Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-19T04:23:56.081Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  02 February 2023

Gordon Bonan
Affiliation:
National Center for Atmospheric Research, Boulder, Colorado
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
Seeing the Forest for the Trees
Forests, Climate Change, and Our Future
, pp. 256 - 307
Publisher: Cambridge University Press
Print publication year: 2023

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

Abatzoglou, J. T., and Williams, A. P. (2016). Impact of anthropogenic climate change on wildfire across western US forests. Proceedings of the National Academy of Sciences USA, 113, 1177011775.Google Scholar
Abbe, C. (1889). Is our climate changing? The Forum, 6(February), 678688.Google Scholar
Abbe, C. (1891a). Cloud observations at sea. American Meteorological Journal, 8(6), 250264.Google Scholar
Abbe, C. (1891b). A plea for terrestrial physics. Proceedings of the American Association for the Advancement of Science, 39, 6579.Google Scholar
Abbe, C. (1893). Determination of the true amount of precipitation and its bearing on theories of forest influences. In Forest Influences (US Department of Agriculture, Forestry Division Bulletin Number 7), edited by Fernow, B. E.. Washington, DC: Government Printing Office, pp. 175186.Google Scholar
Abbe, C. (1894a). The relation of forests to climate and health. Proceedings of the American Forestry Association, 10, 4557.Google Scholar
Abbe, C. (1894b). Schools of meteorology. Nature, 50, 576577.Google Scholar
Abbe, C. (1895). Meteorology in the university. Science, 2(48), 709714.Google Scholar
Abbe, C. (1899). The rain gage and the wind. Monthly Weather Review, 27, 464468.Google Scholar
Abbe, C. (1905). A First Report on the Relations between Climates and Crops, US Department of Agriculture, Weather Bureau Bulletin Number 36. Washington, DC: Government Printing Office.Google Scholar
Abrams, M. D. (2001). Eastern white pine versatility in the presettlement forest. BioScience, 51, 967979.Google Scholar
Abu-Izzeddin, F. (2013). Memoirs of a Cedar: A History of Deforestation; A Future of Conservation. Lebanon: Shouf Biosphere Reserve.Google Scholar
Adams, H. D., Zeppel, M. J. B., Anderegg, W. R. L., et al. (2017). A multi-species synthesis of physiological mechanisms in drought-induced tree mortality. Nature Ecology and Evolution, 1, 12851291.Google Scholar
Albert, B., Chandès, H., and Gaudefroy, I. (2019). Trees. New York: Thames & Hudson.Google Scholar
Albion, R. G. (1926). Forests and Sea Power: The Timber Problem of the Royal Navy, 1652–1862. Cambridge, MA: Harvard University Press.Google Scholar
Alkama, R., and Cescatti, A. (2016). Biophysical climate impacts of recent changes in global forest cover. Science, 351, 600604.Google Scholar
Allan, R. P., Arias, P. A., Berger, S., et al. (2021). Summary for Policymakers. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, edited by Masson-Delmotte, V., Zhai, P., Pirani, A., et al. Cambridge, UK: Cambridge University Press, pp. 3–32.Google Scholar
Allen, C. D., Macalady, A. K., Chenchouni, H., et al. (2010). A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. Forest Ecology and Management, 259, 660684.Google Scholar
Allen, J. R. M., Huntley, B., and Watts, W. A. (1996). The vegetation and climate of northwest Iberia over the last 14 000 yr. Journal of Quaternary Science, 11, 125147.Google Scholar
Alter, R. E., Douglas, H. C., Winter, J. M., and Eltahir, E. A. B. (2018). Twentieth century regional climate change during the summer in the central United States attributed to agricultural intensification. Geophysical Research Letters, 45, 15861594.Google Scholar
Amiro, B. D., Orchansky, A. L., Barr, A. G., et al. (2006). The effect of post-fire stand age on the boreal forest energy balance. Agricultural and Forest Meteorology, 140, 4150.Google Scholar
Anderegg, W. R. L., Trugman, A. T., Badgley, G., et al. (2020). Climate-driven risks to the climate mitigation potential of forests. Science, 368, eaaz7005, DOI: https://doi.org/10.1126/science.aaz7005.Google Scholar
Anderegg, W. R. L., Trugman, A. T., Bowling, D. R., Salvucci, G., and Tuttle, S. E. (2019). Plant functional traits and climate influence drought intensification and land–atmosphere feedbacks. Proceedings of the National Academy of Sciences USA, 116, 1407114076.Google Scholar
Anders, J. M. (1878a). On the transpiration of plants. American Naturalist, 12, 160171.Google Scholar
Anders, J. M. (1878b). The beneficial influence of plants. American Naturalist, 12, 793807.Google Scholar
Anders, J. M. (1882). Forests: Their influence upon climate and rainfall. American Naturalist, 16, 1930.Google Scholar
Anders, J. M. (1887). House-Plants as Sanitary Agents; or, the Relation of Growing Vegetation to Health and Disease. Philadelphia: J. B. Lippincott.Google Scholar
Anderson-Teixeira, K. J., Herrmann, V., Morgan, R. B., et al. (2021). Carbon cycling in mature and regrowth forests globally. Environmental Research Letters, 16, 053009, DOI: https://doi.org/10.1088/1748-9326/abed01.Google Scholar
Anderson-Teixeira, K. J., Snyder, P. K., Twine, T. E., et al. (2012). Climate-regulation services of natural and agricultural ecoregions of the Americas. Nature Climate Change, 2, 177181.Google Scholar
Andreae, M. O., Rosenfeld, D., Artaxo, P., et al. (2004). Smoking rain clouds over the Amazon. Science, 303, 13371342.Google Scholar
Andréassian, V. (2004). Waters and forests: From historical controversy to scientific debate. Journal of Hydrology, 291, 127.Google Scholar
Anthes, R. A. (1984). Enhancement of convective precipitation by mesoscale variations in vegetative covering in semiarid regions. Journal of Climate and Applied Meteorology, 23, 541554.Google Scholar
