odorico, P., and Rulli, M. C. (2013). The fourth food revolution. Nat. Geosci., 6(6), 417–418, doi:10.1038/ngeo1842.Google Scholar
D'Odorico, P. and Rulli, M.C. (2014), The land and its people. Nature Geoscience, 4: 324–325.Google Scholar
Doerr, S. H., Shakesby, R. A., and Walsh, R. P. D. (2000). Soil water repellency: Its causes, characteristics and hydrogeomorphological significance. Earth Sci. Rev., 51, 33–65.Google Scholar
Don, A., Schumacher, J., and Freibauer, A. (2011). Impact of tropical land-use change on soil organic carbon stocks–a meta-analysis. Global Change Biology, 17(4), 1658–1670.Google Scholar
Dornburg, V. et al. (2010). Bioenergy revisited: Key factors in global potentials of bioenergy. Energy & Environmental Science, 3(3), 258–267.Google Scholar
Dosskey, M. G., and Bertsch, P. M. (1997). Transport of dissolved organic matter through a sandy forest soil. Soil Science Society of America Journal, 61(3), 920–927.Google Scholar
Drury, W. H. (1956). Bog Flats and Physiographic Processes in the Upper Kuskokwim River Region, Alaska, Contributions from the Gray Herbarium of Harvard University, No. CLXXVIII, p. 130.
Dubayah, R. O., Drake, J. B. (2000). Lidar remote sensing for forestry. Journal of Forestry, 98, 44–46.Google Scholar
Dubé, S., and Plamondon, A. P. (1995). Watering up after clearcutting on forested wetlands of the St. Lawrence lowland. Water Resour. Res., 31(7), 1741–1750.Google Scholar
Dublin, H. T., Sinclair, A. R. E., and McGlade, J. (1990). Elephants and fire as causes of multiple stable states in the Serengeti-Mara woodlands. Journal of Animal Ecology, 59, 1147–1164.Google Scholar
Dunham, A. E. (2008). Above and below-ground impacts of terrestrial mammals and birds in a tropical forest. Oikos, 117(4), 571–579.Google Scholar
Dunne, T. (1978). Field studies of hillsope flow processes. In Kirkby, M. J. (ed.), Hillslope Hydrology. J. Wiley & Sons, New York, pp. 227–293.
Dunne, T., and Black, R. D. (1970a). An experimental investigation of runoff prediction in permeable soils. Water Resour. Res., 6(2), 478–490.Google Scholar
Dunne, T., and Black, R. D. (1970b). Partial area contributions to storm runoff in a small New England watershed. Water Resour. Res., 6(5), 1296–1311.Google Scholar
Dunne, T., and Leopold, L. B. (1978). Water in Environmental Planning. W. H. Freeman and Co., New York.
Dutch, J., and Ineson, P. (1990). Denitrification of an upland forest site. Forestry, 63(4), 363–37Google Scholar
du Toit, J. T., Rogers, K. H., and Biggs, H. C. (2003). eds. The Kruger experience: Ecology and management of savanna heterogeneity. Island, Washington, DC.
Dutschke, M., and Wolf, R. (2007). Reducing emissions from deforestation and forest degradation in developing countries: The way forward. Foresty, Risk and Climate Policy, 125.Google Scholar
Dvořàk, K. A. (1992). Resource management by West African farmers and the economics of shifting cultivation. American Journal of Agricultural Economics, 74(3), 809–815.Google Scholar
Dy, G., and Payette, S. (2007). Frost hollows of the boreal forest as extreme environments for black spruce tree growth. Canadian Journal of Forest Research, 37, 492–504.Google Scholar
Dyrness, C. T. (1982). Control of depth to permafrost and soil temperature by the forest floor in black spruce/feathermoss communities, US For. Serv. Res. Note PNW-396;1–9 pp.
Easterling, D. R., Karl, T. R., Gallo, K. P., Robinson, D. A., Trenberth, K. E., and Dai, A. (2000). Observed climate variability and change of relevance to the biosphere. J. Geophys. Res., 105(D15), 20, 101–120, 114, doi:10.1029/2000JD900166.Google Scholar
Easterling, D. R., Peterson, T. C., and Karl, T. R. (1996). On the development and use of homogenized climate datasets. J. Clim., 9, 1429–1434.Google Scholar
Easterwood, G. W., and Sartain, J. B. (1990). Clover residue effectiveness in reducing orthophosphate sorption on ferric hydroxide coated soil. Soil Science Society of America Journal, 54(5), 1345–1350.Google Scholar
Eberbach, P. L. (2003). The eco-hydrology of partly cleared, native ecosystems in southern Australia: A review. Plant and Soil, 257(2), 357–369.Google Scholar
Eden, M. J., Furley, P. A., McGregor, D. F. M., Milliken, W., and Ratter, J. A., (1991). Effect of forest clearance and burning on soil properties in northern Roraima, Brazil. For. Ecol. Manage., 38, 283–290.CrossRefGoogle Scholar
Edmonds, R. L., and Lebo, D. S. (1998). Diversity, production, and nutrient dynamics of fungal sporocarps on logs in an oldgrowth temperate rain forest, Olympic National Park, Washington. Canadian Journal of Forest Research, 28, 665–673.Google Scholar
Egerton, J. J. G., Banks, J. C. G., Gibson, A., Cunningham, R. B., and Ball, Marilyn C. (2000). Facilitation of seedling establishment: Reduction in irradiance enchances winter growth of eucalypus pauciflora. Ecology, 81, 1437–1449.Google Scholar
Ellison, D., Futter, M. N., and Bishop, K. (2012). On the forest cover-water yield debate: From demand-to-supply-side thinking. Global Change Biology, 18, 806–820.Google Scholar
Elmqvist, T., Folke, C., Nyström, M., Peterson, G., Bengtsson, J., Walker, B., and Norberg, J. (2003). Response diversity, ecosystem change, and resilience. Frontiers in Ecology and the Environment, 1(9), 488–494.Google Scholar
Eltahir, E. A. B., and Bras, R. L. (1994). Precipitation recycling in the Amazon Basin. Quarterly Journal of the Royal Meteorological Society, 120, 861–880.Google Scholar
Eltahir, E. A. B., and Bras, R. L. (1996). Precipitation recycling. Rev. Geophys. 34(3), 367–379.Google Scholar
Elton, C. S. (1958). The Ecology of Invasions by Plants and Animals. Methuen, London, 18.
Engels, C., and Marschner, H. (1955). Plant uptake and utilization of nitrogen. In Bacon, P. E. (ed.), Nitrogen Fertilization in the Environment. Marcel Dekker, New York, pp. 41–81.
Ensminger, J. (1996). Making a market: The institutional transformation of an African society. Cambridge University Press, Cambridge.
Epstein, H. E. et al. (2004). The nature of spatial transitions in the Arctic. Journal of Biogeography, 31, 1917–1933.Google Scholar
Estes, J. A. et al. (2011). Trophic downgrading of planet Earth. Science, 333(6040), 301–306.Google Scholar
E. U. (2009). Directive 2009/28/EC of the European Parliament and of the Council of 23 April 2009 on the promotion of the use of energy from renewable sources and amending and subsequently repealing Directives 2001/77/EC and 2003/30/EC.
Eugster, W. et al. (2000). Land–atmosphere energy exchange in Arctic tundra and boreal forest: Available data and feedbacks to climate. Global Change Biology, 6, 84–115.Google Scholar
Eviner, V. T., and ChapinIII, F. S. (2003). Functional matrix: A conceptual framework for predicting multiple plant effects on ecosystem processes. Annu. Rev. Ecol. Evol. Syst., 34, 455–485.Google Scholar
Ewel, J., Berish, C., Brown, B., Price, N., and Raich, J. (1981). Slash and burn impacts on a Costa Rican wet forest site. Ecology, 816–829.Google Scholar
Ewers, R. M., Laurance, W. F., and Souza, C. M. (2008). Temporal fluctuations in Amazonian deforestation rates. Environmental Conservation, 35(04), 303–310.Google Scholar
Fairhead, J., Leach, M., and Scoones, I. (2012). Green grabbing: A new appropriation of nature?Journal of Peasant Studies 39(2), 237–261.Google Scholar
Fairman, J. G., Jr., Nair, U. S., Christopher, S. A., and Mölg, T. (2011). Land use change impacts on regional climate over Kilimanjaro. J. Geophys. Res., 116, D03110, doi:10.1029/2010JD014712.Google Scholar
Falkenmark, M. J., and Rockström, J. (2004). Balancing Water for Humans and Nature: The New Approach in Ecohydrology. Earthscan, London.
Falkenmark, M., and Rockstrom, J. (2006). The new blue and green water paradigm: Breaking new ground for water resources planning and management. J. Water Resour. Plann. Manage., 132(3), 129–132.Google Scholar
FAO. (1982). Tropical forest resource, by J.P. Lanly. FAO Forestry Paper No. 30. Rome.
FAO. (1995). Forest Resources Assessment 1990. Global Synthesis. FAO Forestry Paper No. 124. FAO, Rome.
FAO. (1996). Forest resources assessment 1990: Survey of tropical forest cover and study of change processes. FAO Forestry Paper 130, Rome.
FAO. (2001). Global forest resources assessment 2000. FRA 2000 Main Report. FAO Forestry Paper 140, Rome.
FAO. (2005). FAOSTAT database. Rome, Italy: FAO.
FAO (2008). FAOSTAT – Production – ForesSTAT, available at: http://faostat.fao.org/DesktopDefault.aspx?PageID=381&lang=en.
FAO. (2009). State of the world's forests. Rome: FAO.
FAO. (2010). Global forest resource assessment 2010: Key findings, available at: http://www.fao.org/forestry/fra/fra2010/en/.
FAO. (2011a). State of the world's forests. Rome: FAO.
FAO. (2011b). Looking ahead in world food and agriculture: Perspectives to 2050. Edited by Piero Conforti. Agricultural Development Economics Division Economic and Social Development Department Food and Agriculture Organization of the United Nations 2011, Paris.
FAO. (2011c). Voluntary guidelines on the responsible governance of tenure of land and other natural resources (Zero Draft, Released 18 April 2011). Rome: Food and Agriculture Organization of the United Nations. www.fao.org/fileadmin/user_upload/nr/land_tenure/econsultation/english/Zero_Draft_VG_Final.pdf.
FAO. (2012). State of the world's forests. Rome: FAO.
FAO and JRC.(2012). Global forest land-use change 1990–2005, by E. J. Lindquist, R. D'Annunzio, A. Gerrand, K. MacDicken, F. Achard, R. Beuchle, A. Brink, H. D. Eva, P. Mayaux, J. San-Miguel-Ayanz & H-J. Stibig. FAO Forestry Paper No. 169. Food and Agriculture Organization of the United Nations and European Commission Joint Research Centre. Rome, FAO.
FAOSTAT. (2011). Rome, FAO, available at: http://www.faostat.fao.org.proxy1.library.jhu.edu/default.aspx.
FAO/UNEP. (1981). Tropical Forest Resources Assessment Project. FAO, Rome.
FAO-UNESCO. (2006). Soil Map of the World, digitized by ESRI. Soil climate map, USDA-NRCS, Soil Science Division, World Soil Resources, Washington DC. Soil Pedon database, USDA-NRCS National Soil Survey Center, Lincoln, NE.
Fargione, J., Hill, J., Tilman, D., Polasky, S., and Hawthorne, P. (2008). Land clearing and the biofuel carbon debt. Science, 319, 1235–1238.Google Scholar
Farzin, Y. H. (1984). The effect of the discount rate on depletion of exhaustible resources. Journal of Political Economy, 841–851.Google Scholar
Fearnside, P. M. (2005). Deforestation in Brazilian Amazonia: History, rates, and consequences. Conservation Biology, 19, 680–688, doi:10.1111/j.1523-1739.2005.00697.x.Google Scholar
Fearnside, P. M. (2007). Deforestation in Amazonia. In Cleveland, C. J. (ed.), Encyclopedia of Earth. Washington, DC: Environmental Information Coalition, National Council of Science and the Environment, available at: www.eoearth.org/article/Deforestation_in_Amazonia.
Fearnside, P. M., and Imbrozio Barbosa, R. (1998). Soil carbon changes from conversion of forest to pasture in Brazilian Amazonia. Forest Ecology and Management, 108(1), 147–166.Google Scholar
Feller, C., and Beare, M. H. (1997). Physical control of soil organic matter dynamics in the tropics. Geoderma, 79(1), 69–116.Google Scholar
Feller, I. C., Lovelock, C. E., Berger, U., McKee, K. L., Joye, S. B., and Ball, M. C. (2010). Biocomplexity in mangrove ecosystems. Annual Review of Marine Science, 2, 395–417.Google Scholar
Ferraz, S. F.d., Vettorazzi, C. A., Theobald, D. M., and Ballester, M. A. R. (2005). Landscape dynamics of Amazonian deforestation between 1984 and 2002 in central Rondonia, Brazil: Assessment and future scenarios. Forest Ecology and Management, 204, 67–83.Google Scholar
Field, C. B., van der Werf, G. R., and Shen, S. S. P. (2009). Human amplification of drought induced biomass burning in Indonesia since 1960. Nature Geoscience, 2, 185–188.Google Scholar
Finér, L., Mannerkoski, H., Piirainen, S., and Starr, M. (2003). Carbon and nitrogen pools in an old-growth, Norway spruce mixed forest in eastern Finland and changes associated with clear-cutting. Forest Ecology and Management, 174(1), 51–63.Google Scholar
Fisch, G. et al. (2004). The convective boundary layer over pasture and forest in Amazonia. Theoretical and Applied Climatology, 78(1–3), 47–59.Google Scholar
Fischer, J., Lindenmayer, D. B., and Manning, A. D. (2006). Biodiversity, ecosystem function, and resilience: Ten guiding principles for commodity production landscapes. Frontiers in Ecology and the Environment, 4(2), 80–86.Google Scholar
Fisher, B. (2010). African exception to drivers of deforestation. Nature Geoscience, 3(6), 375–376.Google Scholar
Fisher, J. B., Badgley, G., and Blyth, E. (2012). Global nutrient limitation in terrestrial vegetation. Global Biogeochemical Cycles, 26(3).Google Scholar
Fisher, R. A., Williams, M., Lola Da Costa, A., Malhi, Y., Da Costa, R. F., Almeida, S., and Meir, P. (2007). The response of an E. Amazonian rain forest to drought stress: Results and modelling analyses from a throughfall exclusion experiment. Global Change Biol., 13, 2361–2378.Google Scholar
Fitzherbert, E. B. et al. (2008). How will oil palm expansion affect biodiversity?Trends Ecol. Evol., 23, 538–545.Google Scholar
Flannery, T. (2002). The future eaters: An ecological history of the Australasian lands and people. Grove Press.
Flint, E. P. (1994). Changes in land use in South and Southeast Asia from 1880 to 1980: A data base prepared as part of a coordinated research program on carbon fluxes in the tropics. Chemosphere, 29(5), 1015–1062.Google Scholar
Foley, J. A., Kutzbach, J. E., Coe, M. T., and Levis, S. (1994). Feedbacks between climate and boreal forests during the Holocene epoch. Nature, 371, 52–55.Google Scholar
Foley, J. A., Prentice, I. C., Ramankutty, N., Levis, S., Polard, D, Sitch, S., and Haxeltine, A. (1996). An integrated biosphere model of land surface processes, terrestrial carbon balance and vegetation dynamics. Global Biogeophysical Cycle, 10, 603–629.Google Scholar
Foley, J. A. et al. (2007). Amazonia revealed: Forest degradation and loss of ecosystem goods and services in the Amazon Basin. Frontiers in Ecology and the Environment, 5(1), 25–32.Google Scholar
Foley, J. A. et al. (2011). Solutions for a cultivated planet. Nature, 478, 337–342, doi:10.1038/nature10452.Google Scholar
Folke, C., Carpenter, S., Walker, B., Scheffer, M., Elmqvist, T., Gunderson, L., and Holling, C. S. (2004). Regime shifts, resilience, and biodiversity in ecosystem management. Annu. Rev. Ecol. Evol. Syst., 35, 557–581.Google Scholar
Folland, C. K. et al. (2001). Observed Climate Variability and Change. In Houghton, J. T. (ed.), Climate Change 2001: The Scientific Basis. Cambridge University Press, New York, pp. 99–181.
Folliott, P. F., and Thrud, D. B. (1977). Water resources and multiple-use forestry in the Southwest. J. For., 75, 469–472.Google Scholar
Food and Agriculture Organization (FAO) (2002). World Agriculture: Towards 2015/2030. Food and Agriculture Organization of the United Nations, Rome.
Food and Agriculture Organization (FAO) (2007). FAOSTAT Online Statistical Service. United Nations Food and Agriculture Organization (FAO), Rome, available at: http://faostat.fao.org (accessed October 2007).
Food and Agriculture Organization (FAO). (2010). Global Forest Resources Assessment 2010: Main Report. United Nations Food and Agriculture Organization, Rome.
Foster, P. (2001). The potential negative impacts of global climate change on tropical montane cloud forests. Earth-Science Reviews, 55, 73–106.Google Scholar
Fraedrich, K., Kleidon, A., and Lunkeit, F. (1999). A green planet versus a desert world: Estimating the effect of vegetation extremes on the atmosphere. J. Clim. 12, 3156–3163.Google Scholar
Fredriksen, R. (1975). Nitrogen, phosphorus and particulate matter budgets of five coniferous forest ecosystems in the western Cascades Range, Oregon, Ph.D. dissertation, Oregon State University, Corvallis.
Freeze, R. A. (1972). Role of subsurface flow in generating surface runoff. 2. Upstream source areas. Water Resour. Res., 8(5), 1272–1283, doi:10.1029/WR008i005p01272.Google Scholar
Froment, A. (2009). Biodiversity and health: The place of parasitic and infectious diseases. In Sala, O. E., Meyerson, L. A., and Parmesan, C. (eds.), Biodiversity Change and Human Health. Island Press, Washington, DC, pp. 211–227.
Fujiwara, K. (1970). A study on landslides by aerial photographs. Res. Bull. Exp. Forests Hakkaido University, 27(2), 297–345.Google Scholar
Fuller, D. O. (2006). Tropical forest monitoring and remote sensing: A new era of transparency in forest governance?Singapore Journal of Tropical Geography, 27(1), 15–29.Google Scholar
Galindo-Leal, C., and de Gusmao Camara, I. (2003). The Atlantic Forest of South America: Biodiversity Status, Threats, and Outlook. Island Press, Washington, DC.
Galloway, J. N., and Cowling, E. B. (2002). Reactive nitrogen and the world: 200 years of change. AMBIO: A Journal of the Human Environment, 31(2), 64–71.Google Scholar
Galloway, J. N., Schlesinger, W. H., Levy, H., Michaels, A., and Schnoor, J. L. (1995). Nitrogen fixation: Anthropogenic enhancement-environmental response. Global Biogeochemical Cycles, 9(2), 235–252.Google Scholar
Galloway, J. N. et al. (2004). Nitrogen cycles: Past, present, and future. Biogeochemistry, 70(2), 153–226.Google Scholar
Galloway, J. N. et al. (2008). Transformation of the nitrogen cycle: Recent trends, questions, and potential solutions. Science, 320(5878), 889–892.Google Scholar
Garcia-Montiel, D. C., Neill, C., Melillo, J., Thomas, S., Steudler, P. A., and Cerri, C. C. (2000). Soil phosphorus transformations following forest clearing for pasture in the Brazilian Amazon.
García-Oliva, F., Casar, I., Morales, P., and Maass, J. M. (1994). Forest-to-pasture conversion influences on soil organic carbon dynamics in a tropical deciduous forest. Oecologia, 99(3–4), 392–396.Google Scholar
Gash, J. H. C., and Nobre, C. A. (1997). Climatic effects of Amazonian deforestation: Some results from ABRACOS. Bull. American Meteorol. Soc., 78(5), 823–830.Google Scholar
Gebremedhin, B., and Schwab, G. (1998). The economic importance of crop rotation systems: Evidence from the literature (No. 11690). Michigan State University, Department of Agricultural, Food, and Resource Economics, Lansing.
Geiger, R. (1965). The Climate near the Ground. Harvard University Press, Cambridge, MA.
Geist, H. J., and Lambin, E. F. (2002). Proximate causes and underlying driving forces of tropical deforestation: Tropical forests are disappearing as the result of many pressures, both local and regional, acting in various combinations in different geographical locations. BioScience, 52(2), 143–150.Google Scholar
Gelfand, I., Zenone, T., Jasrotia, P., Chen, J., Hamilton, S. K., and Robertson, G. P. (2011). Carbon debt of Conservation Reserve Program (CRP) grasslands converted to bioenergy production. Proceedings of the National Academy of Sciences, 108(33), 13864–13869.Google Scholar
Gerber, S., Hedin, L. O., Oppenheimer, M., Pacala, S. W., and Shevliakova, E. (2010). Nitrogen cycling and feedbacks in a global dynamic land model. Global Biogeochemical Cycles, 24(1).Google Scholar
Germino, M. J., and Smith, W. K. (1999). Sky exposure, crown architecture, and low-temperature photoinhibition in conifer seedlings at alpine treeline. Plant, Cell and Environment, 22, 407–415.Google Scholar
Germino, M. J., and Smith, W. K. (2000). Differences in microsite, plant form, and low-temperature photoinhibition in Alpine plants. Arctic, Antarctic, and Alpine Research, 32, 388–396.Google Scholar
Germino, M. J., Smith, W. K., and Resor, A. C. (2002). Conifer seedling distribution and survival in an alpine-treeline ecotone. Plant Ecology, 162, 157–168.Google Scholar
Ghassemi, F., Jakeman, A. J., and Nix, H. A. (1995). Salinization of Land and Water Resources: Human Causes, Extent, Management and Case Studies. CABI, Canberra.