Arago, F. (1859). De l’influence du déboisement sur les climats. In Oeuvres complètes de François Arago, vol. 12, edited by Barral, J.-A.. Paris: Gide, pp. 432443.Google Scholar
Arias, P. A., Bellouin, N., Coppola, E., et al. (2021). Technical Summary. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, edited by Masson-Delmotte, V., Zhai, P., Pirani, A., et al. Cambridge, UK: Cambridge University Press, pp. 33–144.Google Scholar
Arora, V. K., and Montenegro, A. (2011). Small temperature benefits provided by realistic afforestation efforts. Nature Geoscience, 4, 514518.Google Scholar
Arora, V. K., Katavouta, A., Williams, R. G., et al. (2020). Carbon–concentration and carbon–climate feedbacks in CMIP6 models and their comparison to CMIP5 models. Biogeosciences, 17, 41734222.Google Scholar
Arrington, L. J. (1997). The Price of Prejudice: The Japanese-American Relocation Center in Utah during World War II, 2nd ed. Logan: Utah State University.Google Scholar
Artaxo, P., Hansson, H.-C., Andreae, M. O., et al. (2022). Tropical and boreal forest–atmosphere interactions: A review. Tellus B, 74, 24163.Google Scholar
Artaxo, P., Rizzo, L. V., Brito, J. F., et al. (2013). Atmospheric aerosols in Amazonia and land use change: From natural biogenic to biomass burning conditions. Faraday Discussions, 165, 203235.Google Scholar
Astrup, R., Bernier, P. Y., Genet, H., Lutz, D. A., and Bright, R. M. (2018). A sensible climate solution for the boreal forest. Nature Climate Change, 8, 1112.Google Scholar
Aubinet, M., Vesala, T., and Papale, D. (2012). Eddy Covariance: A Practical Guide to Measurement and Data Analysis. Dordrecht: Springer.Google Scholar
Aubrecht, D. M., Helliker, B. R., Goulden, M. L., et al. (2016). Continuous, long-term, high-frequency thermal imaging of vegetation: Uncertainties and recommended best practices. Agricultural and Forest Meteorology, 228–229, 315326.Google Scholar
Aughey, S. (1878). Geology of Nebraska. In Fourth Annual Report of the President and Secretary of the Nebraska State Board of Agriculture. Lincoln: Journal Company, pp. 6785.Google Scholar
Aughey, S. (1880). Sketches of the Physical Geography and Geology of Nebraska. Omaha: Daily Republican Book and Job Office.Google Scholar
Badger, A. M., and Dirmeyer, P. A. (2016). Remote tropical and sub‑tropical responses to Amazon deforestation. Climate Dynamics, 46, 30573066.Google Scholar
Bailey, N. (1736). Dictionarium Britannicum: Or a More Compleat Universal Etymological English Dictionary Than Any Extant, 2nd ed. London: T. Cox.Google Scholar
Baker, J. C. A., and Spracklen, D. V. (2019). Climate benefits of intact Amazon forests and the biophysical consequences of disturbance. Frontiers in Forests and Global Change, 2, 47, DOI: https://doi.org/10.3389/ffgc.2019.00047.Google Scholar
Bala, G., Caldeira, K., Wickett, M., et al. (2007). Combined climate and carbon-cycle effects of large-scale deforestation. Proceedings of the National Academy of Sciences USA, 104, 65506555.Google Scholar
Baldocchi, D. (2014). Measuring fluxes of trace gases and energy between ecosystems and the atmosphere – the state and future of the eddy covariance method. Global Change Biology, 20, 36003609.Google Scholar
Baldocchi, D., and Penuelas, J. (2019). The physics and ecology of mining carbon dioxide from the atmosphere by ecosystems. Global Change Biology, 25, 11911197.Google Scholar
Baldocchi, D., Kelliher, F. M., Black, T. A., and Jarvis, P. (2000). Climate and vegetation controls on boreal zone energy exchange. Global Change Biology, 6(s1), 6983.Google Scholar
Balfour, E. (1849). Notes on the influence exercised by trees in inducing rain and preserving moisture. Madras Journal of Literature and Science, 15(36), 402448.Google Scholar
Bamba, A., Diallo, I., Touré, N. E., et al. (2019). Effect of the African greenbelt position on West African summer climate: A regional climate modeling study. Theoretical and Applied Climatology, 137, 309322.Google Scholar
Banerjee, T., De Roo, F., and Mauder, M. (2017). Explaining the convector effect in canopy turbulence by means of large-eddy simulation. Hydrology and Earth System Sciences, 21, 29873000.Google Scholar
Banerjee, T., Brugger, P., De Roo, F., et al. (2018). Turbulent transport of energy across a forest and a semiarid shrubland. Atmospheric Chemistry and Physics, 18, 1002510038.Google Scholar
Bar-On, Y. M., Phillips, R., and Milo, R. (2018). The biomass distribution on Earth. Proceedings of the National Academy of Sciences USA, 115, 65066511.Google Scholar
Bartlein, P. J., Harrison, S. P., Brewer, S., et al. (2011). Pollen-based continental climate reconstructions at 6 and 21 ka: A global synthesis. Climate Dynamics, 37, 775802.Google Scholar
Barton, A. (2017). The Shakespearean Forest. Cambridge, UK: Cambridge University Press.Google Scholar
Barton, B. S. (1807). A Discourse on Some of the Principal Desiderata in Natural History, and on the Best Means of Promoting the Study of this Science, in the United-States. Philadelphia: Denham & Town.Google Scholar
Barton, C. V. M., Ellsworth, D. S., Medlyn, B. E., et al. (2010). Whole-tree chambers for elevated atmospheric CO2 experimentation and tree scale flux measurements in south-eastern Australia: The Hawkesbury Forest Experiment. Agricultural and Forest Meteorology, 150, 941951.Google Scholar
Barton, G. A. (2002). Empire Forestry and the Origins of Environmentalism. Cambridge, UK: Cambridge University Press.Google Scholar
Bastin, J.-F., Finegold, Y., Garcia, C., et al. (2019). The global tree restoration potential. Science, 365, 7679.Google Scholar
Bates, C. G., and Henry, A. J. (1928). Forest and stream-flow experiment at Wagon Wheel Gap, Colo.: Final report, on completion of the second phase of the experiment. Monthly Weather Review, Supplement Number 30, 179.Google Scholar
Battle, M., Bender, M. L., Tans, P. P., et al. (2000). Global carbon sinks and their variability inferred from atmospheric O2 and δ13C. Science, 287, 24672470.Google Scholar
Baudrillart, J.-J. (1823). Traité général des eaux et forêts, chasses et pêches. 2nd partie. Dictionnaire général, raisonné et historique des eaux et forêts, vol. 1. Paris: Huzard; Artus Bertrand; Warée oncle.Google Scholar
Baum, L. F. (1900). The Wonderful Wizard of Oz. Chicago: George M. Hill.Google Scholar
Beard, J. S. (1949). The Natural Vegetation of the Windward & Leeward Islands. Oxford: Clarendon Press.Google Scholar
Beatson, A. (1816). Tracts Relative to the Island of St. Helena; Written During a Residence of Five Years. London: W. Bulmer.Google Scholar
Beattie, J. (2003). Environmental anxiety in New Zealand, 1840–1941: Climate change, soil erosion, sand drift, flooding and forest conservation. Environment and History, 9, 379392.Google Scholar
Beattie, J. (2009). Climate change, forest conservation and science: A case study of New Zealand, 1860s–1920. History of Meteorology, 5, 118.Google Scholar
Beattie, J., and Star, P. (2010). Global influences and local environments: Forestry and forest conservation in New Zealand, 1850s–1925. British Scholar, 3, 191218.Google Scholar
Becquerel, A.-C. (1853). Des climats et de l’influence qu’exercent les sols boisés et non boisés. Paris: Firmin Didot Frères.Google Scholar
Becquerel, A.-C. (1860). Recherches sur la température des végétaux et de l’air et sur celle du sol a diverses profondeurs. Paris: Firmin Didot Frères, Fils et Cie.Google Scholar
Becquerel, A.-C. (1865). Mémoire sur les forêts et leur influence climatérique. Paris: Firmin Didot Frères, Fils, et Cie.Google Scholar
Becquerel, A.-C. (1867). Mémoire sur les principales causes qui influent sur les pluies. Comptes rendus hebdomadaires des séances de l’Académie des sciences, 64, 837843.Google Scholar
Becquerel, A.-C. (1872). Forests and their climatic influence. In Annual Report of the Board of Regents of the Smithsonian Institution, Showing the Operations, Expenditures, and Condition of the Institution for the Year 1869. Washington, DC: Government Printing Office, pp. 394416.Google Scholar
Becquerel, A.-C. (1878). Memoir upon forests, and their climatic influence. In Report upon Forestry, edited by Hough, F. B.. Washington, DC: Government Printing Office, pp. 310333.Google Scholar
Becquerel, A.-C., and Becquerel, E. (1847). Éléments de physique terrestre et de météorologie. Paris: Firmin Didot Frères.Google Scholar
Becquerel, A.-C., and Becquerel, E. (1866). Des pluies dans les lieux boisés et non boisés. Comptes rendus hebdomadaires des séances de l’Académie des sciences, 62, 855858.Google Scholar
Becquerel, A.-C., and Becquerel, E. (1867). Extrait d’un mémoire sur les températures de l’air et les quantités d’eau tombées hors du bois et sous bois. Comptes rendus hebdomadaires des séances de l’Académie des sciences, 64, 1619.Google Scholar
Becquerel, A.-C., and Becquerel, E. (1869a). Mémoire sur la température de l’air sous bois et hors des bois. Comptes rendus hebdomadaires des séances de l’Académie des sciences, 68, 677682.Google Scholar
Becquerel, A.-C., and Becquerel, E. (1869b). Des quantités d’eau tombées près et loin des bois. Comptes rendus hebdomadaires des séances de l’Académie des sciences, 68, 789793.Google Scholar
Beerling, D. (2007). The Emerald Planet: How Plants Changed Earth’s History. Oxford: Oxford University Press.Google Scholar
Beerling, D. J., and Osborne, C. P. (2002). Physiological ecology of Mesozoic polar forests in a high CO2 environment. Annals of Botany, 89, 329339.Google Scholar
Bennett, B. M., and Barton, G. A. (2018). The enduring link between forest cover and rainfall: A historical perspective on science and policy discussions. Forest Ecosystems, 5, 5, DOI: https://doi.org/10.1186/s40663-017-0124-9.Google Scholar
Berg, A., Findell, K., Lintner, B., et al. (2016). Land-atmosphere feedbacks amplify aridity increase over land under global warming. Nature Climate Change, 6, 869874.Google Scholar
Berger, Dr. (1865). Wald und Witterung. Annalen der Physik und Chemie, 124, 528568.Google Scholar
Berkeley, G. (1734). A Treatise Concerning the Principles of Human Knowledge… to Which Are Added Three Dialogues between Hylas and Philonous, in Opposition to Scepticks and Atheists. London: Jacob Tonson.Google Scholar
Bernardin de Saint-Pierre, J.-H. (1784). Etudes de la nature, 3 vols. Paris: Pierre-François Didot.Google Scholar
Berner, L. T., Law, B. E., Meddens, A. J. H., and Hicke, J. A. (2017). Tree mortality from fires, bark beetles, and timber harvest during a hot and dry decade in the western United States (2003–2012). Environmental Research Letters, 12, 065005, DOI: https://doi.org/10.1088/1748-9326/aa6f94.Google Scholar
Berner, R. A. (2012). Jacques-Joseph Ébelmen, the founder of earth system science. Comptes Rendus Geoscience, 344, 544548.Google Scholar
Berry, C. M. (2019). Palaeobotany: The rise of the Earth’s early forests. Current Biology, 29, R792–R794.Google Scholar
Betts, R. A. (2000). Offset of the potential carbon sink from boreal forestation by decreases in surface albedo. Nature, 408, 187190.Google Scholar
Betts, R. A., Boucher, O., Collins, M., et al. (2007). Projected increase in continental runoff due to plant responses to increasing carbon dioxide. Nature, 448, 10371041.Google Scholar
Beven, K. (2006). A manifesto for the equifinality thesis. Journal of Hydrology, 320, 1836.Google Scholar
Biard, P. (1616). Relation de la Nouvelle France, de ses terres, naturel du païs, & de ses habitans. Lyon: Louys Muguet.Google Scholar