Giardina, C. P., SanfordJr., R. L., Dockersmith, I. C., and Jaramillo, V. J. (2000). The effects of slash and burning on ecosystem nutrients during the land preparation phase of shifting cultivation. Plant and Soil, 220, 247–260.Google Scholar
Gibbs, H. K., Brown, S., Niles, J. O., and Foley, J. A. (2007). Monitoring and estimating tropical forest carbon stocks: Making REDD a reality. Environmental Research Letters, 2(4), 045023.Google Scholar
Gibbs, H. K., Ruesch, A. S., Achard, F., Clayton, M. K., Holmgren, P., Ramankutty, N., and Foley, J. A. (2010). Tropical forests were the primary sources of new agricultural land in the 1980s and 1990s. Proceedings of the National Academy of Sciences, 107(38), 16732–16737.Google Scholar
Gibson, C. C., Williams, J. T., and Ostrom, E. (2005). Local enforcement and better forests. World Development, 33(2), 273–284.Google Scholar
Gibson, L. et al. (2011). Primary forests are irreplaceable for sustaining tropical biodiversity. Nature, 478(7369), 378–381.Google Scholar
Giller, K. E., Beare, M. H., Lavelle, P., Izac, A. M., and Swift, M. J. (1997). Agricultural intensification, soil biodiversity and agroecosystem function. Applied Soil Ecology, 6(1), 3–16.Google Scholar
Glade, T. (2003). Landslide occurrence as a response tol and use change: A review of evidence from New Zealand. Catena, 51, 297–314.Google Scholar
Global Forest Watch. (2000). Canada's Forests at a Crossroads: An Assessment in the Year 2000. World Resources Institute, Washington, DC.
Godfray, H. C. J. et al. (2010). Food security: The challenge of feeding 9 billion people. Science, 327, 812–818.Google Scholar
Godfray, H. C. (2011). Food for thoughts. Proc. Natl. Acad. Sci. USA, 108, 19845–6.Google Scholar
Godfray, H. C. J., and Garnett, T. (2014). Food security and sustainable intensification. Philosophical Transactions of the Royal Society B: Biological Sciences, 369(1639), 20120273.Google Scholar
Goeschl, T., and Swanson, T. (2002). The social value of biodiversity for RandD. Environ. Resource Econ. 22, 477–504.Google Scholar
Goetz, S., and Dubayah, R. (2011). Advances in remote sensing technology and implications for measuring and monitoring forest carbon stocks and change. Carbon Management, 2(3), 231–244.Google Scholar
Goetz, S. J. et al. (2009). Mapping and monitoring carbon stocks with satellite observations: A comparison of methods. Carbon Balance and Management, 4(1), 2.Google Scholar
GOFC-GOLD (2011). A sourcebook of methods and procedures for monitoring and reporting anthropogenic greenhouse gas emissions and removals caused by deforestation, gains and losses of carbon stocks in forests remaining forests, and forestation. GOFC-GOLD Report Version COP17-1. GOFC-GOLD Project Office, Natural Resources Canada, Alberta.
Goldewijk, K. K. (2001). Estimating global land use change over the past 300 years: The HYDE database. Global Biogeochemical Cycles, 15(2), 417–433.Google Scholar
Goldewijk, K., Beusen, A., Van Drecht, G., and De Vos, M. (2011). The HYDE 3.1 spatially explicit database of human-induced global land-use change over the past 12,000 years. Global Ecology and Biogeography, 20(1), 73–86.Google Scholar
Goll, D. S. et al. (2012). Nutrient limitation reduces land carbon uptake in simulations with a model of combined carbon, nitrogen and phosphorus cycling. Biogeosciences Discussions, 9(3), 3173–3232.Google Scholar
Gordon, L. J., Dunlop, M., and Foran, B. (2003). Land cover change and water vapour flows: Learning from Australia. Philos. Trans. R. Soc. London, Ser. B, 358, 1973–1984.Google Scholar
Gordon, L. J., Steffen, W., Jönsson, B. F., Folke, C., Falkenmark, M., and Johannessen, Å. (2005). Human modification of global water vapor flows from the land surface. Proceedings of the National Academy of Sciences of the United States of America, 102(21), 7612–7617.Google Scholar
Gosz, J. R. (1981). Nitrogen cycling in coniferous ecosystems. Ecological Bulletins, Sweden.
Grace, J. (1989). Tree lines. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 324, 233–245.Google Scholar
Grace, J., Allen, S. J., and Wilson, C. (1989). Climate and the meristem temperatures of plant communities near the tree-line. Oecologia, 79, 198–204.Google Scholar
Grau, H. R., Gasparri, N. I., and Aide, T. M. (2005). Agriculture expansion and deforestation in seasonally dry forests of north-west Argentina. Environmental Conservation, 32(02), 140–148.Google Scholar
Greeley, W. B. (1925). The relation of geography to timber supply. Economic Geography, 1, 1–11.Google Scholar
Greene, R. S. B. (1992). Soil physical-properties of three geomorphic zones in a semiarid Mulga woodland. Aust. J. Soil Res., 30(1), 55–69.Google Scholar
Greene, R. S. B., Kinnell, P. I. A., and Wood, J. T. (1994). Role of plant cover and stock trampling on runoff and soil-erosion from semiarid wooded rangelands. Aust. J. Soil Res., 32(5), 953–973.Google Scholar
Greene, R. S. B., Valentin, C., and Esteves, M. (2001). Runoff and erosion processes, in Banded Vegetation Patterning in Arid and Semiarid Environments: Ecological Processes and Consequences for Management. Ecol. Stud., 149, 52–76.Google Scholar
Grey, M. J., Clarke, M. F., and Loyn, R. H. (1998). Influence of the Noisy Miner Manorina melanocephala on avian diversity and abundance in remnant Grey Box woodland. Pacific Conservation Biology, 4(1), 55.Google Scholar
Griffin, R. C., and McCarl, B. A. (1989). Brushland management for increased water yield in Texas. Journal of the American Water Resources Association, 25, 175–186. DOI: 10.1111/j.1752-1688.1989.tb05679.x
Grimmond, C. S. B., Robeson, S. M., and Schoof, J. T. (2000). Spatial variability of micro-climatic conditions within a mid-latitude deciduous forest. Climate Research, 15, 137–149.Google Scholar
Groot, A., and Carlson, D. W. (1996). Influence of shelter on night temperatures, frost damage, and bud break of white spruce seedlings. Canadian Journal of Forest Research, 26, 1531–1538.Google Scholar
Gruber, N., and Galloway, J. N. (2008). An Earth-system perspective of the global nitrogen cycle. Nature, 451(7176), 293–296.Google Scholar
Guerra, C. A., Snow, R. W., and Hay, S. I. (2006). A global assessment of closed forests, deforestation and malaria risk. Ann. Trop. Med. Parasitol., 100(3), 189–204.Google Scholar
Gutierrez, A. G., Barbosa, O., Christie, D. A., del-Val, E. K., Ewing, H. A., Jones, C. G., ... & Armesto, J. J. (2008). Regeneration patterns and persistence of the fog-dependent Fray Jorge forest in semiarid Chile during the past two centuries.Global Change Biology, 14(1), 161–176.Google Scholar
Guo, Z. W., Xiao, X. M., and Li, D. M. (2000). An assessment of ecosystem services: Water flow regulation and hydroelectric power production. Ecol. Appl., 10, 925–936.Google Scholar
Gupta, R. K., and Abrol, I. P. (2000). Salinity build-up and changes in the rice-wheat system of the Indo-Gangetic Plains. Exp. Agric., 36, 273–84.Google Scholar
Gurr, G. M., Wratten, S. D., and Luna, J. M. (2003). Multi-function agricultural biodiversity: Pest management and other benefits. Basic and Applied Ecology, 4(2), 107–116.Google Scholar
Guthrie, R. H. (2002). The effects of logging on frequency and distribution of landslides in three watersheds on Vancouver Island, British Columbia. Geomorphology, 43(3), 273–292.Google Scholar
Gutierrez, A. G., Barbosa, O., and Christie, D. A. (2008). DE Lequency and distribution of landslides in three watersheds on Vancouver Island, British Columbia. Geomorphology, 43(3), 273–292.Google Scholar
Haberl, H., Beringer, T., Bhattacharya, S. C., Erb, K. H., and Hoogwijk, M. (2010). The global technical potential of bio-energy in 2050 considering sustainability constraints. Current Opinion in Environmental Sustainability, 2(5), 394–403.Google Scholar
Haberl, H., Erb, K. H., Krausmann, F., Running, S., Searchinger, T. D., and Smith, W. K. (2013). Bioenergy: How much can we expect for 2050?Environmental Research Letters, 8(3), 031004.Google Scholar
Haigh, M. J., Rawat, J. S., Rawat, M. S., Bartarya, S. K., and Rai, S. P. (1995). Interactions between forest and landslide activity along new highways in the Kumaun Himalaya. Forest Ecology and Management, 78(1), 173–189.Google Scholar
Halladay, K., Malhi, Y., and New, M. (2012). Cloud frequency climatology at the Andes/Amazon transition. 2. Trends and variability. J. Geophys. Res., 117, D23103, doi:10.1029/2012JD017789.Google Scholar
Hammond, D. S., and ter Steege, H. (1998). Propensity for fire in Guianan rainforests. Conserv. Biol. 12, 944–947.Google Scholar
Hannam, K. D., Quideau, S. A., and Kishchuk, B. E. (2007). The microbial communities of aspen and spruce forest floors are resistant to changes in litter inputs and microclimate. Applied Soil Ecology 35, 635–647.Google Scholar
Hansen, M. C., Stehman, S. V., Potapov, P. V., Arunarwati, B., Stolle, F., and Pittman, K. (2009). Quantifying changes in the rates of forest clearing in Indonesia from 1990 to 2005 using remotely sensed data sets. Environmental Research Letters, 4(3), 034001.Google Scholar
Hansen, M. C. et al. (2008). Humid tropical forest clearing from 2000 to 2005 quantified by using multitemporal and multiresolution remotely sensed data. Proceedings of the National Academy of Sciences, 105(27), 9439–9444.Google Scholar
Hansen, M. C. et al. (2013). High-resolution global maps of 21st-century forest cover change. Science, 342(6160), 850–853.Google Scholar
Hanson, J. D., Liebig, M. A., Merrill, S. D., Tanaka, D. L., Krupinsky, J. M., and Stott, D. E. (2007). Dynamic cropping systems. Agronomy Journal, 99(4), 939–943.Google Scholar
Hargrave, J., and Kis-Katos, K. (2013). Economic causes of deforestation in the Brazilian Amazon: A panel data analysis for the 2000s. Environmental and Resource Economics, 54(4), 471–494.Google Scholar
Harr, R. D., and McCorison, F. M. (1979). Initial effects of clearcut logging on size and timing of peak flows in a small watershed in western Oregon, Forest Service Res. Paper PNW-249 Pacific Northwest Forest and Rangeland Experimental Station.
Harrison, A. F. (1987), Mineralisation of organic phosphorus in relation to soil factors, determined using isotopic 32P labeling. In Rowland, A. P. (eds.), Chemical Analysis in Environmental Research. NERC/ITE, Abbotts Ripton, pp. 84–87.
Hart, S. C. (1999). Nitrogen transformations in fallen tree boles and mineral soil of an old-growth forest. Ecology, 80(4), 1385–1394.Google Scholar
Hartwick, J. M. (1978). Substitution among exhaustible resources and intergenerational equity. The Review of Economic Studies, 45(2), 347–354.Google Scholar
Harvey, C. A. et al. (2008). Integrating agricultural landscapes with biodiversity conservation in the Mesoamerican hotspot. Conserv. Biol., 22, 8–15.Google Scholar
Hasler, N., Werth, D., and Avissar, R. (2009). Effects of tropical deforestation on global hydroclimate: A multimodel ensemble analysis. J. Clim. 22, 1124–1141.Google Scholar
Hauggaard-Nielsen, H., Ambus, P., and Jensen, E. S. (2001). Interspecific competition, N use and interference with weeds in pea–barley intercropping. Field Crops Research, 70(2), 101–109.Google Scholar
Hauggaard-Nielsen, H., and Jensen, E. S. (2005). Facilitative root interactions in intercrops. Plant and Soil, 274(1–2), 237–250.Google Scholar
Hayes, T. M. (2006). Parks, people, and forest protection: An institutional assessment of the effectiveness of protected areas. World Development, 34(12), 2064–2075.Google Scholar
Hazell, P., and Wood, S. (2008). Drivers of change in global agriculture. Philosophical Transactions of the Royal Society B: Biological Sciences, 363(1491), 495–515.Google Scholar
He, Y., De Wekker, S. F., Fuentes, J. D., and D'Odorico, P. (2011). Coupled land-atmosphere modeling of the effects of shrub encroachment on nighttime temperatures. Agric. For. Meteorol., 151, 1690–1697, doi:10.1016/j.agrformet.2011.07.005.Google Scholar
He, Y., D'Odorico, P., DeWekker, S. (2015).The role of vegetation-microclimate feedback in promoting shrub encroachment in the northern Chihuahuan desert. Global Change Biology, doi:10.1111/gcb.12856.Google Scholar
Heath, J. A., and Huebert, B. J. (1999). Cloudwater deposition as a source of fixed nitrogen in a Hawaiian montane forest. Biogeochemistry, 44, 119–134.Google Scholar
Hecht, S. B. (2005). Soybeans, development and conservation on the Amazon frontier. Development and Change, 36(2), 375–404.Google Scholar
Henderson-Sellers, A., and Gornitz, V. (1984). Possible climatic impacts of land cover transformations, with particular emphasis on tropical deforestation. Clim. Change, 6, 231–257.Google Scholar
Henrot, J., and Robertson, G. P. (1994). Vegetation removal in two soils of the humid tropics: Effect on microbial biomass. Soil Biol. Biochem., 26(1), 111–116.Google Scholar
Hermele, K. (2014). The Appropriation of Ecological Space. Agrofuels, unequal exchange and environmental load displacements, Rutledge, New York, pp. 158.
Hese, S. et al. (2005). Global biomass mapping for an improved understanding of the CO2 balance – the Earth observation mission Carbon-3D. Remote Sensing of Environment, 94(1), 94–104.Google Scholar
Hewlett, J. D. (1969). Principles of Forest Hydrology. University of Georgia University Press, Athens.
Hietz, P., Turner, B. L., Wanek, W., Richter, A., Nock, C. A., and Wright, S. J. (2011). Long-term change in the nitrogen cycle of tropical forests. Science, 334(6056), 664–666.Google Scholar
Hillel, D. (2000). Salinity Management for Sustainable Irrigation: Integrating Science, Environment and Economics. The World Bank, Washington, DC.
Hirota, M., Holmgren, M., Van Nes, E. H., and Scheffer, M. (2011). Global resilience of tropical forest and savanna to critical transitions. Science, 334(6053), 232–235.Google Scholar
Hobbie, E. A., Weber, N. S., and Trappe, J. M. (2001). Mycorrhizal vs saprotrophic status of fungi: The isotopic evidence. New Phytologist, 150(3), 601–610.Google Scholar
Hobbie, S. E. (1992). Effects of plant species on nutrient cycling. Tree, 7(10), 336–339.Google Scholar
Hobbie, S. E., and Vitousek, P. M. (2000). Nutrient limitation of decomposition in Hawaiian forests. Ecology, 81(7), 1867–1877.Google Scholar
Hobson, K. A., Bayne, E. M., & Van Wilgenburg, S. L. (2002). Large-scale conversion of forest to agriculture in the boreal plains of Saskatchewan.Conservation Biology, 16(6), 1530–1541.Google Scholar
Hoekman, D. H. (1997). Radar monitoring system for sustainable forest management in Indonesia IGARSS ‘97: 1997 Int. Geoscience and Remote Sensing Symposium (Singapore, 3–8 August 1997) pp. 1731–1733, doi:10.1109/IGARSS.1997.609048.
Hoffman, W. A., Schroeder, W., and Jackson, R. B. (2003). Regional feedbacks among fire, climate and tropical deforestation. Journal of Geophysical Research, 108(D23), 4721, doi:10.1029/2003JD003494.Google Scholar
Holden, S. T. (1993). Peasant household modelling: Farming systems evolution and sustainability in northern Zambia. Agricultural Economics, 9(3), 241–267.Google Scholar
Holder, C. D. (2003). Fog precipitation in the Sierra de las Minas Biosphere Reserve, Guatemala. Hydrol. Process., 17, 2001–2010.Google Scholar
Holder, C. D. (2006). The hydrological significance of cloud forests in the Sierra de las Minas Biosphere Reserve, Guatemala. Geoforum, 37(1), 82–93.Google Scholar
Holling, C. S. (1973). Resilience and stability of ecological systems. Annual Review of Ecology and Systematics, 4, 1–23.Google Scholar
Hong, Y., Adler, R., and Huffman, G. (2006). Evaluation of the potential of NASA multi-satellite precipitation analysis in global landslide hazard assessment. Geophysical Research Letters, 33, L22402, doi:10.1029/2006GL028010.Google Scholar
Hooper, D. U. et al. (2012). A global synthesis reveals biodiversity loss as a major driver of ecosystem change. Nature, 486(7401), 105–108.Google Scholar
Hopkins, A. D. (1938), Bioclimatics: A science of life and climate relations. U.S. Dept. of Agriculture.
Horsthemke, W., and Léfèver, R. (1984). Noise-Induced Transitions: Theory and Applications in Physics, Chemistry, and Biology. Springer, Berlin.
Hosonuma, N. et al. (2012). An assessment of deforestation and forest degradation drivers in developing countries. Environmental Research Letters, 7(4), 044009.Google Scholar
Houghton, R. A. (1996). Land-use change and terrestrial carbon: The temporal record. In Apps, M. J., and Price, D. T. (eds.), Forest Ecosystems, Forest Management and the Global Carbon Cycle. NATO ASI Series I, Vol. 40, Springer-Verlag, Berlin, Heidelberg, pp. 117–134.
Houghton, R. A. (1999). The annual net flux of carbon to the atmosphere from changes in land use 1850–1990. Tellus B, 51(2), 298–313.Google Scholar
Houghton, R. A. (2002). Temporal patterns of land-use change and carbon storage in China and tropical Asia. Science in China Series C Life Sciences-English Edition, 45(Suppl), 10–17.Google Scholar
Houghton, R. A. (2003). Revised estimates of the annual net flux of carbon to the atmosphere from changes in land use and land management 1850–2000. Tellus B, 55, 378–390, doi:10.1034/j.1600-0889.2003.01450.x.Google Scholar
Houghton, R. A. (2005). Aboveground forest biomass and the global carbon balance. Global Change Biology, 11(6), 945–958.Google Scholar
Houghton, R. A., Lefkowitz, D. S., and Skole, D. L. (1991). Changes in the landscape of Latin America between 1850 and 1985. I. Progressive loss of forests. For. Ecol. Manag., 38, 143–72.Google Scholar
Hudak, A. T. et al. (2012). Quantifying aboveground forest carbon pools and fluxes from repeat LiDAR surveys. Remote Sensing of Environment, 123, 25–40.Google Scholar
Huenneke, L. F., Hamburg, S. P., Koide, R., Mooney, H. A., and Vitousek, P. M. (1990). Effects of soil resources on plant invasion and community structure in Californian serpentine grassland. Ecology, 71, 478–491.Google Scholar
Hughes, M. A., and Dunn, M. A. (1996).The molecular biology of plant acclimation to low temperature. Journal of Experimental Botany, 47, 291–305.Google Scholar
Hulm, S. C., and Killham, K. (1988). Gaseous nitrogen losses from soil under Sitka spruce following the application of fertilizer 15N urea. Journal of Soil Science, 39(3), 417–424.Google Scholar
Hurtt, G. C. et al. (2006). The underpinnings of land-use history: Three centuries of global gridded land-use transitions, wood-harvest activity, and resulting secondary lands. Global Change Biology, 12(7), 1208–1229.Google Scholar
Hyde, W. F., Belcher, B. M., and Xu, J. (eds.). (2003). China's forests: Global lessons from market reforms.Resources for the Future. and CIFOR, Washington, DC.Google Scholar
IEA. (2009). Bioenergy: A Sustainable and Reliable Energy Source. Main Report. Paris: International Energy Agency.
IEA. (2011). Biofuels for Transport. International Energy Agency, Paris.
IEA. (2012). World Energy Outlook. The International Energy Agency, Paris.