Biederman, J. A., Somor, A. J., Harpold, A. A., et al. (2015). Recent tree die-off has little effect on streamflow in contrast to expected increases from historical studies. Water Resources Research, 51, 97759789.Google Scholar
Billings, W. D. (1938). The structure and development of old field shortleaf pine stands and certain associated physical properties of the soil. Ecological Monographs, 8, 437499.Google Scholar
Black, J. F. (1963). Weather control: Use of asphalt coatings to tap solar energy. Science, 139, 226227.Google Scholar
Black, J. F., and Tarmy, B. L. (1963). The use of asphalt coatings to increase rainfall. Journal of Applied Meteorology, 2, 557564.Google Scholar
Blanford, H. F. (1886). Influence of forests on rainfall. Indian Meteorological Memoirs, 3, 135145.Google Scholar
Blanford, H. F. (1887). On the influence of Indian forests on the rainfall. Journal of the Asiatic Society of Bengal, Part II, 56, 115.Google Scholar
Blanford, H. F. (1888). Influence of forests on rainfall. The Indian Forester, 14, 3447.Google Scholar
Blodget, L. (1857). Climatology of the United States, and of the Temperate Latitudes of the North American Continent. Philadelphia: J. B. Lippincott.Google Scholar
Blodget, L. (1874). Forest cultivation on the plains: The climate and cultivable capacity of the plains considered in regard to the ameliorations possible through greater protection by forests. In Report of the Commissioner of Agriculture for the Year 1872. Washington, DC: Government Printing Office, pp. 316332.Google Scholar
Blome, R. (1672). A Description of the Island of Jamaica; with the Other Isles and Territories in America, to which the English are Related. London: T. Milbourn.Google Scholar
Blyth, E. M., Arora, V. K., Clark, D. B., et al. (2021). Advances in land surface modelling. Current Climate Change Reports, 7, 4571.Google Scholar
Boisier, J. P., de Noblet-Ducoudré, N., Pitman, A. J., et al. (2012). Attributing the impacts of land-cover changes in temperate regions on surface temperature and heat fluxes to specific causes: Results from the first LUCID set of simulations. Journal of Geophysical Research, 117, D12116, DOI: https://doi.org/10.1029/2011JD017106.Google Scholar
Bolin, B. (1977). Changes of land biota and their importance for the carbon cycle. Science, 196, 613615.Google Scholar
Bonan, G. B. (1997). Effects of land use on the climate of the United States. Climatic Change, 37, 449486.Google Scholar
Bonan, G. B. (1999). Frost followed the plow: Impacts of deforestation on the climate of the United States. Ecological Applications, 9, 13051315.Google Scholar
Bonan, G. B. (2008). Forests and climate change: Forcings, feedbacks, and the climate benefits of forests. Science, 320, 14441449.Google Scholar
Bonan, G. B. (2016a). Ecological Climatology: Concepts and Applications, 3rd ed. Cambridge, UK: Cambridge University Press.Google Scholar
Bonan, G. B. (2016b). Forests, climate, and public policy: A 500-year interdisciplinary odyssey. Annual Review of Ecology, Evolution, and Systematics, 47, 97121.Google Scholar
Bonan, G. B. (2019). Climate Change and Terrestrial Ecosystem Modeling. Cambridge, UK: Cambridge University Press.Google Scholar
Bonan, G. B., and Doney, S. C. (2018). Climate, ecosystems, and planetary futures: The challenge to predict life in Earth system models. Science, 359, eaam8328, DOI: https://doi.org/10.1126/science.aam8328.Google Scholar
Bonan, G. B., and Shugart, H. H. (1989). Environmental factors and ecological processes in boreal forests. Annual Review of Ecology and Systematics, 20, 128.Google Scholar
Bonan, G. B., Davis, K. J., Baldocchi, D., Fitzjarrald, D., and Neumann, H. (1997). Comparison of the NCAR LSM1 land surface model with BOREAS aspen and jack pine tower fluxes. Journal of Geophysical Research, 102D, 2906529075.Google Scholar
Bonan, G. B., Pollard, D., and Thompson, S. L. (1992). Effects of boreal forest vegetation on global climate. Nature, 359, 716718.Google Scholar
Bonneuil, C., and Fressoz, J.-B. (2016). The Shock of the Anthropocene: The Earth, History and Us, translated by D. Fernbach. London: Verso.Google Scholar
Bormann, F. H., and Likens, G. E. (1979). Pattern and Process in a Forested Ecosystem. New York: Springer-Verlag.Google Scholar
Bosch, J. M., and Hewlett, J. D. (1982). A review of catchment experiments to determine the effect of vegetation changes on water yield and evapotranspiration. Journal of Hydrology, 55, 323.Google Scholar
Bosson, M[onsieur] (1825). Second mémoire en réponse a cette question: Quels sont les changemens que peut occasioner le déboisement de forêts considérables sur les contrées et communes adjacentes…. Brussels: P. J. de Mat.Google Scholar
Boulton, C. A., Lenton, T. M., and Boers, N. (2022). Pronounced loss of Amazon rainforest resilience since the early 2000s. Nature Climate Change, 12, 271278.Google Scholar
Boussingault, J.-B. (1833). Mémoire sur la profondeur à laquelle se trouve la couche de température invariable entre les tropiques. Détermination de la température moyenne de la zône torride au niveau de la mer. Observations sur le décroissement de la chaleur dans les Cordilières. Annales de chimie et de physique, 53, 225247.Google Scholar
Boussingault, J.-B. (1834). Recherches sur la composition de l’atmosphère. Premier Mémoire. Sur la possibilité de constater l’existence des miasmes. – Sur la présence d’un principe hydrogéné dans l’air. Annales de chimie et de physique, 57, 148182.Google Scholar
Boussingault, J.-B. (1837). Mémoire sur l’influence des défrichemens dans la diminution des cours d’eau. Annales de chimie et de physique, 64, 113141.Google Scholar
Boussingault, J.-B. (1838). Memoir concerning the effect which the clearing of land has in diminishing the quantity of water in the streams of the district. Edinburgh New Philosophical Journal, 24, 85106.Google Scholar
Boussingault, J.-B. (1845). Rural Economy, in its Relations with Chemistry, Physics, and Meteorology, translated by G. Law. London: H. Bailliere.Google Scholar
Boussingault, J.-B. (1851). Économie rurale considérée das ses rapports avec la chimie, la physique et la météorologie, 2nd ed., 2 vols. Paris: Béchet Jeune.Google Scholar
Bowditch, E., Santopuoli, G., Binder, F., et al. (2020). What is Climate-Smart Forestry? A definition from a multinational collaborative process focused on mountain regions of Europe. Ecosystem Services, 43, 101113, DOI: https://doi.org/10.1016/j.ecoser.2020.101113.Google Scholar
Boyce, C. K., and Lee, J.-E. (2017). Plant evolution and climate over geological timescales. Annual Review of Earth and Planetary Sciences, 45, 6187.Google Scholar
Boyle, R. (1671). Cosmicall suspitions (subjoyned as an appendix to the discourse of the cosmicall qualities of things). In Tracts. Oxford: W. H. for R. Davis, pp. 128.Google Scholar
Boysen, L. R., Brovkin, V., Arora, V. K., et al. (2014). Global and regional effects of land-use change on climate in 21st century simulations with interactive carbon cycle. Earth System Dynamics, 5, 309319.Google Scholar
Boysen, L. R., Brovkin, V., Pongratz, J., et al. (2020). Global climate response to idealized deforestation in CMIP6 models. Biogeosciences, 17, 56155638.Google Scholar
Braconnot, P., Joussaume, S., de Noblet, N., and Ramstein, G. (2000). Mid-Holocene and Last Glacial Maximum African monsoon changes as simulated within the Paleoclimate Modelling Intercomparison Project. Global and Planetary Change, 26, 5166.Google Scholar
Branch, O., and Wulfmeyer, V. (2019). Deliberate enhancement of rainfall using desert plantations. Proceedings of the National Academy of Sciences USA, 116, 1884118847.Google Scholar
Brandis, D. (1887). Regen und Wald in Indien. Meteorologische Zeitschrift, 4, 369376.Google Scholar
Brandis, D. (1888). The influence of forests on rainfall. The Indian Forester, 14, 1020.Google Scholar
Brando, P. M., Paolucci, L., Ummenhofer, C. C., et al. (2019). Droughts, wildfires, and forest carbon cycling: A pantropical synthesis. Annual Review of Earth and Planetary Sciences, 47, 555581.Google Scholar
Breil, M., Rechid, D., Davin, E. L., et al. (2020). The opposing effects of reforestation and afforestation on the diurnal temperature cycle at the surface and in the lowest atmospheric model level in the European summer. Journal of Climate, 33, 91599179.Google Scholar
Brewer, S., Giesecke, T., Davis, B. A. S., et al. (2017). Late-glacial and Holocene European pollen data. Journal of Maps, 13, 921928.Google Scholar
Brewster, D. (1830). The Edinburgh Encyclopædia, 4th ed., 18 vols. Edinburgh: William Blackwood; John Waugh; and others.Google Scholar
Bright, B. C., Hicke, J. A., and Meddens, A. J. H. (2013). Effects of bark beetle-caused tree mortality on biogeochemical and biogeophysical MODIS products. Journal of Geophysical Research: Biogeosciences, 118, 974982.Google Scholar
Bright, R. M., Allen, M., Antón-Fernández, C., et al. (2020). Evaluating the terrestrial carbon dioxide removal potential of improved forest management and accelerated forest conversion in Norway. Global Change Biology, 26, 50875105.Google Scholar
Bright, R. M., Antón-Fernández, C., Astrup, R., et al. (2014). Climate change implications of shifting forest management strategy in a boreal forest ecosystem of Norway. Global Change Biology, 20, 607621.Google Scholar
Bright, R. M., Davin, E., O’Halloran, T., et al. (2017). Local temperature response to land cover and management change driven by non-radiative processes. Nature Climate Change, 7, 296302.Google Scholar
Bright, R. M., and Lund, M. T. (2021). CO2-equivalence metrics for surface albedo change based on the radiative forcing concept: A critical review. Atmospheric Chemistry and Physics, 21, 98879907.Google Scholar
Bright, R. M., Zhao, K., Jackson, R. B., and Cherubini, F. (2015). Quantifying surface albedo and other direct biogeophysical climate forcings of forestry activities. Global Change Biology, 21, 32463266.Google Scholar
Brincken, J. (1828). Mémoire descriptif sur la forêt impériale de Białowieża, en Lithuanie. Varsovie: N. Glücksberg.Google Scholar
Broecker, W. S. (1970). Man’s oxygen reserves. Science, 168, 15371538.Google Scholar
Brooks, C. E. P. (1928). The influence of forests on rainfall and run-off. Quarterly Journal of the Royal Meteorological Society, 54, 117.Google Scholar
Brouard, N. R. (1963). A History of Woods and Forests in Mauritius. Port Louis, Mauritius: J. E. Félix.Google Scholar
Brovkin, V., Boysen, L., Arora, V. K., et al. (2013). Effect of anthropogenic land-use and land-cover changes on climate and land carbon storage in CMIP5 projections for the twenty-first century. Journal of Climate, 26, 68596881.Google Scholar
Brovkin, V., Claussen, M., Petoukhov, V., and Ganopolski, A. (1998). On the stability of the atmosphere-vegetation system in the Sahara/Sahel region. Journal of Geophysical Research, 103D, 3161331624.Google Scholar
Brovkin, V., Raddatz, T., Reick, C. H., Claussen, M., and Gayler, V. (2009). Global biogeophysical interactions between forest and climate. Geophysical Research Letters, 36, L07405, DOI: https://doi.org/10.1029/2009GL037543.Google Scholar
Brovkin, V., Sitch, S., von Bloh, W., et al. (2004). Role of land cover changes for atmospheric CO2 increase and climate change during the last 150 years. Global Change Biology, 10, 12531266.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, 2861.Google Scholar