Ikerd, J. E. (1991). A decision support system for sustainable farming. Northeastern J. Agricult. Resource Economics, 20, 109–113.Google Scholar
Ingebo, P. A. (1971). Suppression of channel-side chaparral cover increases streamflow. J. Soil Water Conserv., 26, 79–81.Google Scholar
INPE (2014). Annual Amazonian deforestation rates, available at: http://www.obt.inpe.br/prodes/index.php (accessed on March 23, 2015).
International Groundwater Resources Assessment Centre (IGARC). (2015). Saline and brackish groundwater by genesis, Available at: http://www.un-igrac.org/publications/344 (accessed on June 25, 2015).
IPCC. (2000). Land-use, land-use change and forestry: Summary for policymakers, available at: http://www.ipcc.ch/ipccreports/sres/land_use/index.php?idp=1 (accessed on June 25, 2015).
IPCC.(2014). Summary for Policymakers. InField, C. B. (eds.), Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge and New York, pp. 1–32.
IUCN. (2014). The IUCN red list of threatened species, available at: http://www.iucnredlist.org. (Accessed on June 1, 2015).
Jactel, H., and Brockerhoff, E. G. (2007). Tree diversity reduces herbivory by forest insects. Ecology Letters, 10(9), 835–848.Google Scholar
Jakob, M. (2000). The impacts of logging on landslide activity at Clayoquot Sound, British Columbia. Catena, 38(4), 279–300.Google Scholar
Jakobsen, I., Chen, B. D., Munkvold, L., Lundsgaard, T., and Zhu, Y. G. (2005). Contrasting phosphate acquisition of mycorrhizal fungi with that of root hairs using root hairless barley mutant. Plant Cell Environ., 28, 928–938.Google Scholar
Janos, D. P. (1988). Mycorrhiza applications in tropical forestry: Are temperate-zone approaches appropriate? In Ng, S. P. (ed.), Trees and Mycorrhiza. Forest Research Institute, Kuala Lumpur, Malaysia, pp. 133–188
Janzen, D. H. (1988). Tropical Dry Forests: The Most Endangered Major Tropical Ecosystem. in Biodiversity, EO Wilson (ed.), Washington, D.C., National Academy Press, pp 130-137.
Jenkins, M. (2003). Prospects for biodiversity. Science, 302(5648), 1175–1177.Google Scholar
Jetz, W., Wilcove, D. S., and Dobson, A. P. (2007). Projected impacts of climate and land-use change on the global diversity of birds. PLoS Biology, 5(6), e157.Google Scholar
Jipp, P. H., Nepstad, D. C., Cassel, D. K., and Reis De Carvalho, C. (1998). Deep soil moisture storage and transpiration in forests and pastures of seasonally-dry Amazonia. Climatic Change, 39, 2–3, 395–412.Google Scholar
Joffre, R., and Rambal, S. (1988). Soil-water improvement by trees in the rangelands of Southern Spain. Acta Oecologica – Oecologia Plantarum, 9, 405–422.Google Scholar
Johnson, D. W., and Henderson, P. (1995). Effects of forest management and elevated carbon dioxide on soil carbon storage. Soil Management and the Greenhouse Effects, 137–145.Google Scholar
Jones, J. A. (2000). Hydrologic processes and peak discharge response to forest removal, regrowth, and roads in 10 small experimental basins, western Cascades, Oregon. Water Resour. Res., 36(9), 2621–2642.Google Scholar
Jones, D. L., and Oburger, E. (2011). Solubilization of phosphorus by soil microorganisms. In Buenemann, E. K., Oberson, A., and Frossard, E. (eds.), Phosphorus in Action: Biological Processes in Phosphorus Cycling, Soil Biology. 26.Springer, Heidelberg, 169–198.
Jones, D. T., Sah, J. P., Ross, M. S., Oberbauer, S. F., Hwang, B., and Jayachandran, K. (2006). Growth and physiological responses of twelve tree species common in Everglades tree islands to simulated hydrologic regimes. Wetlands, 26, 830–844.Google Scholar
Jorgenson, M. T., Racine, C. H., Walters, J. C., and Osterkamp, T. E. (2001). Permafrost degradation and ecological changes associated with a warming climate in Central Alaska. Climatic Change, 48, 551–579.Google Scholar
Ju, X. T. et al. (2009). Reducing environmental risk by improving N management in intensive Chinese agricultural systems. Proceedings of the National Academy of Sciences, 106(9), 3041–3046.Google Scholar
Juang, J.-Y. et al. (2007). Hydrologic and atmopsheric controls on initiation of convective precipitation events. Water Resour. Res., 43, W03421, doi:10.1029/2006WR004954.CrossRefGoogle Scholar
Jull, C. (ed.) (2007). Recent trends in the law and policy of bioenergy production, promotion and use (No. 95). Food & Agriculture Organization, Rome.
Kaimowitz, D., Mertens, B., Wunder, S., and Pacheco, P. (2004). Hamburger connection fuels Amazon destruction. Centre for International Forestry Research, Bogor, Indonesia.
Kaiser, B., and Roumasset, J. (2002). Valuing indirect ecosystem services: the case of tropical watersheds.Environment and Development Economics, 7(04), 701–714.Google Scholar
Kallio, A., and Rieger, S. (1969). Recession of permafrost in a cultivated soil of interior Alaska. Soil Sci. Soc. Am. Proc. 33, 430–432.Google Scholar
Kaplan, J. O., Krumhardt, K. M., & Zimmermann, N. (2009). The prehistoric and preindustrial deforestation of Europe. Quaternary Science Reviews, 28(27), 3016–3034.Google Scholar
Karlen, D. L., Varvel, G. E., Bullock, D. G., and Cruse, R. M. (1994). Crop rotations for the 21st century. Advances in Agronomy, 53, 1–45.Google Scholar
Karp, A., and Shield, I. (2008). Bioenergy from plants and the sustainable yield challenge. New Phytologist, 179(1), 15–32.Google Scholar
Kasischke, E. S., Melack, J. M., Dobson, M. C. (1997). The use of imaging radars for ecological applications – a review. Remote Sensing of Environment, 59, 141–156.Google Scholar
Kastner, T., Rivas, M. J. I., Koch, W., and Nonhebel, S. (2012). Global changes in diets and the consequences for land requirements for food. Proc Natl Acad Sci USA, 109(18): 6868–6872.Google Scholar
Katul, G. G., and Novick, K. A. (2009). Evapotranspiration. In Likens, G. E. (ed.), Encyclopedia of Inland Waters. Elsevier, Oxford, UK, pp. 661–667. doi:10.1016/B978-012370626-3.00012-0.
Katul, G. G., Oren, R., Manzoni, S., Higgins, C., and Parlange, M. B. (2012). Evapotranspiration: A process driving mass transport and energy exchange in the soil-plant-atmosphere-climate system. Rev. Geophys., 50(3), RG3002, doi:10.1029/2011RG000366.Google Scholar
Kauffman, J. B., Cummings, D. L., and Ward, D. E. (1994). Relationships of fire, biomass and nutrient dynamics along a vegetation gradient in the Brazilian cerrado. Journal of Ecology, 82(3), 519–531.Google Scholar
Kauffman, J. B., SanfordJr., R. L., Cummings, D. L., Salcedo, I. H., and Sampaio, E. V. (1993). Biomass and nutrient dynamics associated with slash fires in neotropical dry forests. Ecology, 74(1), 140–151.Google Scholar
Kazianga, H., and Masters, W. A. (2006). Property rights, production technology, and deforestation: Cocoa in Cameroon. Agricultural Economics, 35(1), 19–26.Google Scholar
Kellman, M. (1979). Soil enrichment by Neotropical savanna trees. J. Ecol., 67, 565-577Google Scholar
Kendall, C., and McDonnell, J. J (eds.). (1998). Isotope Tracers in Catchment Hydrology. Elsevier Science B.V., Amsterdam.
Kennedy, G., Nantel, G., and Shetty, P. (2004). Globalization of food systems in developing countries: Impacts on food security and nutrition. FAO Food and Nutrition Paper, 83, 1–26.
Khanna, P. K., Raison, R. J., and Falkiner, R. A. (1994). Chemical properties of ash derived from Eucalyptus litter and its effects on forest soils. Forest Ecology Management, 66, 107–125.Google Scholar
Kinzig, A. P., Perrings, C., Chapin, F. S., Polasky, S., Smith, V. K., Tilman, D., and Turner, B. L. (2011). Paying for ecosystem services promise and peril. Science, 334(6056), 603–604.
Kirschbaum, M. U. (1995). The temperature dependence of soil organic matter decomposition, and the effect of global warming on soil organic C storage. Soil Biology and Biochemistry, 27(6), 753–760.Google Scholar
Kleidon, A., Fraedrich, K., and Heimann, M. (2000). A green planet versus a desert world: Estimating the maximum effect of vegetation on the land surface climate. Climat. Change, 44, 471–493.Google Scholar
Kleidon, A., and Heimann, M. (1999). Deep-rooted vegetation, Amazonian deforestation, and climate: Results from a modeling study. Global Ecol. Biogeogr., 8, 397–405.Google Scholar
Kleinen, T., Held, H., and Petschel-Held, G. (2003). The potential role of spectral properties in detecting thresholds in the Earth system: Application to the thermohaline circulation. Ocean Dynam., 53, 53–63.Google Scholar
Kleinman, P., Pimentel, D., and Bryant, R. (1996), The ecological sustainability of slash-and-burn agriculture. Agric. Ecosyst. Environ., 52, 235–249.Google Scholar
Klepeis, P. (2003). Development policies and tropical deforestation in the southern Yucatan peninsula: Centralized and decentralized approaches. Land Degradation and Development, 14(6), 541–561.Google Scholar
Knight, R. (2015). Balancing the numbers: Using grassroots land valuation to empower communities in land investment negotiations, 2015 Land and Poverty Conference. World Bank, Washington, DC.
Knowles, R. (1982). Denitrification. Microbiological Reviews, 46(1), 43.Google Scholar
Koh, L. P. (2008). Birds defend oil palms from herbivorous insects. Ecol. Appl., 18, 821–825.Google Scholar
Koh, L. P., Miettinen, J., Liew, S. C., and Ghazoul, J. (2011). Remotely sensed evidence of tropical peatland conversion to oil palm. Proceedings of the National Academy of Sciences, 108(12), 5127–5132.Google Scholar
Koh, L. P., and Wilcove, D. S. (2008). Is oil palm agriculture really destroying tropical biodiversity?Conservation Letters, 1(2), 60–64.Google Scholar
Koop, G., and Tole, L. (1999). Is there an environmental Kuznets curve for deforestation?Journal of Development Economics, 58(1), 231–244.Google Scholar
Körner, C. (1998). A re-assessment of high elevation treeline positions and their explanation. Oecologia, 115, 445–459.Google Scholar
Körner, C. (2006). Plant CO2 responses: An issue of definition, time and resource supply. New Phytologist, 172(3), 393–411.Google Scholar
Körner, C., and Paulsen, J. (2004). A world-wide study of high altitude treeline temperatures. Journal of Biogeography, 31, 713–732.Google Scholar
Kozlowski, T. T. (1965). Responses of woody plants to flooding. In Kozlowski, T. T. (ed.), Flooding and Plant Growth. Elsevier, New York, pp. 129–163.
Krammers, J. S., and DeBano, L. F. (1965). Soil wettability: A neglected factor in watershed management. Water Resources Research, 1, 283–286.Google Scholar
Krause, G. H. (1994). Photoinhibition Induced by Low Temperatures. Photoinhibition of Photosynthesis: From Molecular Mechanisms to the Field, edited by Baker, N. R. and Bowyer, J. R.. BIOS Scientific, Oxford, pp. 331–348.
Krauss, K. W., Lovelock, Catherine E., McKee, Karen L., López-Hoffman, L., Ewe, S. M. L., and Sousa, W. P. (2008). Environmental drivers in mangrove establishment and early development: A review. Aquatic Botany, 89, 105–127.Google Scholar
Kremen, C. et al. (2000). Economic incentives for rain forest conservation across scales. Science, 288(5472), 1828–1832.Google Scholar
Kress, M. R., Graves, M. R., and Bourne, S. G. (1996). Loss of bottomland hardwood forests and forested wetlands in the Cache River basin, Arkansas. Wetlands 16, 258–263.Google Scholar
Krupinsky, J. M., Bailey, K. L., McMullen, M. P., Gossen, B. D., and Turkington, T. K. (2002). Managing plant disease risk in diversified cropping systems. Agronomy Journal, 94(2), 198–209.Google Scholar
Kullman, L. (1988). Holocene history of the forest – Alpine tundra ecotone in the Scandes Mountains (central Sweden). New Phytologist, 108, 101–110.Google Scholar
Kumari, K. (1994). Sustainable forest management in Peninsular Malaysia: Towards a total economic valuation, Doctoral dissertation, University of East Anglia.
Kummu, M., de Moel, H., Porkka, M., Siebert, S., Varis, O., and Ward, P. J. (2012). Lost food, waste resources: Global food supply chain losses and their impacts on freshwater, cropland, and fertilizer use. Sci. Total Environ., 438, 477–489.Google Scholar
Kuriakose, S. L., van Beek, L. P. H, and van Westen, C. J. (2009). Parameterizing a physically based shallow landslide model in a data poor region. Earth Surface Processes and Landforms, 34(6), 867–881.Google Scholar
Kuznetsov, Y. A. (1995). Elements of applied bifurcation theory. Springer-Verlag, New York.
Lade, S. J., Tavoni, A., Levin, S. A., and Schlüter, M. (2013). Regime shifts in a social-ecological system. Theoretical Ecology, 6(3), 359–372.Google Scholar
Lamarque, J. F. et al. (2005). Assessing future nitrogen deposition and carbon cycle feedback using a multimodel approach: Analysis of nitrogen deposition. Journal of Geophysical Research: Atmospheres (1984–2012), 110(D19).Google Scholar
Lambers, H., ChapinIII, F. S., and Pons, T. L. (2008b). Interactions among Plants. In Plant Physiological Ecology. Springer, New York, pp. 505–531.
Lambers, H., Raven, J. A., Shaver, G. R., and Smith, S. E. (2008a). Plant nutrient-acquisition strategies change with soil age. Trends in Ecology & Evolution, 23(2), 95–103.Google Scholar
Lambin, E. F. (1994). Modeling deforestation processes: A review. TREES (Tropical Ecosystem Environment Observation by Satellites). Research Report 1. European Commission Joint Research Centre/European Space Agency, Brussels.
Lambin, E. F., and Geist, H. J. (2003). Regional differences in tropical deforestation. Environment: Science and Policy for Sustainable Development, 45(6), 22–36.Google Scholar
Lambin, E. F., Geist, H. J., and Lepers, E. (2003). Dynamics of land-use and land-cover change in tropical regions. Annu. Rev. Environ. Resour., 28, 205–241.Google Scholar
Lambin, E. F., Rounsevell, M. D. A., and Geist, H. J. (2000). Are agricultural land-use models able to predict changes in land-use intensity?Agriculture, Ecosystems and Environment, 82(1), 321–331.Google Scholar
Lane, J. (2013). IEA says cellulstic biofuels capacity has tripled since 2010, International Energy Agency, New Task 39 Global report, Biofuels Digest, available at: www.biofuelsdigest.com.
Langford, A. O., Fehsenfeld, F. C., Zachariassen, J., and Schimel, D. S. (1992). Gaseous ammonia fluxes and background concentrations in terrestrial ecosystems of the United States. Global Biogeochemical Cycles, 6(4), 459–483.Google Scholar
Langvall, O., and Örlander, G. (2001). Effects of pine shelterwoods on microclimate and frost damage to Norway spruce seedlings. Canadian Journal of Forest Research, 31, 155–164.Google Scholar
Langvall, O., and Ottosson Löfvenius, M. (2002). Effect of shelterwood density on nocturnal near-ground temperature, frost injury risk and budburst date of Norway spruce. Forest Ecology and Management, 168, 149–161.Google Scholar
Lapola, D. M., Schaldach, R., Alcamo, J., Bondeau, A., Koch, J., Koelking, C., and Priess, J. A. (2010). Indirect land-use changes can overcome carbon savings from biofuels in Brazil. Proceedings of The national Academy of Sciences, 107(8), 3388–3393.Google Scholar
Larcher, W. (1995). Physiological Plant Ecology, 3rd ed. Springer, Berlin.
Larsen, J. A. (1980). The Boreal Ecosystem. Academic Press, New York.
Laurance, W. F., Albernaz, A. K., Schroth, G., Fearnside, P. M., Bergen, S., Venticinque, E. M., & Da Costa, C. (2002). Predictors of deforestation in the Brazilian Amazon.Journal of Biogeography, 29(5–6), 737–748.Google Scholar
Laurance, W. F., Cochrane, M. A., Bergen, S., Fearnside, P. M., Delamônica, P., Barber, C., ... & Fernandes, T. (2001). The future of the Brazilian Amazon. Science (Washington), 291(5503), 438–439.Google Scholar
Laurance, W. F. et al. (2012). Averting biodiversity collapse in tropical forest protected areas. Nature, 489(7415), 290–294.Google Scholar
Lawrence, D., D'Odorico, P., Diekmann, L., DeLonge, M., Das, R., and Eaton, J. (2007). Ecological feedbacks following deforestation create the potential for a catastrophic ecosystem shift in tropical dry forest. Proc. Natnl Acad. Sci, USA, PNAS, 104(52), 20696–20701.Google Scholar
Lawrence, D., and Schlesinger, W. H. (2001). Changes in soil phosphorus during 200 years of shifting cultivation in Indonesia. Ecology, 82, 2769–2780.Google Scholar
Lawrence, D., and Vandecar, K. (2015). Effects of tropical deforestation on climate and agriculture. Nature Climate Change, 5(1), 27–36.Google Scholar
Lawton, J. H., and Brown, V. K. (1993). Redundancy in ecosystems. In Schulze, E. D. and H. A. Mooney (eds.), Biodiversity and Ecosystem Function. pp. 255–270, Springer, Berlin.
Lawton, R. O., Nair, U. S., PielkeSr., R. A., and Welch, R. M. (2001). Climatic impact of tropical lowland deforestation on nearby montane cloud forests. Science, 294, 584–587.Google Scholar
Le Maitre, D. C., Scott, D. F., and Colvin, C. (1999). A review of information on interactions between vegetation and groundwater. – Water SA, 25(2), 137–152.Google Scholar
Le Toan, T. et al. (2011). The BIOMASS mission: Mapping global forest biomass to better understand the terrestrial carbon cycle. Remote Sensing of Environment, 115(11), 2850–2860.Google Scholar
Lean, J., and Rowntree, P. R. (1999). Correction note on “Understanding the sensitivity of a GCM simulation of Amazonian deforestation to the specification of vegetation and soil characteristics.” J. Clim., 12, 1549–1551.Google Scholar
Lean, J., and Warrilow, D. A. (1989). Simulation of the regional climatic impact of Amazon deforestation. Nature, 342, 411–413.Google Scholar
LeBauer, D. S., and Treseder, K. K. (2008). Nitrogen limitation of net primary productivity in terrestrial ecosystems is globally distributed. Ecology, 89(2), 371–379.Google Scholar
Lee, J. E., Lintner, B. R., Boyce, C. K., and Lawrence, P. J. (2011). Land use change exacerbates tropical South American drought by sea surface temperature variability. Geophys. Res. Lett., 38, L19706, doi:10.1029/2011GL049066.Google Scholar
Lee, J. E. et al. (2012). Reduction of tropical land region precipitation variability via transpiration. Geophys. Res. Lett., 39, L19704, doi:10.1029/2012GL053417.Google Scholar
Lee, R. (1978). Forest Microclimatology. Columbia University Press, New York.
Lee, R. (1980). Forest Hydrology. Columbia University Press, New York.
Lee, R. (2011). The outlook for population growth. Science, 333, 569–573.Google Scholar
Leighton, M., and Wirawan, N. (1986). Catastrophic drought and fire in Borneo tropical rain forest associated with the 1982–83 El Nino southern oscillation event, 75–102. In Prance, G. T. (ed.) Tropical rain forests and the world atmosphere. AAAS Selected Symposium 101. West-view, Boulder, CO.
Lemus, R., and Lal, R. (2005). Bioenergy crops and carbon sequestration. Critical Reviews in Plant Sciences, 24(1), 1–21.Google Scholar
Leopoldo, P. R., Chaves, J. G., and Franken, W. K. (1993). Solar energy budgets in central Amazonian ecosystems: A comparison between natural forest and bare soil areas. Forest Ecology and Management, 59(3–4), 313–328.Google Scholar
Letey, J. (2001). Causes and consequences of fire-induced soil water repellency. Hydrol. Process., 15, 2867–2875.Google Scholar
Lewis, S. L. et al. (2013). Above-ground biomass and structure of 260 African tropical forests. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1625), 20120295.Google Scholar
Li, Y., Zhao, M., Motesharrei, S., Mu, Q., Kalnay, E., and Li, S. (2015). Local cooling and warming effects of forests based on satellite observations. Nature Communications, 2015; 6: 6603 DOI: 10.1038/ncomms7603.Google Scholar
Liebman, M., and Altieri, M. A. (1986). Insect, weed and plant disease management in multiple cropping systems. MacMillan, New York.