Brown, H. T., and Escombe, F. (1905). Researches on some of the physiological processes of green leaves, with special reference to the interchange of energy between the leaf and its surroundings. Proceedings of the Royal Society of London B, 76, 29111.Google Scholar
Brown, H. T., and Wilson, W. E. (1905). On the thermal emissivity of a green leaf in still and moving air. Proceedings of the Royal Society London B, 76, 122137.Google Scholar
Brown, J. C. (1875). Hydrology of South Africa; or Details of the Former Hydrographic Condition of the Cape of Good Hope, and of Causes of its Present Aridity, with Suggestions of Appropriate Remedies for this Aridity. London: Henry S. King.Google Scholar
Brown, J. C. (1876). Reboisement in France: Or, Records of the Replanting of the Alps, the Cevennes, and the Pyrenees with Trees, Herbage, and Bush. London: Henry S. King.Google Scholar
Brown, J. C. (1877a). Forests and Moisture; or Effects of Forests on Humidity of Climate. Edinburgh: Oliver & Boyd.Google Scholar
Brown, J. C. (1877b). Water Supply of South Africa and Facilities for the Storage of It. Edinburgh: Oliver & Boyd.Google Scholar
Brown, J. C. (1883). French Forest Ordinance of 1669; with Historical Sketch of Previous Treatment of Forests in France. Edinburgh: Oliver & Boyd.Google Scholar
Brown, J. C. (1887). Management of Crown Forests at the Cape of Good Hope under the Old Regime and under the New. Edinburgh: Oliver & Boyd.Google Scholar
Brown, R. H. (1951). The seaboard climate in the view of 1800. Annals of the Association of American Geographers, 41, 217232.Google Scholar
Brückner, E. (1890). Klimaschwankungen seit 1700 nebst Bemerkungen über die Klimaschwankungen der Diluvialzeit. Penck’s Geographische Abhandlungen, 4, 153484.Google Scholar
Brugger, P., De Roo, F., Kröniger, K., et al. (2019). Contrasting turbulent transport regimes explain cooling effect in a semi-arid forest compared to surrounding shrubland. Agricultural and Forest Meteorology, 269270, 1927.Google Scholar
Bruijnzeel, L. A., Mulligan, M., and Scatena, F. N. (2011). Hydrometeorology of tropical montane cloud forests: Emerging patterns. Hydrological Processes, 25, 465498.Google Scholar
Buffon, G.-L. Leclerc, comte de (1778). Histoire naturelle, générale et particulière. Supplément, tome cinquième. Des époques de la nature. Paris: Imprimerie royale.Google Scholar
Burakowski, E. A., Ollinger, S. V., Bonan, G. B., et al. (2016). Evaluating the climate effects of reforestation in New England using a Weather Research and Forecasting (WRF) Model multiphysics ensemble. Journal of Climate, 29, 51415156.Google Scholar
Burakowski, E., Tawfik, A., Ouimette, A., et al. (2018). The role of surface roughness, albedo, and Bowen ratio on ecosystem energy balance in the Eastern United States. Agricultural and Forest Meteorology, 249, 367376.Google Scholar
Burgess, G. S. (2004). The History of the Norman People: Wace’s “Roman de Rou,” translated by G. S. Burgess, notes by Burgess, G. S. and van Houts, E.. Woodbridge, UK: Boydell Press.Google Scholar
Butter, D. (1839). Outlines of the Topography and Statistics of the Southern Districts of Oud’h, and of the Cantonment of Sultanpur-Oud’h. Calcutta: G. H. Huttmann.Google Scholar
Caldeira, K., Bala, G., and Cao, L. (2013). The science of geoengineering. Annual Review of Earth and Planetary Sciences, 41, 231256.Google Scholar
Calder, I. R. (2002). Forests and hydrological services: Reconciling public and science perceptions. Land Use and Water Resources Research, 2, 112.Google Scholar
Calder, I. R. (2007). Forests and water: Ensuring forest benefits outweigh water costs. Forest Ecology and Management, 251, 110120.Google Scholar
Calder, I., Amezaga, J., Aylward, B., et al. (2004). Forest and water policies. The need to reconcile public and science perceptions. Geologica Acta, 2, 157166.Google Scholar
Campbell Walker, I. (1876a). State forestry: Its aim and object. Transactions and Proceedings of the New Zealand Institute, 9, 187203.Google Scholar
Campbell Walker, I. (1876b). The climatic and financial aspect of forest conservancy as applicable to New Zealand. Transactions and Proceedings of the New Zealand Institute, 9, Appendix, pp. xxvii–xlix.Google Scholar
Campbell Walker, I. (1877). Report of the conservator of state forests. In Appendix to the Journals of the House of Representatives of New Zealand, vol. 1. Wellington: George Didsbury, pp. C3:159.Google Scholar
Canadell, J. G., Monteiro, P. M. S., Costa, M. H., et al. (2021). Global carbon and other biogeochemical cycles and feedbacks. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, edited by Masson-Delmotte, V., Zhai, P., Pirani, A., et al. Cambridge, UK: Cambridge University Press, pp. 673–816.Google Scholar
Capron, H. (1869). Report of the Commissioner of Agriculture for the Year 1868. Washington, DC: Government Printing Office.Google Scholar
Capron, H. (1870). Report of the Commissioner of Agriculture for the Year 1869. Washington, DC: Government Printing Office.Google Scholar
Capron, H. (1871). Report of the Commissioner of Agriculture for the Year 1870. Washington, DC: Government Printing Office.Google Scholar
Carey, F. (2012). The Tree: Meaning and Myth. Burlington, VT: Lund Humphries.Google Scholar
Carlson, D. W., and Groot, A. (1997). Microclimate of clear-cut, forest interior, and small openings in trembling aspen forest. Agricultural and Forest Meteorology, 87, 313329.Google Scholar
Carlyle-Moses, D. E., and Gash, J. H. C. (2011). Rainfall interception loss by forest canopies. In Forest Hydrology and Biogeochemistry: Synthesis of Past Research and Future Directions, edited by Levia, D. F., Carlyle-Moses, D., and Tanaka, T.. Dordrecht: Springer, pp. 407423.Google Scholar