Lieffers, V. J., and Rothwell, R. L. (1986). Effects of depth of water table and substrate temperature on root and top growth of Picea mariana and Larix lancina seedlings. Canadian Journal of Forest Research, 16(6), 1201–1206.Google Scholar
Lieth, H., and Werger, M. J. A. (1989). Ecosystems of the world. 14B. Tropical Rain Forest Ecosystems: Biogeographical and Ecological Studies. Elsevier, Amsterdam.
Likens, G. E., Bormann, F. H., and Johnson, N. M. (1969). Nitrification: Importance to nutrient losses from a cutover forested ecosystem. Science, 163(3872), 1205–1206.Google Scholar
Likens, G. E., Bormann, F. H., Johnson, N. M., Fisher, D. W., and Pierce, R. S. (1970). Effects of forest cutting and herbicide treatment on nutrient budgets in the Hubbard Brook watershed-ecosystem. Ecological Monographs, 40(1), 23–47.Google Scholar
Lima, M., Skutsch, M., and de Medeiros Costa, G. (2011). Deforestation and the social impacts of soy for biodiesel: Perspectives of farmers in the south Brazilian Amazon. Ecology and Society, 16(4), 04, available at: http://dx.doi.org/10.5751/ES-04366-160404.Google Scholar
Linde, M., Galbe, M., and Zacchi, G. (2008). Bioethanol production from non-starch carbohydrate residues in process streams from a dry-mill ethanol plant. Bioresource Technology, 99(14), 6505–6511.Google Scholar
Lindenmayer, D. B., and Franklin, J. F. (2002). Conserving Forest Biodiversity: A Comprehensive Multiscaled Approach. Island Press, Washington, DC.
Lindenmayer, D. B., Franklin, J. F., and Fischer, J. (2006). General management principles and a checklist of strategies to guide forest biodiversity conservation. Biological conservation, 131(3), 433–445.Google Scholar
Linn, D. M., and Doran, J. W. (1984). Effect of water-filled pore space on carbon dioxide and nitrous oxide production in tilled and nontilled soils. Soil Science Society of America Journal, 48(6), 1267–1272.Google Scholar
Liscow, Z. D. (2013). Do property rights promote investment but cause deforestation? Quasi-experimental evidence from Nicaragua. Journal of Environmental Economics and Management, 65(2), 241–261.Google Scholar
Liski, J., Ilvesniemi, H., Mäkelä, A., and Starr, M. (1998). Model analysis of the effects of soil age, fires and harvesting on the carbon storage of boreal forest soils. European Journal of Soil Science, 49(3), 407–416.Google Scholar
Liu, W., Meng, F., Zhang, Y., Liu, Y., and Li, H. (2004). Water input from fog drip in the tropical seasonal rain forest of Xishuangbanna, South-West China. Journal of Tropical Ecology, 20, 517–524.Google Scholar
Lloyd, A. H., Yoshikawa, K., Fastie, C. L., Hinzman, L., and Fraver, M. (2003). Effects of permafrost degradation on woody vegetation at Arctic treeline on the Seaward Peninsula. Permafrost and Periglacial Processes, 14, 93–101.Google Scholar
Loarie, S. R., Lobell, D. B., Asner, G. P., and Field, C. B. (2011). Land-cover and surface water change drive large albedo increases in South America. Earth Interact, 15, 1–16, doi:10.1175/2010EI342.1.Google Scholar
Loik, M. E., and Nobel, P. S. (1993). Freezing tolerance and water relations of Opuntia fragilis from Canada and the United States. Ecology, 74, 1722–1732.Google Scholar
Lopez, R. (1994). The environment as a factor of production: The effects of economic growth and trade liberalization. Journal of Environmental Economics and Management, 27(2), 163–184.Google Scholar
Lovett, G. M. (1994). Atmospheric deposition of nutrients and pollutants in North America: An ecological perspective. Ecological Applications, 4(4), 629–650.Google Scholar
Lovett, G. M., and Kinsman, J. D. (1990). Atmospheric pollutant deposition to high-elevation. Atmos. Environ. 24(11), 2767–2786.Google Scholar
Lovett, G. M., and Reiners, W. A. (1986). Canopy structure and cloud water deposition in subalpine coniferous forests. Tellus TELLAL, 38B (5), 319–327.Google Scholar
Lu, D., Batistella, M., and Moran, E. (2005). Satellite estimation of aboveground biomass and impacts of forest stand structure. Photogrammetric Engineering & Remote Sensing, 71(8), 967–974.Google Scholar
Luck, G. W., Daily, G. C., and Ehrlich, P. R. (2003). Population diversity and ecosystem services. Trends in Ecology and Evolution, 18(7), 331–336.Google Scholar
Lundmark, T., and Hällgren, J.-E. (1987). Effects of frost on shaded and exposed spruce and pine seedlings planted in the field. Canadian Journal of Forest Research, 17, 1197–1201.Google Scholar
Luo, Y. et al. (2004). Progressive nitrogen limitation of ecosystem responses to rising atmospheric carbon dioxide. Bioscience, 54(8), 731–739.Google Scholar
Luyssaert, S. et al. (2007). CO2 balance of boreal, temperate, and tropical forests derived from a global database. Global Change Biology, 13(12), 2509–2537.Google Scholar
MacArthur, R. H., and Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton University Press, Princeton, NJ.
MacDonald, G. E. (2004). Cogongrass (Imperata cylindrica) – biology, ecology, and management. Critic. Rev. Plant Sci. 23, 367–380.Google Scholar
Machimura, T., Kobayashi, Y., Iwahana, G., Hirano, T., Lopez, L., Fukuda, M., and Fedorov, A. N. (2005). Change of carbon dioxide budget during three years after deforestation in Eastern Siberian Larch Forest. J. Agric. Meteorol., 60(5), 653–656.Google Scholar
Madi, M. A. C. (2004). Financial liberalization and macroeconomic policy options: Brazil, 1994–2003. Campinas: Instituto de Economia da UNICAMP, p. 32, available at: www.eco.unicamp.br/Downloads/Publicacoes/TextosDiscussao/texto117.pdf.
Maertens, M., Zeller, M., and Birner, R. (2006). Sustainable agricultural intensification in forest frontier areas. Agricultural Economics, 34(2), 197–206.Google Scholar
Magnusson, T. (1994). Studies of the soil atmosphere and related physical characteristics in peat forest soils. For. Ecol. Manage., 67, 203–224.Google Scholar
Mahar, D. J. (1979). Frontier Development Policy in Brazil: A Study of Amazonia. Praeger, New York.
Maher, E. L., and Germino, Matthew J. (2006). Microsite differentiation among conifer species during seedling establishment at alpine treeline. Ecoscience, 13, 334–341.Google Scholar
Maher, E. L., Germino, Matthew J., and Hasselquist, N. J. (2005). Interactive effects of tree and herb cover on survivorship, physiology, and microclimate of conifer seedlings at the alpine tree-line ecotone. Canadian Journal of Forest Research, 35, 567–574.Google Scholar
Maherali, H., Pockman, W. T., and Jackson, R. B. (2004). Adaptive variation in the vulnerability of woody plants to xylem cavitation. Ecology, 85, 2184–2199.Google Scholar
Mahowald, N. et al. (2008). The global distribution of atmospheric phosphorus deposition and anthropogenic impacts. Global Biogeochem. Cycles, 22, GB4026, doi:10.1029/2008GB003240.Google Scholar
Malézieux, E. et al. (2009). Mixing plant species in cropping systems: Concepts, tools and models: A review. Agronomy for Sustainable Development, 29(1), 43–62.Google Scholar
Malhi, Y., Adu-Bredu, S., Asare, R. A., Lewis, S. L., and Mayaux, P. (2013). African rainforests: Past, present and future. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1625), 20120312.Google Scholar
Malhi, Y., Baldocchi, D. D., and Jarvis, P. G. (1999). The carbon balance of tropical, temperate and boreal forests. Plant, Cell & Environment, 22(6), 715–740.Google Scholar
Malhi, Y., and Grace, J. (2000). Tropical forests and atmospheric carbon dioxide. Trends in Ecology & Evolution, 15(8), 332–337.Google Scholar
Malhi, Y., Roberts, J. T., Betts, R. A., Killeen, T. J., Li, W., and Nobre, C. A. (2008). Climate change, deforestation and the fate of the Amazon. Science, 319, 169–172.Google Scholar
Malhi, Y. et al. (2009). Exploring the likelihood and mechanism of a climate-change-induced dieback of the Amazon rainforest. PNAS, 106(49), 20610–20615.Google Scholar
Malingreau, J. P., Stephens, G., and Fellows, L. (1985). Remote sensing of forest fires: Kalimantan and North Borneo in 1982–83, Ambio, 14(6), 314–321.Google Scholar
Man, R., and Lieffers, V. J. (1999). Effects of shelterwood and site preparation on microclimate and establishment of white spruce seedlings in a boreal mixedwood forest. The Forestry Chronicle, 75, 837–844.Google Scholar
Manzoni, S., Trofymow, J. A., Jackson, R. B., and Porporato, A. (2010). Stoichiometric controls on carbon, nitrogen, and phosphorus dynamics in decomposing litter. Ecological Monographs, 80(1), 89–106.Google Scholar
Markewitz, D., Davidson, E. A., Figueiredo, R. D. O., Victoria, R. L., and Krusche, A. V. (2001). Control of cation concentrations in stream waters by surface soil processes in an Amazonian watershed. Nature, 410(6830), 802–805.Google Scholar
Marshall, E., Schreckenberg, K., and Newton, A. C. (eds.) (2006). Commercialization of Non-Timber Forest Products: Factors Influencing Success: Lessons Learned from Mexico and Bolivia and Policy Implications for Decision-Makers. UNEP World Conservation Monitoring Centre, Cambridge.
Martinez-Meze, E., and Whitford, W. G. (1996). Stemflow, throughfall and channelization of stemflow by roots in the three Chihuahuan desert shrubs. J. Arid Environ., 32, 271–287.Google Scholar
Martínez-Yrízar, A., Bullock, S. H., Mooney, H. A., and Medina, E. (1995). Seasonally Dry Tropical Forests. Cambridge University Press, Cambridge.
Mas, J. F., Sorani, V., and Alvarez, R. (1996). Elaboración de un modelo de simulación del proceso de deforestación., Investigaciones Geográficas, 5, 43–57.Google Scholar
Masera, Díaz-Jiménez, O. R., , R., and Berrueta, V. (2005). From cookstoves to cooking systems: The integrated program on sustainable household energy use in Mexico. In Energy for Sustainable Development, vol. IX(1). IEI, India.
Mather, A. S. (1992). The forest transition. Area, 367–379.Google Scholar
Mather, A. S., Needle, C. L., and Fairbairn, J. (1999). Environmental Kuznets curves and forest trends. Geography, 55–65.Google Scholar
Matson, P. (1990). Plant-soil interactions in primary succession at Hawaii Volcanoes National Park. Oecologia, 85, 241–246.Google Scholar
Matson, P. A., Billow, C., Hall, S., and Zachariassen, J. (1996). Fertilization practices and soil variations control nitrogen oxide emissions from tropical sugar cane. Journal of Geophysical Research: Atmospheres (1984–2012), 101(D13), 18533–18545.Google Scholar
Matson, P. A., McDowell, W. H., Townsend, A. R., and Vitousek, P. M. (1999). The globalization of N deposition: Ecosystem consequences in tropical environments. Biogeochemistry, 46(1–3), 67–83.Google Scholar
Matson, P. A., Naylor, R., and Ortiz-Monasterio, I. (1998). Integration of environmental, agronomic, and economic aspects of fertilizer management. Science, 280 (5360), 112–115.Google Scholar
Matson, P. A., and Vitousek, P. M. (1981). Nitrogen mineralization and nitrification potentials following clearcutting in the Hoosier National Forest, Indiana. Forest Science, 27(4), 781–791.Google Scholar
Matson, P. A., and Vitousek, P. M. (2006). Agricultural intensification: Will land spared from farming be land spared for nature?Conservation Biology, 20(3), 709–710.Google Scholar
Matthews, E. (1983). Global vegetation and land use: New high-resolution data bases for climate studies. Journal of Climate and Applied Meteorology, 22(3), 474–487.Google Scholar
May, R.M. (1973). Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton, New Jersey.
May, R. M. (1977). Thresholds and breakpoints in ecosystems with a multiplicity of stable states. Nature, 269, 471–477.Google Scholar
Mayaux, P. et al. (2013). State and evolution of the African rainforests between 1990 and 2010. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1625), 20120300.Google Scholar
McBirney, A. R. (1993). Igneous Petrology, 2nd ed. Jones & Bartlett, Boston.
McCann, K. S. (2000). The diversity–stability debate. Nature, 405(6783), 228–233.Google Scholar
McDowell, N. et al. (2008). Mechanisms of plant survival and mortality during drought: Why do some plants survive while others succumb to drought?New Phytologist, 178(4), 719–739.Google Scholar
McGill, W. B., and Cole, C. V. (1981). Comparative aspects of cycling of organic C, N, S and P through soil organic matter. Geoderma, 26, 267–286.Google Scholar
McGrady-Steed, J., and Morin, P. J. (2000). Biodiversity, density compensation, and the dynamics of populations and functional groups. Ecology, 81(2), 361–373.Google Scholar
McGroddy, M. E., Silver, W. L., and de Oliveira, R. C. (2004). The effect of phosphorus availability on decomposition dynamics in a seasonal lowland Amazonian forest. Ecosystems, 7, 172–179.Google Scholar
McGuffie, K., Henderson-Sellers, A., Zhang, H., Durbridge, T. B., and Pitman, A. J. (1995). Global climate sensitivity to tropical deforestation. Global Planet. Change, 10, 97–128.Google Scholar
McGuire, A. D. et al. (2001). Carbon balance of the terrestrial biosphere in the twentieth century: Analyses of CO2, climate and land use effects with four process-based ecosystem models. Global Biogeochemical Cycles, 15(1), 183–206.Google Scholar
McJannet, D., Wallace, J., and Reddell, P. (2007). Precipitation interception in Australian tropical rainforests. II. Altitudinal gradients of cloud interception, stemflow, throughfall and interception. Hydrological Processes, 21(13), 1703–1718.Google Scholar
McKee, K. L., and Rooth, J. E. (2008). Where temperate meets tropical: Multi-factorial effects of elevated CO2, nitrogen enrichment, and competition on a mangrove-salt marsh community. Global Change Biology, 14, 971–984.Google Scholar
McNaughton, S. J. (1977). Diversity and stability of ecological communities: A comment on the role of empiricism in ecology. American Naturalist, 515–525.Google Scholar
McRae, L., Freeman, R., and Deinet, S. (2014). The Living Planet Index. In McLellan, R., L.Iyengar, B.Jeffries, , and Oerlemans, N. (eds.), Living Planet Report 2014: Species and Spaces, People and Places. WWF, Gland, Switzerland.
Medeiros, J. S., and Pockman, W. T. (2011). Drought increases freezing tolerance of both leaves and xylem of Larrea tridentata. Plant, Cell and Environment, 34, 43–51.Google Scholar
Medvigy, D., Walko, R. L., and Avissar, R. (2011). Effects of deforestation on spatiotemporal distributions of precipitation in South America. J. Clim., 24, 2147–2163, doi:10.1175/2010JCLI3882.1.Google Scholar
Megevand, C. (2013). Deforestation trends in the Congo Basin: Reconciling economic growth and forest protection. World Bank Publications.
Meggers, B. J. (1994). Archeological evidence for the impact of Mega-Nino events on Amazonia during the past two millennia. Clim. Change, 28, 321–338.Google Scholar
Melillo, J. M. et al. (2009). Indirect emissions from biofuels: How important?Science, 326(5958), 1397–1399.Google Scholar
Melo, F. P. L, Pinto, S. R. R., Brancalion, P. H. S., Castro, P. S., Rodrigues, R. R., Aronson, J., and Tabarelli, M. (2013). Priority setting for scaling up tropical forest restoration projects: Early lessons from Atlantic Forest Restoration Pact. Environmental Science and Policy, 33, 395–404.Google Scholar
Mendelsohn, R. (1994). Property rights and tropical deforestation. Oxford Economic Papers, 750–756.Google Scholar
Mendez, M., and Popkin, B. (2004). Globalization, urbanization and nutritional change in the developing world: Globalization of food systems in developing countries. Impact on Food Security and Nutrition, 55–80.Google Scholar
Merriam, R. A. (1973). Fog drip from artificial leaves in a fog wind tunnel. Water Resources Research, 9, 1591–1598.Google Scholar
Merry, F. D., Hildebrand, P. E., Pattie, P., & Carter, D. R. (2002). An analysis of land conversion from sustainable forestry to pasture: a case study in the Bolivian Lowlands.Land Use Policy, 19(3), 207–215.Google Scholar
Meusburger, K., and Alewell, C. (2008). Impacts of anthropogenic and environmental factors on the occurrence of shallow landslides in an alpine catchment (Urseren Valley, Switzerland). Natural Hazards and Earth System Science, 8(3), 509–520.Google Scholar
Meyer, W. B. (1995). Past and present land use and land cover in the USA. Consequences, 1(1), 25–33.Google Scholar
Meyfroidt, P., and Lambin, E. F. (2011). Global forest transition: Prospects for an end to deforestation. Annual Review of Environment and Resources, 36(1), 343–371.
Meyfroidt, P., Lambin, E. F., Erb, K.-H., and Hertel, T. W. (2013). Globalization of land use: Distant drivers of land change and geographic displacement of land use. Curr. Opin. Environ. Sustain, 5, 438–444.Google Scholar
Meyfroidt, P., Rudel, T. K., and Lambin, E. F. (2010). Forest transitions, trade, and the global displacement of land use. Proc. Natl. Acad. Sci. USA 107, 20917–20922.Google Scholar
Miettinen, J., Shi, C., and Liew, S. C. (2011). Deforestation rates in insular Southeast Asia between 2000 and 2010. Global Change Biology, 17(7), 2261–2270.Google Scholar
Minetti, J. L., and Lamelas, C. M. (1997). Respuesta regional de la soja en Tucuman a la variabilidad clim´atica. Revista Industrial y Agr´ıcola de Tucuman, 72, 63–68.Google Scholar
Minetti, J. L., and Vargas, W. M. (1997). Trends and jumps in the annual rainfall in South America, south of 15° S. Atmosfera, 11, 205–221.Google Scholar
Minh, D. H. T., Le Toan, T., Rocca, F., Tebaldini, S., d'Alessandro, M. M., and Villard, L. (2014). Relating P-band synthetic aperture radar tomography to tropical forest biomass. Geoscience and Remote Sensing, IEEE Transactions on, 52(2), 967–979.Google Scholar
Mitchard, E. T. et al. (2011). Measuring biomass changes due to woody encroachment and deforestation/degradation in a forest–savanna boundary region of central Africa using multi-temporal L-band radar backscatter. Remote Sensing of Environment, 115(11), 2861–2873.Google Scholar
Mitchard, E. T. et al. (2013). A novel application of satellite radar data: Measuring carbon sequestration and detecting degradation in a community forestry project in Mozambique. Plant Ecology & Diversity, 6(1), 159–170.Google Scholar
Mitsch, W. J., and Gosselink, J. (2000). Wetlands, 3rd ed. John Wiley, New York.
Mittermeier, R. A., Myers, N., Thomsen, J. B., Da Fonseca, G. A., and Olivieri, S. (1998). Biodiversity hotspots and major tropical wilderness areas: Approaches to setting conservation priorities. Conservation Biology, 12(3), 516–520.Google Scholar
Miyake, S., Renouf, M., Peterson, A., McAlpine, C., and Smith, C. (2012). Land-use and environmental pressures resulting from current and future bioenergy crop expansion: A review. Journal of Rural Studies, 28(4), 650–658.Google Scholar
Molotch, N. P., Blanken, P. D., and Link, T. E. (2011). Snow: Hydrological and ecological feedbacks in forests. In Levia, Delphis F., Carlyle-Moses, Darryl, and Tanaka, Tadashi (eds.), Forest Hydrology and Biogeochemistry. Synthesis of Past Research and Future Directions Series: Ecological Studies, Vol. 216. Springer, New York.
Monela, G. C. (1995). Tropical rainforest deforestation, biodiversity benefits and sustainable landuse: Analysis of economic and ecological aspects related to the Nguru Mountains, Tanzania. Doctor Scientiarum Theses (Norway), Agricultural University of Norway.
Montgomery, D. R. (2008). Dirt: The erosion of civilizations. University of California Press, Berkeley.
Montgomery, D. R., and Dietrich, W. E. (1994). A physically-based model for the topographic control on shallow landsliding. Water Resour. Res., 30, 1153–1171.Google Scholar
Mora, C., Tittensor, D. P., Adl, S., Simpson, A. G., and Worm, B. (2011). How many species are there on Earth and in the ocean?PLoS Biology, 9(8), e1001127.Google Scholar
Morton, D. C. et al. (2006). Cropland expansion changes deforestation dynamics in the southern Brazilian Amazon. Proceedings of the National Academy of Sciences, 103(39), 14637–14641.Google Scholar
Mueller, N. D., Gerber, J. S., Johnston, M., Ray, D. K., Ramankutty, N., and Foley, J. A. (2012). Closing yield gaps through nutrient and water management. Nature, 490, 254–257, doi:10.1038/nature11420.Google Scholar
Muller, S. W. (1945). Permafrost or perennially frozen ground and related engineering problems, 2nd ed., U.S. Geol. Surv. Spec. Rep. Strategic Eng., Study No. 62.
Muller, S. W. (1947). Permafrost or permanently frozen ground and related engineering problems. Edwards, Ann Arbor, MI.
Müller, M. M. L., Guimaraes, M. F., Desjardins, T., and Mitja, D. (2004). The relationship between pasture degradation and soil properties in the Brazilian Amazon: A case study. Agriculture, Ecosystems and Environment, 103(2), 279–288.Google Scholar
Mulvaney, T. J. (1851). On the use of self-registering rain and flood gauges in making observations of the relations of rainfall and of flood discharges in a given catchment. Proceedings of the Institute of Civil Engineers of Ireland, 4, pp. 18–31.
Munns, R., and Tester, M. (2008). Mechanisms of salinity tolerance. Annu. Rev. Plant Biol., 59, 651–81.Google Scholar
Murdiyarso, D., Hergoualc'h, K., and Verchot, L. V. (2010). Opportunities for reducing greenhouse gas emissions in tropical peatlands. Proceedings of the National Academy of Sciences, 107(46), 19655–19660.Google Scholar
Murphy, P. G., and Lugo, A. E. (1986). Ecology of tropical dry forest. Annu. Rev. Ecol. Syst., 17, 67–88, doi:10.1146/annurev.es.17.110186.000435.Google Scholar
Murray, M. B., Smith, R. I., Friend, A., and Jarvis, P. G. (2000). Effect of elevated [CO2] and varying nutrient application rates on physiology and biomass accumulation of Sitka spruce (Picea sitchensis). Tree Physiology, 20(7), 421–434.Google Scholar
Murty, D., Kirschbaum, M. U., Mcmurtrie, R. E., and Mcgilvray, H. (2002). Does conversion of forest to agricultural land change soil carbon and nitrogen? A review of the literature. Global Change Biology, 8(2), 105–123.Google Scholar
Myers, N. (1980). Role of forest farmers in conversion of tropical moist forests. A report prepared for the committee on research priorities in tropical Biology of the National Research Council. National Academy of Sciences, Washington, DC
Myers, N. (1988). Tropical forests: Much more than stocks of wood. Journal of Tropical Ecology, 4, 209–21.Google Scholar
Myers, N., Mittermeier, R. A., Mittermeier, C. G., Da Fonseca, G. A., and Kent, J. (2000). Biodiversity hotspots for conservation priorities. Nature, 403(6772), 853–858.Google Scholar
Myers, N., and Tucker, R. (1987). Deforestation in Central America: Spanish legacy and North American consumers. Environmental Review: 11, 55–71.Google Scholar
Myers, S. S. et al. (2013). Human health impacts of ecosystem alteration. Proceedings of the National Academy of Science, USA, PNAS, 110(47), 18753–18760, doi:10.1073/pnas.1218656110.Google Scholar
Myrold, D. D., Matson, P. A., and Peterson, D. L. (1989). Relationships between soil microbial properties and aboveground stand characteristics of conifer forests in Oregon. Biogeochemistry, 8(3), 265–281.Google Scholar
Naeem, S. (1998). Species redundancy and ecosystem reliability. Conservation Biology, 12(1), 39–45.Google Scholar
Naeem, S., and Li, S. (1997). Biodiversity enhances ecosystem reliability. Nature, 390(6659), 507–509.Google Scholar
Naiman, R. J., Decamps, H., and McClain, M. E. (2010). Riparia: Ecology, Conservation, and Management of Streamside Communities. Academic Press, New York.
Nair, U. S., Lawton, R. O., Welch, R. M., and PielkeSr., R. A. (2003). Impact of land use on Costa Rican tropical montane cloud forests: Sensitivity of cumulus cloud field characteristics to lowland deforestation. J. Geophys. Res., 108(D7), 4206, doi:10.1029/2001JD001135.Google Scholar
Nannipieri, P., Giagnoni, L., Landi, L., and Renella, G. (2011). Role of phosphatase enzymes in soil. In Buenemann, E. K., Oberson, A., and Frossard, E. (eds.), Phosphorus in Action: Biological Processes in Phosphorus Cycling. Soil Biology, 26. Springer, Heidelberg, 215–241.
Nassar, A. M., Alvesde, B. F. T., R., and Aguiar, D., M., Bacchi, R. P., and Adami, M. (2008). Prospects of the sugarcane expansion in Brazil: Impacts on direct and indirect land use changes. In Zuurbier, P., and van de Vooren, J. (eds.), Sugarcane Ethanol: Contributions to Climate Change Mitigation and the Environment. Wageningen Academic Publishers, Wageningen.
National Research Council (2008). Hydrologic effects of a changing forest landscape. National Academies Press, Washington, DC.
Naumburg, E., Mata-Gonzales, R., Hunter, R. G., McLendon, T., and Martin, D. W. (2005). Phreatophytic vegetation and groundwater fluctuations: A review of current research and application of ecosystem response modeling with an emphasis on Great Basin vegetation. Environ. Manag., 35, 726–740.Google Scholar
Návar, J., and Bryan, R. (1990). Interception loss and rainfall redistribution by three semi-arid growing shrubs in northeastern Mexico. Journal of Hydrology, 115(1), 51–63.Google Scholar
Naylor, R. L. (1996). Energy and resource constraints on intensive agricultural production. Annual Review of Energy and the Environment, 21(1), 99–123.Google Scholar
Naylor, R. (2011). Expanding the boundaries of agricultural development. Food Sec., 3, 233–251.Google Scholar
Negri, A. J., Adler, R. F., Xu, L., and Surratt, J. (2004). The impact of Amazonian deforestation on dry season rainfall. J. Clim. 17, 1306–1319.Google Scholar
Nelson, P. N. et al. (2014). Oil palm and deforestation in Papua New Guinea. Conservation Letters, 7(3), 188–195.Google Scholar
Nepstad, D. C., Stickler, C. M., and Almeida, O. T. (2006). Globalization of the Amazon soy and beef industries: Opportunities for conservation. Conserv. Biol., 20, 1595–1602.Google Scholar
Nepstad, D. C., Stickler, C. M., Soares-Filho, B., and Merry, F. (2008). Interactions among Amazon land use, forests and climate: Prospects for a near-term forest tipping point. Phil. Trans. R. Soc. B., 363, 1737–1746.Google Scholar
Nepstad, D. C., Tohver, I. M., Ray, D., Moutinho, P., and Cardinot, G. (2007). Mortality of large trees and lianas following experimental drought in an amazon forest. Ecology, 88, 2259–2269.Google Scholar
Nepstad, D. C., Uhl, C., and Serraõ, E. A. S. (1991). Recuperation of a degraded Amazonian landscape: Forest recovery and agricultural restoration. Ambio, 20, 248–255.Google Scholar
Nepstad, D. C. et al. (1994). The role of deep roots in the hydrological and carbon cycles of Amazonian forests and pastures. Nature, 372, 666–669.Google Scholar
Nepstad, D. et al. (2009). The end of deforestation in the Brazilian Amazon. Science, 326(5958), 1350–1351.Google Scholar
Neumann, R. P., and Hirsch, E. (2000). Commercialisation of Non-Timber Forest Products: Review and Analysis of Research. CIFOR, Bogor.
Newmark, W. D. (2001). Tanzanian forest edge microclimatic gradients: Dynamic patterns. Biotropica, 33, 2–11.Google Scholar
Nickel, J. L. (1973). Pest situation in changing agricultural systems – a review. Bulletin of the ESA, 19(3), 136–142.Google Scholar
Niklaus, P. A., Spinnler, D., and Körner, C. (1998). Soil moisture dynamics of calcareous grassland under elevated CO2. Oecologia, 117(1–2), 201–208.Google Scholar
Nobre, C. A., Sellers, P. J., and Shukla, J. (1991). Amazonian deforestation and regional climate change. J. Climate, 4, 957–988.Google Scholar
Norton, T. W. (1996). Conserving biological diversity in Australia's temperate eucalypt forests. For. Ecol. Manage., 85, 21–33.Google Scholar
Noy-Meir, I. (1973). Desert ecosystems: Environment and producers. Ann. Rev. Ecol. Systemat., 4, 25–51.Google Scholar
Noy-Meir, I. (1975). Stability of grazing systems: Application of predator-prey graphs. J. Ecol., 63, 459–481.Google Scholar
Nunez, D., Nahuelhual, L., and Oyarzun, C. (2006). Forests and water: The value of native temperate forests in supplying water for human consumption. Ecol. Econ., 58, 606–616.Google Scholar
Nusslein, K., and Tiedje, J. M. (1999). Soil bacterial community shift correlated with change from forest to pasture vegetation in a tropical soil. Appl. Environ. Microbiol., 65, 3622–3626.Google Scholar
Nye, P. H., and Greenland, D. J. (1964). Changes in the soil after clearing tropical forest. Plant and Soil, 21(1), 101–112.Google Scholar
O'Loughlin, C. L., and Pearce, A. J. (1976). Influence of Cenozoic geology on mass movement and sediment yield response to forest removal, North Westland, New Zealand. Bull. Int. Assoc. Eng. Geol., 14, 41–46.Google Scholar
Oberson, A., and Joner, E. J. (2005). Microbial turnover of phosphorus in the soil. In Turner, B. L., Frossard, E., and Baldwin, D. S., (eds.), Organic Phosphorus in the Environment. CABIL, Wallingford, pp. 133–164.
Oerke, E. C. (2006). Crop losses to pests. Journal of Agricultural Science, 144(01), 31–43.Google Scholar
Oliveira, G. D. T. (2013). Land regularization in Brazil and the global land grab. Development and Change, 44(2), 261–283.Google Scholar
Oliveira-Filho, A. T., and Fontes, M. A. L. (2000). Patterns of foristic differentiation among atlantic forests in southeastern Brazil and the infuence of climate. Biotropica, 32(4b), 793–810.Google Scholar
Olschewski, R., and Benitez, P. C. (2005). Secondary forests as temporary carbon sinks? The economic impact of accounting methods on reforestation projects in the tropics. Ecological Economics, 55(3), 380–394.Google Scholar
Örlander, G. (1993). Shading reduces both vsible and invisible frost damage to Norway spruce seedlings in the field. Forestry, 66, 27–36.Google Scholar
Oslisly, R., White, L., Bentaleb, I., Favier, C., Fontugne, M., Gillet, J. F., and Sebag, D. (2013). Climatic and cultural changes in the west Congo Basin forests over the past 5000 years. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 368(1625), 20120304.Google Scholar
Osterkamp, T. E. et al. (2000). Observations of Thermokarst and its impact on Boreal Forests in Alaska, U.S.A. Arct. Antart. Alp. Res., 32, 303–315.Google Scholar
Ostrom, E. (1990). Governing the Commons: The Evolution of Institutions for Collective Action. Cambridge University Press, Cambridge.
Ouellet, F. (2007). Cold Acclimation and Freezing Tolerance in Plants. Encyclopedia of Life Sciences, 1-6, DOI: 0.1002/9780470015902.a0020093
Ovington, J. D (1983). Ecosystems of the world. 10. Temperate Broad-Leaved Evergreen Forests. Elsevier, Amsterdam.
Oyama, M. D., and Nobre, C. A. (2003). A new climate-vegetation equilibrium state for Tropical South America. Geophysical Research Letters, 30, 2199, doi:10.1029/2003GL018600.Google Scholar
Palmer, M. A., Lee, J., Matthews, J. H., and D'Odorico, P. (2015). Building water security: The role of natural systems. Science, 349(6248), 584–585.Google Scholar
Pan, Y. et al. (2011). A large and persistent carbon sink in the world's forests. Science, 333(6045), 988–993.Google Scholar
Paquette, A., Bouchard, A., and Cogliastro, A. (2006). Survival and growth of under-planted trees: A meta-analysis across four biomes. Ecological Applications, 16, 1575–1589.Google Scholar
Pastor, J., and Post, W. M. (1988). Response of northern forests to CO2-induced climatic change. Nature, 334, 55–58.Google Scholar
Patenaude, G., Hill, R. A., Milne, R., Gaveau, D. L. A., Briggs, B. B. J., and Dawson, T. P. (2004). Quantifying forest above ground carbon content using LiDAR remote sensing. Remote Sensing of Environment, 93(3), 368–380.Google Scholar
Pattanayak, S. K., and Kramer, R. A. (2001). Pricing ecological services: Willingness to pay for drought mitigation from watershed protection in eastern Indonesia. Water Resources Research, 37(3), 771–778.Google Scholar
Pattanayak, S. K., and Sills, E. O. (2001). Do tropical forests provide natural insurance? The microeconomics of non-timber forest product collection in the Brazilian Amazon. Land Economics, 77(4), 595–612.Google Scholar
Paustian, K. A. O. J. H. et al. (1997). Agricultural soils as a sink to mitigate CO2 emissions. Soil Use and Management, 13(s4), 230–244.Google Scholar
Pearce, D., and Brown, K. (eds.) (1994). The Causes of Tropical Deforestation. University of British Columbia Press, Vancouver, pp. 2–26
Peck, A. J., and Hurle, D. H. (1973). Chloride balance of some farmed and forested catchments in Southwestern Australia. Water Resour. Res., 9(3), 648–657.Google Scholar
Peck, A. J., and Williamson, D. R. (1987). Effects of forest clearings on groundwater. J. Hydrol., 94, 47–65.Google Scholar
Peng, S., Garcia, F. V., Laza, R. C., Sanico, A. L., Visperas, R. M., and Cassman, K. G. (1996). Increased N-use efficiency using a chlorophyll meter on high-yielding irrigated rice. Field Crops Research, 47(2), 243–252.CrossRefGoogle Scholar
Pennington, R. T., Lavin, M., and Oilveira-Filho, A. (2009). Woody plant diversity, evolution, and ecology in the tropics: Perspectives from seasonally dry tropical forests. Annu. Rev. Ecol. Evol. Syst., 40, 437–457.Google Scholar
Peñuelas, J. et al. (2013). Human-induced nitrogen–phosphorus imbalances alter natural and managed ecosystems across the globe. Nature Communications, 4.Google Scholar
Pereira, H. M. et al. (2010). Scenarios for global biodiversity in the 21st century. Science, 330(6010), 1496–1501.Google Scholar
Perez, D., and Gonzáles-Lelong, A. (2003). Estimacíon de rendimiento y produccíon de soja y maíz en Tucuman. Campaña 2001–2002. Avance Agroindustrial 23, 19–23.Google Scholar
Perez, D., Gonzales Lelong, A. & Devani, M. (2002) Evolucion de algunos aspectos economico-productivos de la producci´on de soja en la ultima decada en la provincia de Tucuman.Avance Agroindustrial, 22, 31–34.Google Scholar
Perfecto, I., Mas, A., Dietsch, T., and Vandermeer, J. (2003). Conservation of biodiversity in coffee agroecosystems: A tri-taxa comparison in southern Mexico. Biodivers. Conserv., 12, 1239–1252.Google Scholar
Perfecto, I., and Vandermeer, J. (2008). Biodiversity conservation in tropical agroecosystems. Ann. N.Y. Acad. Sci., 1134, 173–200.Google Scholar
Perrin, R. M. (1977). The role of environmental diversity in crop protection. Prot. Ecology, 2, 77–114.Google Scholar
Perry, D. A., Amaranthus, M. P., Borchers, J. G., Borchers, S. L., and Brainerd, R. E. (1989). Bootstrapping in ecosystems. Bioscience, 39(4), 230–237.Google Scholar
Perry, D. A., Orem, R., and Hart, S. C. (2008). Forest Ecosystems. J. Hopkins University Press, Baltimore.
Perry, D. A., Molina, R., and Amaranthus, M. P. (1987). Mycorrhizae, mycorrhizospheres, and reforestation: Current knowledge and research needs. Can. J. For. Res., 17, 929–940.Google Scholar
Persha, L., Agrawal, A., and Chhatre, A. (2011). Social and ecological synergy: Local rulemaking, forest livelihoods, and biodiversity conservation. Science, 331(6024), 1606–1608.Google Scholar
Petheram, C., Zhang, L., Walker, G. R., and Grayson, R. (2000). Towards a framework for predicting impacts of land-use on recharge. 1. A review of recharge studies in Australia, Aust. J. Soil Res., 40, 397–417.Google Scholar
Petraitis, P. (2013). Multiple Stable States in Natural Ecosystems. Oxford University Press, Oxford.
Pfaff, A., and Walker, R. (2010). Regional interdependence and forest “transitions”: Substitute deforestation limits the relevance of local reversals. Land Use Policy, 27(2), 119–129.Google Scholar
Pfaff, A. S. (1996). What drives deforestation in the Brazilian Amazon? MIT Joint Program on the Science and Policy of Global Change, pp. 16.
Phelps, J., Carrasco, L. R., Webb, E. L., Koh, L. P., and Pascual, U. (2013). Agricultural intensification escalates future conservation costs. Proceedings of the National Academy of Sciences, 110(19), 7601–7606.Google Scholar
Phillips, O. L. et al. (1998). Changes in the carbon balance of tropical forests: Evidence from long-term plots. Science, 282, 439–442.Google Scholar
Pielke, R. A. et al. (2011). Land use/land cover changes and climate: Modelling analysis and observational evidence, WIREs Clim. Change, doi:10.1002/wcc.144.
Pimentel, D., and Kounang, N. (1998). Ecology of soil erosion in ecosystems. Ecosystems, 1: 416–426.
Pimm, S. L., and Brooks, T. M. (2000). The sixth extinction: How large, where, and when. In Nature and Human Society: The Quest for a Sustainable World. National Academy Press, Washington DC, pp. 46–62.
Pimm, S. L., and Raven, P. (2000). Biodiversity: Extinction by numbers. Nature, 403(6772), 843–845.Google Scholar
Pockman, W. T., and Sperry, J. S. (1997). Freezing-induced xylem cavitation and the northern limit of Larrea tridentata. Oecologia, 109, 19–27.Google Scholar
Poeplau, C., Don, A., Vesterdal, L., Leifeld, J., Van Wesemael, B. A. S., Schumacher, J., and Gensior, A. (2011). Temporal dynamics of soil organic carbon after land-use change in the temperate zone-carbon response functions as a model approach. Global Change Biology, 17(7), 2415–2427.Google Scholar
Poffenberger, M. (1989). The deforestation of South Asia: History of management conflicts. In Environmental Education and Sustainable Development. Indian Environmental Soc., New Delhi, pp. 155–172.
Pointing, C. (1991). A New Green History of the World: The Environment and the Collapse of Great Civilizations. Penguin Books, New York.
Pommerening, A., and Murphy, S. T. (2004). A review of the history, definitions and methods of continuous cover forestry with special attention to afforestation and restocking. Forestry, 77, 27–44.Google Scholar
Pongratz, J., Reick, C., Raddatz, T., and Claussen, M. (2008). A reconstruction of global agricultural areas and land cover for the last millennium. Global Biogeochemical Cycles, 22(3).Google Scholar
Popp, J., Lakner, Z., Harangi-Rakos, M., and Fari, M. (2014). The effect of bioenergy expansion: Food, energy, and environment. Renewable and Sustainable Energy Reviews, 32, 559–578.Google Scholar
Porporato, A., D'Odorico, P., Laio, F., and Rodriguez-Iturbe, I. (2003). Hydrologic controls on soil carbon and nitrogen cycles. I. Modeling scheme. Advances in Water Resources, 26(1), 45–58.Google Scholar
Porter-Bolland, L., Ellis, E. A., Guariguata, M. R., Ruiz-Mallén, I., Negrete-Yankelevich, S., and Reyes-García, V. (2012). Community managed forests and forest protected areas: An assessment of their conservation effectiveness across the tropics. Forest Ecology and Management, 268, 6–17.Google Scholar
Poschl, U. et al. (2010). Rainforest Aerosols as Biogenic Nuclei of Clouds and Precipitation in the Amazon. Science, 329, 1513–1516.Google Scholar
Potapov, P., Hansen, M. C., Stehman, S. V., Loveland, T. R., and Pittman, K. (2008). Combining MODIS and Landsat imagery to estimate and map boreal forest cover loss. Remote Sensing of Environment, 112(9), 3708–3719.Google Scholar
Potter, C. S., Matson, P. A., Vitousek, P. M., and Davidson, E. A. (1996). Process modeling of controls on nitrogen trace gas emissions from soils worldwide. Journal of Geophysical Research: Atmospheres (1984–2012), 101(D1), 1361–1377.Google Scholar
Pounds, J. A., Fogden, M. P. L., and Campbell, J. H. (1999). Biological response to climate change on a tropical mountain. Nature, 398, 611–614.Google Scholar
Prandini, L., Guidiini, G., Bottura, J. A., Pancano, W. L., and Santos, A. R. (1977). Behavior of the vegetation in slope stability: A critical review. Bulletin of Engineering Geology and the Environment, 16(1), 51–55.Google Scholar
Prather, M., Derwent, R., Ehhalt, D., Fraser, P., Sanhueza, E., and Zhou, X. (1994). Other trace gases and atmospheric chemistry. Climate Change, 94, 77–126.Google Scholar
Pressland, A. J. (1976). Soil moisture distribution as affected by throughfall and stemflow in an arid zone shrub community. Aust. J. Bot., 24, 641–649.Google Scholar
Pretty, J. (2008). Agricultural sustainability: Concepts, principles and evidence. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 363(1491), 447–465.Google Scholar
Price, T. D. (2000). Europe's First Farmers. Cambridge University Press, Cambridge.