Cerasoli, S., Yin, J., and Porporato, A. (2021). Cloud cooling effects of afforestation and reforestation at midlatitudes. Proceedings of the National Academy of Sciences USA, 118, e2026241118, DOI: https://doi.org/10.1073/pnas.2026241118.Google Scholar
Cézanne, E. (1870–72). Étude sur les torrents des hautes-alpes par Alexandre Surell, 2e édition, avec une suite par Ernest Cézanne, 2 vols. Paris: Dunod.Google Scholar
Chandan, D., and Peltier, W. R. (2020). African Humid Period precipitation sustained by robust vegetation, soil, and lake feedbacks. Geophysical Research Letters, 47, e2020GL088728, DOI: https://doi.org/10.1029/2020GL088728.Google Scholar
Chapin, F. S., III, Oswood, M. W., Van Cleve, K., Viereck, L. A., and Verbyla, D. L. (2006). Alaska’s Changing Boreal Forest. Oxford: Oxford University Press.Google Scholar
Charney, J. G. (1975). Dynamics of deserts and drought in the Sahel. Quarterly Journal of the Royal Meteorological Society, 101, 193202.Google Scholar
Charney, J., Stone, P. H., and Quirk, W. J. (1975). Drought in the Sahara: A biogeophysical feedback mechanism. Science, 187, 434435.Google Scholar
Charney, J., Quirk, W. J., Chow, S.-H., and Kornfield, J. (1977). A comparative study of the effects of albedo change on drought in semi-arid regions. Journal of the Atmospheric Sciences, 34, 13661385.Google Scholar
Chastellux, F.-J., marquis de (1963). Travels in North America in the Years 1780, 1781 and 1782, translated with introduction and notes by H. C. Rice, Jr., 2 vols. Chapel Hill: University of North Carolina Press.Google Scholar
Chekhov, A. (2011). Five Plays. Antov Chekhov, translated by M. Brodskaya, introduction by Wolff, T.. Stanford: Stanford University Press.Google Scholar
Chen, C., Li, D., Li, Y., et al. (2020). Biophysical impacts of Earth greening largely controlled by aerodynamic resistance. Science Advances, 6, eabb1981, DOI: https://doi.org/10.1126/sciadv.abb1981.Google Scholar
Chen, C., Park, T., Wang, X., et al. (2019). China and India lead in greening of the world through land-use management. Nature Sustainability, 2, 122129.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, 219237.Google Scholar
Chen, L., and Dirmeyer, P. A. (2017). Impacts of land-use/land-cover change on afternoon precipitation over North America. Journal of Climate, 30, 21212140.Google Scholar
Chen, L., and Dirmeyer, P. A. (2019). Differing responses of the diurnal cycle of land surface and air temperatures to deforestation. Journal of Climate, 32, 70677079.Google Scholar
Chen, L., and Dirmeyer, P. A. (2020). Reconciling the disagreement between observed and simulated temperature responses to deforestation. Nature Communications, 11, 202, DOI: https://doi.org/10.1038/s41467-019-14017-0.Google Scholar
Chen, L., Dirmeyer, P. A., Guo, Z., and Schultz, N. M. (2018). Pairing FLUXNET sites to validate model representations of land‐use/land‐cover change. Hydrology and Earth System Sciences, 22, 111125.Google Scholar
Cheng, L., Zhang, L., Wang, Y.-P., et al. (2017). Recent increases in terrestrial carbon uptake at little cost to the water cycle. Nature Communications, 8, 110, DOI: https://doi.org/10.1038/s41467-017-00114-5.Google Scholar
Chianese, R. L. (2013). Regeneration on Tree Mountain. American Scientist, 101(5), 350351.Google Scholar
Chilukoti, N., and Xue, Y. (2021). An assessment of potential climate impact during 1948–2010 using historical land use land cover change maps. International Journal of Climatology, 41, 295315.Google Scholar
Chinard, G. (1945). The American Philosophical Society and the early history of forestry in America. Proceedings of the American Philosophical Society, 89, 444488.Google Scholar
Chittenden, H. M. (1909). Forests and reservoirs in their relation to stream flow, with particular reference to navigable rivers. Transactions of the American Society of Civil Engineers, 62, 245546.Google Scholar
Christensen, N. L., and Peet, R. K. (1981). Secondary forest succession on the North Carolina Piedmont. In Forest Succession: Concepts and Application, edited by West, D. C., Shugart, H. H., and Botkin, D. B.. New York: Springer-Verlag, pp. 230245.Google Scholar
Churkina, G., Organschi, A., Reyer, C. P. O., et al. (2020). Buildings as a global carbon sink. Nature Sustainability, 3, 269276.Google Scholar
Ciais, P., Tan, J., Wang, X., et al. (2019). Five decades of northern land carbon uptake revealed by the interhemispheric CO2 gradient. Nature, 568, 221225.Google Scholar
Clark, D. B., Xue, Y., Harding, R. J., and Valdes, P. J. (2001). Modeling the impact of land surface degradation on the climate of tropical North Africa. Journal of Climate, 14, 18091822.Google Scholar
Clarke, W. B. (1835). Instances of the effects of forest vegetation on climate. Magazine of Natural History, and Journal of Zoology, Botany, Mineralogy, Geology, and Meteorology, 8, 473482.Google Scholar
Clarke, W. B. (1876). Effects of forest vegetation on climate. Journal and Proceedings of the Royal Society of New South Wales, 10, 179235.Google Scholar
Claussen, M. (2009). Late Quaternary vegetation–climate feedbacks. Climate of the Past, 5, 203216.Google Scholar
Claussen, M., Brovkin, V., and Ganopolski, A. (2001). Biogeophysical versus biogeochemical feedbacks of large-scale land cover change. Geophysical Research Letters, 28, 10111014.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, 99118.Google Scholar
Claussen, M., Fohlmeister, J., Ganopolski, A., and Brovkin, V. (2006). Vegetation dynamics amplifies precessional forcing. Geophysical Research Letters, 33, L09709, DOI: https://doi.org/10.1029/2006GL026111.Google Scholar
Claussen, M., Kubatzki, C., Brovkin, V., et al. (1999). Simulation of an abrupt change in Saharan vegetation in the mid-Holocene. Geophysical Research Letters, 26, 20372040.Google Scholar