Pritchard, S. G., and Rogers, H. H. (2000). Spatial and temporal deployment of crop roots in CO2-enriched environments. New Phytologist, 147(1), 55–71.CrossRefGoogle Scholar
Putz, F. E., Redford, K. H. (2010). The importance of defining “forest”: Tropical forest degradation, deforestation, long-term phase shifts, and further transitions. Biotropica, 42, 10–20.Google Scholar
Qadir, M., Ghafoor, A., and Murtaza, G. (2000). Amelioration strategies for saline soils: A review. Land. Degrad. Dev., 11, 501–21.Google Scholar
Qadir, M., Oster, J. D., Schubert, S., Noble, A. D., and Sahrawat, K. L. (2007). Phytoremediation of Sodic and Saline-Sodic Soils. Advances in Agronomy, 96, 197–247.
Raison, R. J., Khanna, P. K., and Woods, P. (1985). Mechanisms of element transfer to the atmosphere during vegetation burning. Can. J. For. Res., 15, 132–140.Google Scholar
Ramankutty, N., and Foley, J. A. (1999). Estimating historical changes in global land cover: Croplands from 1700 to 1992. Global Biogeochemical Cycles, 13(4), 997–1027.Google Scholar
Ramankutty, N., Gibbs, H. K., Achard, F., Defries, R., Foley, J. A., and Houghton, R. A. (2007). Challenges to estimating carbon emissions from tropical deforestation. Global Change Biology, 13(1), 51–66.Google Scholar
Randerson, J. T., Chapin III, F. S., Harden, J. W., Neff, J. C., and Harmon, M. E. (2002). Net ecosystem production: A comprehensive measure of net carbon accumulation by ecosystems. Ecological Applications, 12(4), 937–947.Google Scholar
Rasmussen, M. et al. (2011). An Aboriginal Australian genome reveals separate human dispersals into Asia. Science, 334, 94–98.Google Scholar
Rausser, G. C., and Small, A. A. (2000). Valuing research leads: Bioprospecting and the conservation of genetic resource. J. Polit. Economy, 108, 173–206.Google Scholar
Ray, D. K., Mueller, N. D., West, P. C., and Foley, J. A. (2013). Yield trends are insufficient to double global crop production by 2050. PLoS ONE, 8(6), e66428, doi:10.1371/journal.pone.0066428.Google Scholar
Ray, D. K., Nair, U. S., Lawton, R. O., Welch, R. M., and PielkeSr., R. A. (2006). Impact of land use on Costa Rican tropical montane cloud forests: Sensitivity of orographic cloud formation to deforestation in the plains. J. Geophys. Res., 111, D02108, doi:10.1029/2005JD006096.Google Scholar
Ray, D. K., Ramankutty, N., Mueller, N. D., West, P. C., and Foley, J. A. (2012). Recent patterns of crop yield growth and stagnation. Nat. Comm., 3, 1293, doi:10.1038/ncomms2296.Google Scholar
Raynor, G. S. (1971). Wind and temperature structure in a coniferous forest and a contiguous field. Forest Science, 17, 351–363.Google Scholar
Redfield, G. W. (2002). Atmospheric deposition of phosphorus to the Everglades: Concepts, constraints and published deposition rates for ecosystem management. Scientific World Journal, 2, 1843–1873.Google Scholar
Redo, D. J., Grau, H. R., Aide, T. M., and Clark, M. L. (2012). Asymmetric forest transition driven by the interaction of socioeconomic development and environmental heterogeneity in Central America. Proceedings of the National Academy of Sciences, 109(23), 8839–8844.Google Scholar
Reed, S. C., Townsend, A. R., Taylor, P. G., and Cleveland, C. C. (2011). Phosphorus cycling in tropical forests growing on highly weathered soils. In Buenemann, E. K., Oberson, A., and Frossard, E. (eds.), Phosphorus in Action: Biological Processes in Phosphorus Cycling, Soil Biology. 26, Springer, Heidelberg, pp. 339–369.
Regan, H. M., Lupia, R., Drinnan, A. N., and Burgman, M. A. (2001). The currency and tempo of extinction. American Naturalist, 157(1), 1–10.Google Scholar
REN21. (2008). Renewable 2007 Global Status Report. REN21 Secretariat, Paris and Washington, DC.
REN21. (2013). Renewables 2013 Global Status Report. REN21 Secretariat, Paris, p. 177.
Renaud, V., and Rebetez, M. (2009). Comparison between open-site and below-canopy climatic conditions in Switzerland during the exceptionally hot summer of 2003. Agricultural and Forest Meteorology, 149, 873–880.Google Scholar
Rengasamy, P. (2006). World salinization with emphasis on Australia. J. Exp. Bot., 57(5), 1017–1023.Google Scholar
Rengasamy, P., Chittleborough, D., and Helyar, K. (2003). Root-zone constraints and plant-based solutions for dryland salinity. Plant and Soil, 257, 249–260.Google Scholar
Renssen, H., Goosse, H., and Fichefet, T. (2003). On the non-linear response of the ocean thermohaline circulation to global deforestation. Geophysical Research Letters, 30(2).Google Scholar
Resende, J. C. F., Markewitz, D., Klink, C. A., Bustamante, M. M., and Davidson, E. A. (2010). Phosphorus cycling in a small watershed in the Brazilian Cerrado: Impacts of frequent burning. Biogeochemistry, doi:10.1007/s10533-010-9531-5.Google Scholar
RFA (Renewable Fuels Association). (2010). Statistics 2010, available at: http://www.ethanolrfa.org/pages/statistics.
Rice, R. M. (1977). Forest management to minimize landslide risk. Guidelines for Watershed Management, 271–287.Google Scholar
Ricker, M., Mendelsohn, R. O., Daly, D. C., and Angeles, G. (1999). Enriching the rainforest with native fruit trees: An ecological and economic analysis in Los Tuxtlas (Veracruz, Mexico). Ecological Economics, 31(3), 439–448.Google Scholar
Ricklefs, R. (2008). The Economy of Nature. 6th ed., W. H. Freeman, pp. 700
Ridolfi, L., D'Odorico, P., and Laio, F. (2006). Effect of vegetation-water table feedbacks on the stability and resilience of plant ecosystems. Water Resources Research, 42, W01201, doi:10.1029/2005WR004444.Google Scholar
Ridolfi, L, D'Odorico, P., and Laio, F. (2007). Vegetation dynamics induced by phreatophyte-water table interactions. J. Theor. Biol., 248, 301–310.Google Scholar
Ridolfi, L., D'Odorico, P., and Laio, F. (2011). Noise-Induced Phenomena in the Environmental Sciences. Cambridge University Press, New York.
Ridolfi, L., D'Odorico, P., Laio, F., Tamea, F. S., and Rodriguez-Iturbe, I. (2008). Coupled stochastic dynamics of water table and soil moisture in bare soil conditions. Water Resour. Res., 44, W01435, doi:10.1029/2007WR006707.Google Scholar
Riekerk, H. (1989). Influence of silvicultural practices on the hydrology of pine flatwoods in Florida. Water Resour. Res., 25(4), 713–719.Google Scholar
Rietkerk, M., and van de Koppel, J. (1997). Alternate stable states and threshold effects in semiarid grazing systems. Oikos, 79, 69–76.Google Scholar
Robinson, B. E., Holland, M. B., and Naughton-Treves, L. (2013). Does secure land tenure save forests? A meta-analysis of the relationship between land tenure and tropical deforestation. Global Environmental Change. 29, 281-293.
Robertson, F. A., Myers, R. J., and Sagna, P. G. (1993). Carbon and nitrogen mineralisation in cultivated and grassland soils in subtropical grassland. Australian Journal of Soil Research, 31, 611–619.Google Scholar
Robichaud, P. R., and Hungerford, R. D. (2000). Water repellency by laboratory burning of four northern Rocky Mountain forest soils. Journal of Hydrology, 231–232, 207–219.Google Scholar
Rodrigues, A. S., Ewers, R. M., Parry, L., Souza, C., Veríssimo, A., and Balmford, A. (2009). Boom-and-bust development patterns across the Amazon deforestation frontier. Science, 324(5933), 1435–1437.Google Scholar
Rodrigues, R. R., Gandolfi, S., Nave, A. G., Aronson, J., Barreto, T. E., Vidal, C. Y., and Brancalion, P. H. S. (2011). Large-scale ecological restoration of high-diversity tropical forests in SE Brazil. For. Ecol. Manage., 261, 1605–1613.Google Scholar
Rodriguez-Iturbe, I., and Rinaldo, A. (1997) Fractal River Basins: Chance and Self-Organization. Cambridge Univ. Press, New York.
Roering, J. J., Schmidt, K. M., Stock, J. D., Dietrich, W. E., and Montgomery, D. R. (2003). Shallow landsliding, root reinforcement, and the spatial distribution of trees in the Oregon Coast Range. Canadian Geotechnical Journal, 40, 237–253.Google Scholar
Röhrig, E., and Ulrich, B. (1991) Ecosystems of the world. 7. Temperate Deciduous Forests. Elsevier, Amsterdam.
Rojstaczer, S., Sterling, S. M., Moore, N. J. (2001). Human appropriation of photosynthesis products. Science, 294, 2549–2552.Google Scholar
Rolett, B., and Diamond, J. (2004). Environmental predictors of pre-European deforestation on Pacific islands. Nature, 431(7007), 443–446.Google Scholar
Ronco, F. (1970). Influence of high light intensity on survival of planted Engelmann spruce. Forest Science, 16, 331–339.Google Scholar
Rosado, B. H. P., and Holder, C. (2012). The significance of leaf water repellency in ecohydrological research: A review. Ecohydrology, doi:10.1002/eco.1340.Google Scholar
Rosenfeld, D. (1999). TRMM observed first direct evidence of smoke from forest fires inhibiting rainfall. Geophysical Res. Lett. 26(20), 3105–3108.Google Scholar
Rosenqvist, A., Shimada, M., Igarashi, T., Watanabe, M., Tadono, T., and Yamamoto, H. (2003, July). Support to multi-national environmental conventions and terrestrial carbon cycle science by ALOS and ADEOS-II-the Kyoto and Carbon Initiative. In International Geoscience and Remote Sensing Symposium, Vol. 3, pp. III-1471.
Ross, D. J., Kelliher, F. M., and Tate, K. R. (1999). Microbial processes in relation to carbon, nitrogen and temperature regimes in litter and a sandy mineral soil from a central Siberian Pinus sylvestris L. forest. Soil Biology and Biochemistry, 31(5), 757–767.Google Scholar
Roy, V., Ruel, J. C., and Plamondon, A. C. (2000). Establishment, growth and survival of natural regeneration after clearcutting and drainage on forested wetlands. Forest Ecology and Management, 129, 253–267.Google Scholar
Royama, T. (1984). Population dynamics of the spruce budworm Choristoneura fumiferana. Ecological Monographs, 429–462.Google Scholar
Royama, T., MacKinnon, W. E., Kettela, E. G., Carter, N. E., and Hartling, L. K. (2005). Analysis of spruce budworm outbreak cycles in New Brunswick, Canada, since 1952. Ecology, 86(5), 1212–1224.Google Scholar
Royer, P. D., Cobb, N. S., Clifford, M. J., Huang, C., Breshears, D. D., Adams, H. D., and Villegas, J. C. (2011). Extreme climatic event-triggered overstorey vegetation loss increases understorey solar input regionally: Primary and secondary ecological implications. Journal of Ecology, 99, 714–723.Google Scholar
Rozema, J., and Flowers, T. J. (2008). Crops for a salinized world. Science, 322(5907), 1478–1480.Google Scholar
Ruben, R., Kruseman, G., and Hengsdijk, H. (1994). Farm household modelling for estimating the effectiviness of price instruments on sustainable land use in the Atlantic Zone of Costa Rica. DLV report No. 4. AB-DLO, Wageningen.
Rudel, T. K. (1998). Is there a forest transition? Deforestation, reforestation, and development. Rural Sociology, 63(4), 533–552.Google Scholar
Rudel, T. K. (2005). Tropical Forests – Regional Paths of Destruction and Regeneration in the Late Twentieth Century. New York: Columbia University Press, p. 237.
Rudel, T. K. (2012). The human ecology of regrowth in the tropics. Journal of Sustainable Forestry, 31, 4–5, 340–354.Google Scholar
Rudel, T. K. (2013). The national determinants of deforestation in sub-Saharan Africa. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 368(1625), 20120405.Google Scholar
Rudel, T. K., Coomes, O. T., Moran, E., Achard, F., Angelsen, A., Xu, J., and Lambin, E. (2005). Forest transitions: Towards a global understanding of land use change. Global Environmental Change, 15(1), 23–31.Google Scholar
Rudel, T. K., Defries, R., Asner, G. P., and Laurance, W. F. (2009a). Changing drivers of deforestation and new opportunities for conservation. Conserv. Biol., 23, 1396–1405.Google Scholar
Rudel, T. K. et al. (2009b). Agricultural intensification and changes in cultivated areas, 1970–2005. Proceedings of the National Academy of Sciences, 106(49), 20675–20680.Google Scholar
Runge-Metzger, A. (1995). Closing the cycle: Obstacles to efficient P management for improved global food security. InTiessen, H. (ed.), Phosphorus in the Global Environment: Transfers, Cycles and Management. Wiley, New York, pp. 27–42.
Runyan, C. W., and D'Odorico, P. (2010). Ecohydrological feedbacks between salt accumulation and vegetation dynamics: The role of vegetation-groundwater interactions. Water Resources Research, 46, W11561, doi:10.1029/2010WR009464.Google Scholar
Runyan, C. W., and D'Odorico, P. (2012). Hydrologic controls on phosphorus dynamics: A modeling framework. Advances in Water Resources, 35, 94–109.Google Scholar
Runyan, C. W., and D'Odorico, P. (2013). Positive feedbacks and bistability associated with phosphorus-vegetation-microbial interactions. Advances in Water Resources, 52, 151–164.Google Scholar
Runyan, C. W., and D'Odorico, P. (2014). Bistable dynamics between vegetation disturbance and landslide occurrence. Water Resources Research, 50(2), 1112–1130, DOI: 10.1002/2013WR014819.Google Scholar
Runyan, C. W., D'Odorico, P., and Lawrence, D. (2012a). Physical and biological feedbacks of deforestation. Reviews of Geophysics, 50(4).Google Scholar
Runyan, C. W., D'Odorico, P., and Lawrence, D. L. (2012b). The effect of repeated deforestation on vegetation dynamics for phosphorus limited ecosystems. Journal of Geophysical Research, 117, G01008, doi:10.1029/2011JG001841.Google Scholar
Runyan, C. W., D'Odorico, P., and Shobe, W. (2015). The economic impacts of positive feedbacks associated with deforestation, Ecological Economics, 120, 93-99, doi:10.1016/j.ecolecon.2015.10.007.
Runyan, C. W., and D'Odorico, P., Vandecar, K. L., Das, R., Schmook, B., and Lawrence, D. (2013). Positive feedbacks between phosphorus deposition and forest canopy trapping, evidence from Southern Mexico. Journal of Geophysical Research-Biogeosciences, 118(4), 1521–1531.Google Scholar
Ruprecht, J. K., and Schofield, N. J. (1989). Analysis of streamflow generation following deforestation in southwest western Australia. J. Hydrol., 105, 1–17.Google Scholar
Saa, A., Trasar-Cepeda, M. C., Gil-Sotres, F., and Carballas, T. (1993). Changes in soil phosphorus and acid phosphatase activity immediately following forest fires. Soil Biology and Biochemistry, 25(9), 1223–1230.Google Scholar
Saad, S. I., da Rocha, H. R., Silva Dias, M. A. F., and Rosolem, R. (2010). Can the deforestation breeze change the rainfall in Amazonia? A case study for the BR-163 highway region. Earth Interact., 14, 18.Google Scholar
Saatchi, S. S. et al. (2011). Benchmark map of forest carbon stocks in tropical regions across three continents. Proceedings of the National Academy of Sciences, 108(24), 9899–9904.Google Scholar
Sahin, V., and Hall, M. J. (1996). The effects of afforestation and deforestation on water yields. Journal of Hydrology, 178(1), 293–309.Google Scholar
Sala, O. E., Meyerson, L. A., and Parmesan, C. (2009). Biodiversity Change and Human Health. Island Press, Washington, DC.
Sala, O. E. et al. (2000). Global biodiversity scenarios for the year 2100. Science, 287(5459), 1770–1774.Google Scholar
Salati, E., Dall'Olio, A., Matsui, E., and Gat, J. R. (1979). Recycling of water in the Amazon Basin: An isotopic study. Water Resour. Res., 15(5), 1250–1258, doi:10.1029/WR015i005p01250.Google Scholar
Samarakoon, A. B., and Gifford, R. M. (1995). Soil water content under plants at high CO2 concentration and interactions with the direct CO2 effects: A species comparison. Journal of Biogeography, 193–202.Google Scholar
Sample, E. C., Soper, R. J., and Racz, G. J. (1980). Reaction of phosphate fertilizers in soils. In Khasawneh, F. E., Sample, A. C., and Kamprath, E. J., (eds.),The Role of Phosphorus in Agriculture.American Society of Agronomy Crop Science Society of America, Soil Science Society of America. Madison, WI, pp. 263–310.
Santiago, L. S. (2007). Extending the leaf economics spectrum to decomposition: Evidence from a tropical forest. Ecology, 88(5), 1126–1131.Google Scholar
Sardans, J., and Peñuelas, J. (2012). The role of plants in the effects of global change on nutrient availability and stoichiometry in the plant-soil system. Plant Physiology, 160(4), 1741–1761.Google Scholar
Sarraf, M. (2004). Assessing the costs of environmental degradation in the Middle East and North Africa countries. Environment Strategy Notes 9. Environment Department, World Bank, Washington DC.
Savage, S. M., Osborn, J., Letey, J., Heaton, C. (1972). Substances contributing to fire-induced water repellency in soils. Soil Science Society America Proceedings, 36, 674–678.Google Scholar
Scheffer, M., Carpenter, S., Foley, J. A., Folke, C., and Walker, B. H. (2001). Catastrophic shifts in ecosystems. Nature, 413, 591–596.Google Scholar
Scheffer, M., Holgren, M., Brovkin, V., and Claussen, M. (2005). Synergy between small- and large-scale feedbacks of vegetation on the water cycle. Global Change Biology, 11, 1003–1012, doi:10.1111/j.1365-2486.2005.00962.x.Google Scholar
Scheffer, M. et al. (2009). Early-warning signals for critical transitions. Nature, 461(7260), 53–59.Google Scholar
Schimel, J. P., Gulledge, J. M., Clein-Curley, J. S., Lindstrom, J. E., and Braddock, J. F. (1999). Moisture effects on microbial activity and community structure in decomposing birch litter in the Alaskan taiga. Soil Biology and Biochemistry, 31(6), 831–838.Google Scholar
Schlesinger, W. (1997). Biogeochemistry: An Analysis of Global Change. Academic Press, San Diego.
Schlesinger, W. H., and Hartley, A. E. (1992). A global budget for atmospheric NH3. Biogeochemistry, 15(3), 191–211.Google Scholar
Schlesinger, W. H, and Pilmanis, A. M. (1998). Plant-soil interactions in deserts. Biogeochemistry, 42, 169–187.Google Scholar
Schlesinger, W. H. et al. (1990). Biological feedbacks in global desertification. Science, 247, 1043–1048.Google Scholar
Schmitt, C. B. et al. (2008). Global ecological forest classification and forest protected area gap analysis. Analyses and recommendations in view of the 10% target for forest protection under the Convention on Biological Diversity (CBD). University of Freiburg, Freiburg.
Scholes, R. J., and Archer, S. R. (1997). Tree-grass interactions in savannas. Annu. Rev. Ecol. Syst., 28, 517–544.Google Scholar
Schoups, G., Hopmans, J. W., Young, C. A., Vrugt, J. A., Wallender, W. W., Tanji, K. K., and Panday, S. (2005). Sustainability of irrigated agriculture in the San Joaquin Valley, California. Proceedings of the National Academy of Sciences, 102(43), 15352–15356.Google Scholar
Schuur, E. A. G. et al. (2008). Vulnerability of permafrost carbon to climate change: Implications for the global carbon cycle. BioScience, 58(8), 701–714.Google Scholar
Searchinger, T. D. et al. (2009). Fixing a critical climate accounting error. Science, 326(5952), 527.Google Scholar
Serraõ, E. A. S., and Falesi, I. C. (1977). Pastagens do Trópico Úmido Brasileiro. Empresa Brasileira de Pesquisa Agropecuária-Centro de Pesquisas Agro-Pecuárias do Trópico Úmido (EMBRAPA- CPATU), Belém, Pará, Brazil.
Seto, K. C., Güneralp, B., and Hutyra, L. R. (2012). Global forecasts of urban expansion to 2030 and direct impacts on biodiversity and carbon pools. Proceedings of the National Academy of Sciences, 109(40), 16083–16088.Google Scholar
Seubert, C., Sanchez, P., and Valverde, C. (1977). Effects of land clearing methods on soil properties of an ultisol and crop performance in the Amazon jungle of Peru. Tropical Agriculture (Trinidad), 54, 307–321.Google Scholar
Sharratt, B. S. (1998), Radiative exchange, near-surface temperature and soil water of forest cropland in interior Alaska. Agricultural and Forest Meteorology, 89, 269–280.Google Scholar
Shimabukuro, Y. et al. (2007). Near real time detection of deforestation in the Brazilian Amazon using MODIS imagery. Ambiente & Água-An Interdisciplinary Journal of Applied Science, 1(1), 37–47.Google Scholar
Shimada, M., Itoh, T., Motooka, T., Watanabe, M., Shiraishi, T., Thapa, R., and Lucas, R. (2014). New global forest/non-forest maps from ALOS PALSAR data (2007–2010). Remote Sensing of Environment, 155, 13–31.
Shimada, M., Rosenqvist, A., Watanabe, M., and Tadono, T. (2005). The polarimetric and interferometric potential of ALOS PALSAR. In ESA Special Publication, Vol. 586, p. 41.
Shively, G. E., and Pagiola, S. (2004). Agricultural Intensification, local labor markets, and deforestation in the Philippines. Environment and Development Economics, 9(02), 241–266.Google Scholar
Shukla, J., Nobre, C., and Sellers, P. (1990). Amazon deforestation and climate change. Science, 247, 1322–1325.Google Scholar
Shuttleworth, W. J. (1977). The exchange of wind-driven fog and mist between vegetation and the atmosphere. Boundary-Layer Meteorology, 12, 463–489.Google Scholar
Sidle, R. C., Pearce, A. J., and O'Loughlin, C. L. (1985). Hillslope Stability and Land Use. Water Resour. Monogr. Ser., 11.Google Scholar
Sidle, R. C., and Swanston, D. N. (1982). Analysis of a small debris slide in coastal Alaska. Can. Geotech. J., 19, 167–174.Google Scholar
Sidle, R. C., Ziegler, A. D., Negishi, J. N., Nik, A. R., Siew, R., and Turkelboom, F. (2006). Erosion processes in steep terrain – truths, myths, and uncertainties related to forest management in Southeast Asia. Forest Ecology and Management, 224(1), 199–225.Google Scholar
Sills, E., and Pattanayak, S. (2006). Tropical trade-offs: An economic perspective on tropical forests. In Spray, S. and McGlothlin, K (eds.), Tropical Deforestation. Rowman & Littlefield, New York.
Silver, W. (1994). Is nutrient availability related to plant nutrient use in humid tropical forests?Oecologia (Berl), 98, 336–343.Google Scholar
Silver, W. L., and Miya, R. K. (2001). Global patterns in root decomposition: Comparisons of climate and litter quality effects. Oecologia, 129(3), 407–419.Google Scholar
Simpson, R. D., Sedjo, R. A., and Reid, J. W. (1996). Valuing biodiversity for use in pharmaceutical research. J. Polit. Economy 104, 163–185.Google Scholar
Singer, B., de Castro, M. C. (2006). Enhancement and suppression of malaria in the Amazon. Am. J. Trop. Med. Hyg., 74(1), 1–2.Google Scholar
Skiba, U. et al. (2012). UK emissions of the greenhouse gas nitrous oxide. Philosophical Transactions of the Royal Society B: Biological Sciences, 367(1593), 1175–1185.Google Scholar
Skidmore, T. E., and Smith, P. H. (1984). Modern Latin America. Oxford University Press, New York and Oxford, 39–56.
Skole, D. L., Chomentowski, W. H., Salas, W. A., and Nobre, A. D. (1994). Physical and human dimensions of deforestation in Amazonia. BioScience, 314–322.Google Scholar
Skopp, J., Jawson, M. D., and Doran, J. W. (1990). Steady-state aerobic microbial activity as a function of soil water content. Soil Sci. Soc. Am. J., 54, 1619–1625.Google Scholar
Smil, V. (1999). Nitrogen in crop production: An account of global flows. Global Biogeochemical Cycles, 13(2), 647–662.Google Scholar
Smil, V. (2000). Phosphorus in the environment: Natural flows and human interferences. Annual Review of Energy and the Environment, 25(1), 53–88.Google Scholar
Smith, S. E., Smith, F. A., and Jakobsen, I. (2004). Functional diversity in arbuscular mycorrhizal (am) symbioses: The contribution of the mycorrhizal P uptake pathway is not correlated with mycorrhizal responses in growth or total P uptake. New Phytol., 162, 511–524.
Smith, W. K., Germino, M. J., Hancock, T. E., and Johnson, D. M. (2003). Another perspective on altitudinal limits of alpine timberlines. Tree Physiology, 23, 1101–1112Google Scholar
Soares-Filho, B. S. et al. (2006). Modelling conservation in the Amazon basin. Nature, 440(7083), 520–523.Google Scholar
Sollins, P., and McCorison, F. M. (1981). Nitrogen and carbon solution chemistry of an old growth coniferous forest watershed before and after cutting. Water Resources Research, 17(5), 1409–1418.Google Scholar
Solow, R. M. (1974). Intergenerational equity and exhaustible resources: Review of economic studies. symposium of the economics of exhaustible resources, 29–46.
Sorenson, S. K., Dileanis, P. D., and Branson, F. A. (1991). Soil water and vegetation responses to precipitation and changes in depth to ground water in Owens Valley, California. USGS, Denver 54.
Southgate, D. (1990). The causes of land degradation along “spontaneously” expanding agricultural frontiers in the Third World. Land Economics, 93–101.Google Scholar
Souza, E. P., Renno, N. O., Dias, M. A. F. Silva (2000). Convective circulations induced by surface heterogeneities. J. Atmos. Sci., 57, 2915–2922.Google Scholar
Spracklen, D. V., Arnold, S. R., and Taylor, C. M. (2012). Observations of increased tropical rainfall preceded by air passage over forests. Nature, 489, 282–286.Google Scholar
Stark, N. (1972). Nutrient cycling pathways and litter fungi. Bioscience, 355–360.Google Scholar
Staver, A. C., Archibald, S., and Levin, S. A. (2011). The global extent and determinants of savanna and forest as alternative stable states. Science, 334, 230–232.Google Scholar
Steadman, D. W. (1995). Prehistoric extinctions of Pacific island birds: Biodiversity meets zooarchaeology. Science, 267(5201), 1123–1131.Google Scholar
Stednick, J. D. (1996). Monitoring the effects of timber harvest on annual water yield. Journal of Hydrology, 176(1/4), 79–95.Google Scholar
Steininger, M. K., Tucker, C. J., Townshend, J. R. G., Killeen, T. J., Desch, A., Bell, V., and Ersts, P. (2001). Tropical deforestation in the Bolivian Amazon. Environmental Conservation, 28, 127–134.Google Scholar
Stevens, G. C., and Fox, J. F. (1991). The causes of treeline. Annual Review of Ecology and Systematics, 22, 177–191.Google Scholar
Stevenson, C.M., et al. (2015). Variation in Rapa Nui (Easter Island) land use indicates production and population peaks prior to European contact. PNAS, 112(4), 1025–1030; DOI: 10.1073/pnas.1420712112.
Stokes, A. et al. (2008). How vegetation reinforces soil on slopes. In Norris, J. E., Stokes, A., Mickovski, S. B., Cammeraat, E., Beek, L. P. H. van, Nicoll, B., and Achim, A. (eds.), Slope Stability and Erosion Control: Ecotechnological Solutions. Springer, Dordrecht, pp. 65–118.
Stoorvogel, J. J., Smaling, E. M., and Janssen, B. H. (1993). Calculating soil nutrient balances in Africa at different scales I. Supra-National Scale, Fertilizer Reseach, 35, 227–235.Google Scholar
Stork, N. E. (2010). Re-assessing current extinction rates. Biodiversity and Conservation, 19(2), 357–371.Google Scholar
Strogatz, S. H. (1994). Nonlinear Dynamics and Chaos. Westview Press, Cambridge, MA.
Stuart, S. A., Choat, B., Martin, K. C., Holbrook, N. M., and Ball, M. C. (2007). The role of freezing in setting the latitudinal limits of mangrove forests. New Phytologist, 173, 576–583.Google Scholar
Sturm, M., Douglas, T., Racine, C., and Liston, G. E. (2005). Changing snow and shrub conditions affect albedo with global implications. Journal of Geophysical Research, 110, G01004, doi:10.1029/2005JG000013.Google Scholar
Sturm, M., Racine, C., and Tape, K. (2001). Increasing shrub abundance in the Arctic. Nature, 411, 546–547.Google Scholar
Suding, K. N., Collins, S. L., Gough, L., Clark, C., Cleland, E. E., Gross, K. L., Milchunas, D. G., and Pennings, S. (2005). Functional- and abundance-based mechanisms explain diversity loss due to N fertilization. Proceedings of the National Academy of Sciences (USA), 102, 4387–4392.Google Scholar
Sun, G. et al. (2001). Effects of forest management on the hydrology of wetland forests in the southern United States. For. Ecol. Manag., 143, 227–236.Google Scholar
Sutton, M. A., Pitcairn, C. E., and Fowler, D. (1993). The Exchange of Ammonia between the Atmosphere and Plant Communities. Academic Press, San Diego.
Suweis, S., and D'Odorico, P. (2014). Early warning signs in social-ecological networks. PLoS-One, 9(7), e101851, doi:10.1371/journal.pone.0101851.Google Scholar
Sveinbjörnsson, B. (2000). North American and European treelines: External forces and internal processes controlling position. AMBIO: A Journal of the Human Environment, 29, 388–395.Google Scholar
Swank, W. T., and Johnson, C. E. (1994). Small catchment research in the evaluation and development of forest management practices. In Moldan, B., and Cerny, J. (eds.), Biogeochemistry of Small Catchments: A Tool for Environmental Research. Wiley, Chichester, pp. 383–408.
Swanson, F. J., and Dyrness, C. T. (1975). Impact of clear-cutting and road construction on soil erosion by landslides in the western Cascade Range, Oregon. Geology, 3(7), 393–396.Google Scholar
Swanston, D. N. (1988). Timber harvest and progressive deformation of slopes in southwestern Oregon. Bulletin of the Association of Engineering Geologists, 25, 371–381.Google Scholar
Swift, M. J., Izac, A. M., and van Noordwijk, M. (2004). Biodiversity and ecosystem services in agricultural landscapes – are we asking the right questions?Agriculture, Ecosystems and Environment, 104(1), 113–134.Google Scholar
Tarafdar, J. C., and Marschner, H. (1994). Phosphatase activity in the rhizosphere and hyphosphere of VA mycorrhizal wheat supplied with inorganic and organic phosphorus. Soil Biol. Biochem., 26, 387–395.Google Scholar
Tatem, A. J., Goetz, S. J., Hay, S. I. (2008). Fifty years of earth observation satellites. American Scientist, 96, 390–398.Google Scholar
Tavoni, A., Schlueter, M., and Levin, S. (2012). The survival of the conformist: Social pressure and renewable resource management. J. Theor. Biol., 299, 152–161.Google Scholar
Terborgh, J. et al. (2001). Ecological meltdown in predator-free forest fragments. Science, 294(5548), 1923–1926.Google Scholar
Terman, G. L. (1980). Volatilization losses of nitrogen as ammonia from surface-applied fertilizers, organic amendments, and crop residues. Advances in Agronomy, 31, 189–223.Google Scholar
Terwilliger, V. J., and Waldron, L. J. (1991). Effects of root reinforcement on soil-slip patterns in the Transverse Ranges of southern California. Geological Society of America Bulletin, 103, 775–785.Google Scholar
Thompson, I. D. et al. (2011). Forest biodiversity and the delivery of ecosystem goods and services: Translating science into policy. BioScience, 61(12), 972–981.Google Scholar
Tilman, D. (1987). Secondary succession and the pattern of plant dominance along experimental nitrogen gradients. Ecological Monographs, 57, 189–214.Google Scholar
Tilman, D. (1999). Global environmental impacts of agricultural expansion: The need for sustainable and efficient practices. Proceedings of the National Academy of Sciences, 96(11), 5995–6000.Google Scholar
Tilman, D., Lehman, C. L., and Bristow, C. E. (1998). Diversity-stability relationships: Statistical inevitability or ecological consequence?American Naturalist, 151(3), 277–282.Google Scholar
Tilman, D. et al. (2001). Forecasting agriculturally driven global environmental change. Science, 292(5515), 281–284.Google Scholar
Tilman, D., Cassman, K. G., Matson, P. A., Naylor, R., and Polasky, S. (2002). Agricultural sustainability and intensive production practices. Nature, 418(6898), 671–677.Google Scholar
Tilman, D., Hill, J., and Lehman, C. (2006). Carbon-negative biofuels from low-input high-diversity grassland biomass. Science, 314(5805), 1598–1600.Google Scholar
Tilman, D., Isbell, F., and Cowles, J. M. (2014). Biodiversity and Ecosystem Functioning. Annual Review of Ecology, Evolution, and Systematics, 45(1), 471.Google Scholar
Torres, R., Dietrich, W. E., Montgomery, D. R., Anderson, S. P., and Loague, K. (1998). Unsaturated zone processes and the hydrologic response of a steep, unchanneled catchment. Water Resour. Res., 34, 1865–1879Google Scholar
Townsend, A. R., Vitousek, P. M., and Holland, E. A. (1992). Tropical soils could dominate the short-term carbon cycle feedbacks to increased global temperatures. Climatic Change, 22(4), 293–303.Google Scholar
Toy, ADF. (1973). The Chemistry of Phosphorus. Pergamon, Oxford.
Tranquillini, W. (1979). Physiological ecology of the alpine timberline: Tree existence at high altitudes with special reference to the European Alps, Springer, Berlin.
Trappe, J. M. (1987). Phylogenetic and ecologic aspects of mycotrophy in the angiosperms from an evolutionary standpoint. In Safir, G. R. (ed.), Ecophysiology of VA Mycorrhizal Plants. CRC, Boca Raton, FL, pp. 5–25.
Trenbath, B. R. (1993). Intercropping for the management of pests and diseases. Field Crops Research, 34(3), 381–405.Google Scholar
Trenberth, K. E. (1999). Atmospheric moisture recycling: Role of advection and local evaporation. J. Climate, 12, 1368–1381.Google Scholar
Troendle, C. A. (1983). The potential for water yield augmentation from forest management in the Rocky Mountain region. Water Resour. Bull., 19(3), 359–373.Google Scholar
Trumbore, S. E., Davidson, E. A., Camargo, P. B., Nepstad, D. C., and Martinelli, L. A. (1995). Below-ground cycling of carbon in forests and pastures of eastern Amazonia. Global Biogeochem. Cycles, 9(4), 515–528.Google Scholar
Trustrum, N. A., and De Rose, R. C. (1988). Soil depth-age relationship of landslides on deforested hillslopes. Taranaki, New Zeland. Geomorphology, 1, 143–160.Google Scholar
Trustrum, N. A., Lambert, M. G., and Thomas, V. J. (1983). The impact of soil slip erosion on hill country pasture production in New Zealand. Proceedings of the Second International Conference on Soil Erosion and Conservation, Honolulu, January 1983.
Tscharntke, T., Klein, A. M., Kruess, A., Steffan-Dewenter, I., and Thies, C. (2005). Landscape perspectives on agricultural intensification and biodiversity–ecosystem service management. Ecology Letters, 8(8), 857–874.Google Scholar
Turazza, D. (1880). Trattato di idrometria o di idraulica pratica. Padova, Italy: F. Sacchetto.
Turner, B. L., Lambin, E. F., and Reenberg, A. (2007b). The emergence of land change science for global environmental change and sustainability. Proceedings of the National Academy of Sciences, 104(52), 20666–20671.
Turner, B. L., and Robbins, P. (2008). Land-change science and political ecology: Similarities, differences, and implications for sustainability science. Annual review of environment and resources, 33, 295–316.
Turner, B.L., and Sabloff, J.A. (2012). Classic Period collapse of the Central Maya Lowlands: Insights about human–environment relationships for sustainability, Proc. Natnl. Acad. Sci., USA, PNAS, 109(35), 13908–13914.
Turner, I. M., Tan, H. T. W., Wee, Y. C., Ibrahim, A. B., Chew, P. T., and Corlett, R. T. (1994). A study of plant species extinction in Singapore: Lessons for the conservation of tropical biodiversity. Conservation Biology, 8(3), 705–712.Google Scholar
Turner, M. G., Smithwick, E. A., Metzger, K. L., Tinker, D. B., and Romme, W. H. (2007a). Inorganic nitrogen availability after severe stand-replacing fire in the Greater Yellowstone ecosystem. Proceedings of the National Academy of Sciences, 104(12), 4782–4789.Google Scholar
Uhl, C., and Kauffman, H. J. B. (1990). Deforestation, fire susceptibility and potential tree responses to fire in the eastern Amazon. Ecology, 71(2), 437–449.Google Scholar
UN Comtrade. (2014). Available at: http://comtrade.un.org/ (accessed on June 25, 2014).
UNEP. (2007). Global environment outlook 4. United Nations Environment Programme, Nairobi, Kenya.
UNEP/GRID Arendal. (2002). Arctic environmental atlas, available at: http://maps.grida.no/arctic/.
UNPD. UN Population Division, New York (2005). Long-range world population projections: Based on the 1998 revision.
Urli, M., Porté, A. J., Cochard, H., Guengant, Y., Burlett, R., and Delzon, S. (2013). Xylem embolism threshold for catastrophic hydraulic failure in angiosperm trees. Tree Physiology, 33(7), 672–683.Google Scholar
USDA-FAS (2010). Oilseeds: World markets and trade. Circular Series FOP 8–10 August 2010. US Depart of Agric-Foreign Agric Serv, Washington, DC.
US Forest Service. (2008). North American forest outlook study: US country report. Unpublished SOFO 2009 contribution.
Valdes, C. (2006). Brazil's booming agriculture faces obstacles, Amber Waves Economic Research Service/USDA, 4, 28–35, available at: www.ers.usda.gov/AmberWaves/November06/.
van de Koppel, Rietkerk, J., M., and Weissing, F. J. (1997). Catastrophic vegetation shifts and soil degradation in terrestrial grazing systems. Trends in Ecology and Evolution, 12, 352–356.Google Scholar
van der Ent, R. J., Savenije, H. H., Schaefli, B., and Steele-Dunne, S. C. (2010). Origin and fate of atmospheric moisture over continents. Water Resour. Res., 46, W09525, doi:10.1029/2010WR009127.Google Scholar
van der Werf, G. R., Randerson, J. T., Giglio, L., Collatz, G. J., Kasibhatla, P. S., and ArellanoJr., A. F. (2006). Interannual variability in global biomass burning emissions from 1997 to 2004. Atmos. Chem. Phys., 6, 3423–3441, doi:10.5194/acp-6-3423-2006.Google Scholar
van der Werf, G. R. et al. (2009). CO2 emissions from forest loss. Nature Geoscience, 2(11), 737–738.Google Scholar
Van Laerhoven, F. (2010). Governing community forests and the challenge of solving two-level collective action dilemmas—a large-N perspective. Global Environmental Change, 20(3), 539–546.Google Scholar
van Langevelde, F., C. et al. (2003). Effects of fire and herbivory on the stability of savanna ecosystems. Ecology, 84, 337–350.Google Scholar
van Meeteren, M. M., Tietema, A., van Loon, E. E., and Verstraten, J. M. (2008). Microbial dynamics and litter decomposition under a changed climate in a Dutch heathland. Applivan meeed Soil Ecology 38, 119–127.Google Scholar
van Nes, E. H., and Scheffer, M. (2007). Slow recovery from perturbations as a generic indicator of a nearby catastrophic shift. Am. Nat., 169, 738–747.Google Scholar
van Noordwijk, M., Cerri, C., Woomer, P. L., Nugroho, K., Bernoux, M. (1997). Soil carbon dynamics in the humid tropical forest zone. Geoderma, 79, 187–225.Google Scholar
van Vuuren, D. P., Sala, O. E., Pereira, H. M. (2006). The future of vascular plant diversity under four global scenarios. Ecol. Soc., 11(2), 25.Google Scholar
van Wilgen, B. W., Cowling, R. M., and Burgers, C. J. (1996). Valuation of ecosystem services. BioScience, 184–189.Google Scholar
van Wilgen, B. W., Trollope, W. S. W, Biggs, H. C., Potgieter, A. L. F., and Brockett, B. H. (2003). Fire as a driver of ecosystem variability. In Du Toit, J. T., Rogers, K. H., and Biggs, H. C. (eds.), The Kruger Experience: Ecology and Management of Savanna Heterogeneity. Island Press, Washington, DC, pp. 149–170.
Vanacker, V., Vanderschaeghe, M., Govers, G., Willems, E., Poesen, J., Deckers, J., and De Bievre, B. (2003). Linking hydrological, infinite slope stability and land-use change models through GIS for assessing the impact of deforestation on slope stability in high Andean watersheds. Geomorphology, 52(3), 299–315.Google Scholar
Vance, C. P., Uhde-Stone, C., and Allan, D. L. (2003). Phosphorus acquisition and use: Critical adaptations by plants for securing a nonrenewable resource. New Phytologist, 157, 423–447.Google Scholar
Vandermeer, J. H. (1989). The Ecology of Intercropping. Cambridge University Press, Cambridge.
Verhoeven, J. T. A., and Schmitz, M. B. (1991). Control of plant growth by nitrogen and phosphorus in mesotrophic fens. Biogeochemistry 12, 135–148.Google Scholar
Vermeer, E. B. (1998). Population and ecology along the frontier in Qing China, in Elvin, M., and Liu, T. -J. (eds.), Sediments of Time, Environment and Society in Chinese History. Cambridge University Press, Cambridge.
Vetaas, O. R. (1992). Micro-site effects of trees and shrubs in dry savannas. J. Veg. Sci., 3(3), 337–344.Google Scholar
Viana, V. M., Tabanez, A. A. (1996). Biology and conservation of forest fragments in the Brazilian Atlantic moist forest. In Schelhas, J., and Greenberg, R. (eds.), Forest Patches in Tropical Landscapes. Island Press, Washington, DC, pp. 151–167.
Villegas, J. C., Breshears, D. D., Zou, C. B., and Royer, P. D. (2010). Seasonally pulsed heterogeneity in microclimate: Phenology and cover effects along deciduous grassland–forest continuum. Vadose Zone Journal, 9, 537–547.Google Scholar
Villoria, N. B., Byerlee, D., and Stevenson, J. (2014). The effects of agricultural technological progress on deforestation: What do we really know?Applied Economic Perspectives and Policy, 36(2), 211–237.Google Scholar
Vitousek, P. M. (1984). Litterfall, nutrient cycling, and nutrient limitation in tropical ecosystems. Ecology, 65, 285–298.Google Scholar
Vitousek, P. M., and Hobbie, S. (2000). Heterotrophic nitrogen fixation in decomposing litter: Patterns and regulation. Ecology, 81(9), 2366–2376.Google Scholar
Vitousek, P. M., and Howarth, R. W. (1991). Nitrogen limitation on land and in the sea: How can it occur?Biogeochemistry, 13(2), 87–115.Google Scholar
Vitousek, P. M., and Matson, P. A. (1985). Disturbance, nitrogen availability, and nitrogen losses in an intensively managed loblolly pine plantation. Ecology, 1360–1376.Google Scholar
Vitousek, P. M., Mooney, H. A., Lubchenco, J., and Melillo, J. M. (1997b). Human domination of Earth's ecosystems. Science, 277(5325), 494–499.Google Scholar
Vitousek, P. M., Porder, S., Houlton, B. Z., and Chadwick, O. A. (2010). Terrestrial phosphorus limitation: Mechanisms, implications and nitrogen-phosphorus interactions. Ecological Applications, 20(1), 5–15.Google Scholar
Vitousek, P. M., Turner, D. R., Parton, W. J., and Sanford, R. L. (1994). Litter decomposition on the Mauna Loa Environmental Matrix, Hawai'i: Patterns, mechanisms and models. Ecology, 75(2), 418–429.Google Scholar
Vitousek, P. M. et al. (1997a). Human alteration of the global nitrogen cycle: Sources and consequences. Ecological Applications, 7(3), 737–750.Google Scholar
Vitt, D. H., Halsey, L. A., and Zoltai, S. C. (1994). The bog landforms of continental Western Canada in relation to climate and permafrost patterns. Arct. Alp. Res., 26, 1–13.Google Scholar
Vitt, D. H., Halsey, L. A., and Zoltai, S. C. (1999). The changing landscape of Canada's western boreal forest: The current dynamics of permafrost. Can. J. For. Res., 30, 283–287.Google Scholar
Vittor, A. Y. et al. (2006). The effect of deforestation on the human-biting rate of Anopheles darlingi, the primary vector of falciparum malaria in the Peruvian Amazon. American Journal of Tropical Medicine and Hygiene, 74(1), 3–11.Google Scholar
Voicu, M. F., and Comeau, P. G. (2006). Microclimatic and spruce growth gradients adjacent to young aspen stands. Forest Ecology and Management, 221, 13–26.Google Scholar
von Randow, C. et al. (2004). Comparative measurements and seasonal variations in energy and carbon exchange over forest and pasture in South West Amazonia. Theor. Appl. Climatol., 78, 5–26.Google Scholar
Von Uexkull, H. R., and Mutert, E. (1995). Global extent, development and economic impact of acid soils. Plant and Soil, 171, 1–15.
Von Uexkull, H. R., and Mutert, E. (1998). Global extent, development and economic impact of acid soils. In Date, R. A., Grundon, N. J., Rayment, G. E., and , M. E. Probert, (eds.), Plant-Soil Interactions at Low pH: Principles and Management. Dordrecht: Kluwer, 5–19.
Vosti, S. A., Witcover, J., and Carpentier, C. L. (2002). Agricultural intensification by smallholders in the western Brazilian Amazon: From deforestation to sustainable land use. Intl Food Policy Res Inst., 130.Google Scholar
Wagener, S. M., and Schimel, J. P. (1998). Stratification of soil ecological processes: A study of the birch forest floor in the Alaskan taiga. Oikos, 63–74.Google Scholar
Wakker, E. (2006). The Kalimantan border oil palm mega-project. 38(19), International Atomic Energy Agency.
Waldrop, M. P., Balser, T. C., and Firestone, M. K. (2000). Linking microbial community composition to function in a tropical soil. Soil. Biol. Biochem., 24, 317–323.Google Scholar
Walker, B., and Salt, D. (2006). Resilience Thinking. Island Press, Wahington, DC.
Walker, B. H., Ludwig, D., Holling, C. S., and Peterman, R. M. (1981). Stability of semiarid savanna grazing systems. J. Ecol., 69, 473–498.Google Scholar
Walker, J., Bullen, F., and Williams, B. (1993). Ecohydrological changes in the Murray-Darling Basin. I. The number of trees cleared over the last two centuries. J. Appl. Ecol., 30(2), 265–273, doi:10.2307/2404628.Google Scholar
Walker, R., (1993). Deforestation and economic development. Canadian Journal of Regional Science, 16, 481–497.Google Scholar
Walker, T. W., and Syers, J. K. (1976). The fate of phosphorus during pedogenesis. Geoderma, 15, 1–19.Google Scholar
Wan, S., Hui, D., and Luo, Y. (2001). Fire effects on nitrogen pools and dynamics in terrestrial ecosystems: A meta-analysis. Ecological Applications, 11(5), 1349–1365.Google Scholar
Wang, G. L., Eltahir, E. A. B. (2000a). Biosphere-atmosphere interactions over West Africa. Part 1: Development and validation of a coupled dynamic model. Q J R Meteorol. Soc., 126, 1239–1260.Google Scholar
Wang, G. L., Eltahir, E. A. B. (2000b). Biosphere-atmosphere interactions over West Africa. Part 2: Multiple climate equilibria. Q J R Meteorol. Soc., 126, 1261–1280.Google Scholar
Wang, G. L. and Eltahir, E. A. B. (2000C), Ecosystem dynamics and the Sahel drought.Geophys. Res. Lett., 27, 795–798.Google Scholar
Wang, J. et al. (2009). Impact of deforestation in the Amazon basin on cloud climatology. Proc. Natl. Acad. Sci., 106, 3670–3674, doi:10.1073/pnas.0810156106.Google Scholar
Wanner, H. J. et al. (2008). Mid- to Late Holocene climate change: An overview. Quaternary Science Reviews, 27, 1791–1828.Google Scholar
Wardle, D. A. (1998). Controls of temporal variability of the soil microbial biomass: A global-scale synthesis. Soil Biol. Biochem., 30(13), 1627–1637.Google Scholar
Warneck, P. (1988). Chemistry of the Natural Atmosphere. International Geophysics, 41, San Diego, California, 757.
Wasilewska, L. (1995). Differences in development of soil nematode communities in single-and multi-species grass experimental treatments. Applied Soil Ecology, 2(1), 53–64.Google Scholar
Wasowski, J. (1998). Understanding rainfall-landslide relationships in man-modified environments: A case-history from Caramanico Terme, Italy. Environmental Geology, 35(2–3), 197–209.Google Scholar
Wassenaar, T., Gerber, P., Verburg, P. H., Rosales, M., Ibrahim, M., and Steinfeld, H. (2007). Projecting land use changes in the Neotropics: The geography of pasture expansion into forest. Global Environmental Change, 17(1), 86–104.Google Scholar
Weathers, K. C., Cadenasso, M. L., and Pickett, S. T. (2001). Forest edges as nutrient and pollutant concentrators: Potential synergisms between fragmentation, forest canopies, and the atmosphere. Conservation Biology, 15, 1506–1514.Google Scholar
Weathers, K. C., and Likens, G. E. (1997). Clouds in Southern Chile: An important source of nitrogen to nitrogen-limited ecosystems?Environ. Sci. Technol., 31, 210–213.Google Scholar
Webb, E. L., Jachowski, N. R. A.Phelps, J., Friess, D. A., Than, M. M., and Ziegler, A. D. (2014). Deforestation in the Ayeyarwady Delta and the conservation implications of an internationally-engaged Myanmar. Global Environmental Change - Human and Policy Dimensions, 24, 321–333.Google Scholar
Webb, T., Bartlein, P. J., and Kutzbach, J. E. (1987). Climatic Change in Eastern Noth America during the past 18,000 years: Comparisons of pollen data with model results in North America and Adjacent Oceans during the last Glaciation. (eds. Ruddiman, W. F., and Wright), H. E., 447-462, Geological Soc. Am., Boulder, Co.
Webb, T. J., Woodward, F. I., Hannah, L., and Gaston, K. J. (2005). Forest cover–rainfall relationships in a biodiversity hotspot: The Atlantic forest of Brazil. Ecological Applications, 15(6), 2005, 1968–1983.Google Scholar
Weitzman, M. L. (1994). On the “environmental” discount rate. Journal of Environmental Economics and Management, 26(2), 200–209.Google Scholar
Werth, D., and Avissar, R. (2002). The local and global effects of Amazon deforestation. J. Geophys. Res., 107(D20), 8087, doi:10.1029/2001JD000717.Google Scholar
West, P. C., Gibbs, H. K., Monfreda, C., Wagner, J., Barford, C. C., Carpenter, S. R., and Foley, J. A. (2010). Trading carbon for food: Global comparison of carbon stocks vs. crop yields on agricultural land. Proceedings of the National Academy of Sciences, 107(46), 19645–19648.Google Scholar
Westfall, J. (2004). 2003 Summary of forest health conditions in British Columbia. Ministry of Forests, Forest Practices Branch.
Westheimer, F. H. (1987). Why nature chose phosphates. Science, 235(4793), 1173–1178.Google Scholar
Wetzel, P. R. et al. (2005). Maintaining tree islands in the Florida Everglades: Nutrient redistribution is the key. Front. Ecol. Environ., 3, 370–376.Google Scholar
Whitford, W. G., Anderson, J., and Rice, P. M. (1997). Stemflow contribution to the “fertile island” effect in cresotebush, Larrea-tridentata. J. Arid Environ., 35, 451–457.Google Scholar
Wieczorek, G. F., Morgan, B. A., and Campbell, R. H. (2000). Debris-flow hazards in the Blue Ridge of central Virginia. Environ. Eng. Geol., 6, 3–23.Google Scholar
Wieder, W. R., Cleveland, C. C., and Townsend, A. R. (2009). Controls over leaf litter decomposition in wet tropical forests. Ecology, 90(12), 3333–3341.Google Scholar
Wilby, A. et al. (2009). Biodiversity, food provision, and human health. In Sala, O. E., Meyerson, L. A., and Parmesan, C. (eds.), Biodiversity Change and Human Health. Island Press, Washington, DC, pp. 13–39.
Wilde, S. A., Steinbrenner, E. C., Pierce, R. S., Dosen, R. C., and Pronin, D. T. (1953). Influence of forest cover on the state of the ground water table. Soil Sci. Soc. Proc., 17, 65–67.Google Scholar
Williams, E. J., Hutchinson, G. L., and Fehsenfeld, F. C. (1992). NOx and N2O emissions from soil. Global Biogeochemical Cycles, 6(4), 351–388.Google Scholar
Williams, M. (1990). Forests. In Turner, B. L., II, Clark, W. C., Kates, R. W., Richards, J. F., Mathews, J. T., and Meyer, W. B. (eds.), The Earth as Transformed by Human Action. Cambridge University Press, Cambridge, pp. 179–201.
Williams, M. (2000). Dark ages and dark areas: Global deforestation in the deep past. Journal of Historical Geography, 26, 28–46.Google Scholar
Williams, M. (2002). Deforesting the Earth: From Prehistory to Global Crisis. University of Chicago Press, Chicago.
Williams, M. R., Fisher, T. R., and Melack, J. M. (1997). Solute dynamics in soil water and groundwater in a central Amazon catchment undergoing deforestation. Biogeochemistry, 38(3), 303–335.Google Scholar
Wilson, B. A., Neldner, V. J., and Accad, A. (2002). The extent and status of remnant vegetation in Queensland and its implications for statewide vegetation management and legislation. Rangeland J., 24, 6–35.Google Scholar
Wilson, J. B., and Agnew, A. D. Q. (1992). Positive-feedback switches in plant communities. Advances in Ecological Research, 23, 263–336.Google Scholar
Wittemyer, G., Elsen, P., Bean, W. T., Burton, C. O., and Brashares, J. S. (2008). Accelerated human population growth at protected area edges. Science, 321, 123–126.Google Scholar
Woodward, C., Shulmeister, J., Larsen, J., Jacobsen, G. E., and Zawadzki, A. (2014). The hydrological legacy of deforestation on global wetlands. Science, 346(6211), 844–847.Google Scholar
WRI. (1997). The Last Frontier Forests ecosystems and economies on the edge. World Resources Institute. Washington DC, USA, 49 pp.
WRI. (2007). EarthTrends: Environmental information. World Resources Institute (WRI), Washington, DC available at: http://earthtrends.wri.org.
Wright, J. M., and Chambers, J. C. (2002). Restoring riparian meadows currently dominated by Artemisia using alternative state concepts – above-ground response. Appl. Veg. Sci., 5, 237–246.Google Scholar
Wright, S. J. (2010). The future of tropical forests. Ann. N.Y. Acad. Sci., 1195, 1–27.Google Scholar
Wu, T. H., McKinnel, W. P., and Swanston, D. N. (1979). Strength of tree-roots and landslides on Prince of Wales Island, Alaska. Can. Geotech. J., 16(1), 19–33.Google Scholar
Wu, T. H., and Swanston, D. N. (1980). Risk of landslides in shallow soils and its relation to clearcutting in southeastern Alaska. For. Sci., 26(3), 495–510.Google Scholar
Wunder, S., and Dermawan, A. (2007). Cross-sectoral tropical forest cover impacts: What matters. Cross-Sectoral Policy Developments in Forestry, 1–14.Google Scholar
Xue, Y., Sellers, P. J., KinterIII, J. L., and Shukla, J. (1991). A simplified biosphere model for global climate studies. J. Climate, 4, 345–364.Google Scholar
Yachi, S., and Loreau, M. (1999). Biodiversity and ecosystem productivity in a fluctuating environment: The insurance hypothesis. Proceedings of the National Academy of Sciences, 96(4), 1463–1468.Google Scholar
Yang, X., Post, W. M., Thornton, P. E., and Jain, A. (2013). The distribution of soil phosphorus for global biogeochemical modeling. Biogeosciences, 10(4), 2525–2537.Google Scholar
Yang, Y. S., Guo, J., Chen, G., Xie, J., Gao, R., Li, Z., and Jin, Z. (2005). Carbon and nitrogen pools in Chinese fir and evergreen broadleaved forests and changes associated with felling and burning in mid-subtropical China. Forest Ecology and Management, 216(1), 216–226.Google Scholar
Yasuoka, J., and Levins, R. (2007). Impact of deforestation and agricultural development on anopheline ecology and malaria epidemiology. Am. J. Trop. Med. Hyg., 76(3), 450–460.Google Scholar
Young, A., and Mitchell, N. (1994). Microclimate and vegetation edge effects in a fragmented podocarp-broadleaf forest in New Zealand. Biological Conservation, 67, 63–72.Google Scholar
Xiao, H., Ouyang, Z., Zhao, J., and Wang, X. (2000). Forest ecosystem services and their ecological valuation – a case study of tropical forest in Jianfengling of Hainan Island. The Journal of Applied Ecology, 11(4), 481–484.Google Scholar
Zak, D. R., Holmes, W. E., Finzi, A. C., Norby, R. J., and Schlesinger, W. H. (2003). Soil nitrogen cycling under elevated CO2: A synthesis of forest FACE experiments. Ecological Applications, 13(6), 1508–1514.Google Scholar
Zavaleta, E. (2000). Valuing ecosystem services lost to Tamarix invasion in the United States. Invasive Species in a Changing World, 261–300.Google Scholar
Zegada-Lizarazu, W., and Monti, A. (2011). Energy crops in rotation: A review. Biomass and Bioenergy, 35(1), 12–25.Google Scholar
Zen, Z., Barlow, C., and Gondowarsito, R. (2006). Oil palm in Indonesian socio-economic improvement: A review of options. Industry Economic Journal, 6, 18–29.Google Scholar
Zeng, N. et al. (1999). Enhancement of interdecadal climate variability in the Sahel by vegetation interaction. Science, 286, 1537–1540.Google Scholar
Zeng, N., and Neelin, J. D. (2000). The role of vegetation-climate interaction and interannual variability in shaping the African savanna. J. Clim. 13, 2665–2670, doi:10.1175/1520-0442(2000)013,2665:trovci.2.0.co;2.Google Scholar
Zeng, X., Shen, S. S., Zeng, X., and Dickinson, R. E. (2004). Multiple equilibrium states and the abrupt transitions in a dynamical system of soil water interacting with vegetation. Geophys. Res. Lett., 31, L05501, doi:10.1029/2003GL018910.Google Scholar
Zentner, R. P. et al. (2002). Economics of crop diversification and soil tillage opportunities in the Canadian prairies. Agronomy Journal, 94(2), 216–230.Google Scholar
Zhang, D., Hui, D., Luo, Y., and Zhou, G. (2008). Rates of litter decomposition in terrestrial ecosystems: Global patterns and controlling factors. Journal of Plant Ecology, 1(2), 85–93.Google Scholar
Zhang, L., Dawes, W. R., and Walker, G. R. (2001). Response of mean annual evapotranspiration to vegetation changes at catchment scale. Water Resources Research, 37(3), 701–708.Google Scholar
Zhang, Q., Wang, Y. P., Pitman, A. J., and Dai, Y. J. (2011). Limitations of nitrogen and phosphorous on the terrestrial carbon uptake in the 20th century. Geophysical Research Letters, 38(22).Google Scholar
Zhang, T., Barry, R. G., Knowles, K., Ling, F., and Armstrong, R. L. (2003). Distribution of seasonally and perennially frozen ground in the Northern Hemisphere. In Proceedings of the 8th International Conference on Permafrost. AA Balkema, Vol. 2, pp. 1289–1294.
Zheng, D., Wallin, D. O., and Hao, Z. (1997). Rates and patterns of landscape change between 1972 and 1988 in the Changbai Mountain area of China and North Korea. Landscape Ecology, 12, 241–254.Google Scholar
Zhu, Z., Xiong, Z., and Xing, G. (2005). Impacts of population growth and economic development on the nitrogen cycle in Asia. Science in China Series C: Life Sciences, 48(2), 729–737.Google Scholar
Zipperer, W. C. (1993). Deforestation patterns and their effects on patches. Landscape Ecology, 8, 177–184.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.

  • References
  • Christiane Runyan, The Johns Hopkins University, Paolo D'Odorico, University of Virginia
  • Book: Global Deforestation
  • Online publication: 05 April 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316471548.008
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.

  • References
  • Christiane Runyan, The Johns Hopkins University, Paolo D'Odorico, University of Virginia
  • Book: Global Deforestation
  • Online publication: 05 April 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316471548.008
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.

  • References
  • Christiane Runyan, The Johns Hopkins University, Paolo D'Odorico, University of Virginia
  • Book: Global Deforestation
  • Online publication: 05 April 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316471548.008
Available formats
×