Clavé, J. (1862). Études sur l’économie forestière. Paris: Guillaumin.Google Scholar
Clavé, J. (1875). Étude de météorologie forestière. Revue des deux mondes, troisième période, 9(3), 632649.Google Scholar
Clayton, J. (1693). A letter from Mr. John Clayton Rector of Crofton at Wakefield in Yorkshire to the Royal Society, May 12, 1688, giving an account of several observables in Virginia, and in his voyage thither, more particularly concerning the air. Philosophical Transactions, 17(201), 781795.Google Scholar
Cleaveland, P. (1809). Meteorological observations, made at Bowdoin College. Memoirs of the American Academy of Arts and Sciences, 3(1), 119121.Google Scholar
Cleghorn, H., Royle, J. F., Baird Smith, R., and Strachey, R. (1852). Report of the committee appointed by the British Association to consider the probable effects in an economical and physical point of view of the destruction of tropical forests. In Report of the Twenty-First Meeting of the British Association for the Advancement of Science; Held at Ipswich in July 1851. London: John Murray, pp. 78102.Google Scholar
Coates, C., and Degroot, D. (2015). “Les bois engendrent les frimas et les gelées”: compendre le climat en Nouvelle-France. Revue d’histoire de l’Amérique française, 68, 197219.Google Scholar
Coen, D. R. (2018). Climate in Motion: Science, Empire, and the Problem of Scale. Chicago: University of Chicago Press.Google Scholar
Cogbill, C. V., Burk, J., and Motzkin, G. (2002). The forests of presettlement New England, USA: Spatial and compositional patterns based on town proprietor surveys. Journal of Biogeography, 29, 12791304.Google Scholar
Colón, F. (1571). Historie del S. D. Fernando Colombo; nelle quali s’ha particolare, & vera relatione della vita, & de’ fatti dell’Ammiraglio D. Christoforo Colombo, suo padre. Venice: Francesco de’ Franceschi Sanese.Google Scholar
Colón, F. (1959). The Life of the Admiral Christopher Columbus by His Son Ferdinand, translated and annotated by B. Keen. New Brunswick, NJ: Rutgers University Press.Google Scholar
Colyvan, M., and Ginzburg, L. R. (2003). Laws of nature and laws of ecology. Oikos, 101, 649653.Google Scholar
Conant, J. B. (1950). The Overthrow of the Phlogiston Theory: The Chemical Revolution of 1775–1789 (Harvard Case Histories in Experimental Science, Case 2). Cambridge, MA: Harvard University Press.Google Scholar
Cook, B. I., Cook, E. R., Smerdon, J. E., et al. (2016). North American megadroughts in the Common Era: Reconstructions and simulations. WIREs Climate Change, 7, 411432.Google Scholar
Cook, B. I., Miller, R. L., and Seager, R. (2008). Dust and sea surface temperature forcing of the 1930s “Dust Bowl” drought. Geophysical Research Letters, 35, L08710, DOI: https://doi.org/10.1029/2008GL033486.Google Scholar
Cook, B. I., Miller, R. L., and Seager, R. (2009). Amplification of the North American “Dust Bowl” drought through human-induced land degradation. Proceedings of the National Academy of Sciences USA, 106, 49975001.Google Scholar
Cook, B. I., Seager, R., and Miller, R. L. (2011). Atmospheric circulation anomalies during two persistent North American droughts: 1932–1939 and 1948–1957. Climate Dynamics, 36, 23392355Google Scholar
Cooper, E. (1876). Forest Culture and Eucalyptus Trees. San Francisco: Cubery.Google Scholar
Costanza, R., d’Arge, R., de Groot, R., et al. (1997). The value of the world’s ecosystem services and natural capital. Nature, 387, 253260.Google Scholar
Costanza, R., de Groot, R., Braat, L., et al. (2017). Twenty years of ecosystem services: How far have we come and how far do we still need to go? Ecosystem Services, 28 (part A), 116.Google Scholar
Costlow, J. (2003). Imaginations of destruction: The “forest question” in nineteenth-century Russian culture. The Russian Review, 62, 91118.Google Scholar
Cotta, H. (1832). Grundriß der Forstwissenschaft. Dresden & Leipzig: Arnold.Google Scholar
Cowan, I. R. (1977). Stomatal behaviour and environment. Advances in Botanical Research, 4, 117228.Google Scholar
Cowan, I. R., and Farquhar, G. D. (1977). Stomatal function in relation to leaf metabolism and environment. In Integration of Activity in the Higher Plant, edited by Jennings, D. H.. Cambridge, UK: Cambridge University Press, pp. 471505.Google Scholar
Cowan, T., Hegerl, G. C., Schurer, A., et al. (2020). Ocean and land forcing of the record-breaking Dust Bowl heatwaves across central United States. Nature Communications, 11, 2870, DOI: https://doi.org/10.1038/s41467-020-16676-w.Google Scholar
Crane, P. (2013). Ginkgo: The Tree That Time Forgot. New Haven, CT: Yale University Press.Google Scholar
Cronon, W. (1983). Changes in the Land: Indians, Colonists, and the Ecology of New England. New York: Hill & Wang.Google Scholar
Crowther, T. W., Glick, H. B., Covey, K. R., et al. (2015). Mapping tree density at a global scale. Nature, 525, 201205.Google Scholar
Curtis, G. E. (1893). Analysis of the causes of rainfall with special relation to surface conditions. In Forest Influences (US Department of Agriculture, Forestry Division Bulletin Number 7), edited by Fernow, B. E.. Washington, DC: Government Printing Office, pp. 187191.Google Scholar
Curtis, J. T. (1956). The modification of mid-latitude grasslands and forests by man. In Man’s Role in Changing the Face of the Earth, edited by Thomas, W. L., Jr. Chicago: University of Chicago Press, pp. 721736.Google Scholar
Dalhousie, J. A. Ramsay, Broun, marquess (1868). Minute by the most noble the Marquis of Dalhousie, K. G., Governor General of India, dated 20th February, 1851. In Select Papers of the Agri-Horticultural Society of the Punjab, from its Commencement to 1862. Lahore: Lahore Chronicle Press, pp. 15.Google Scholar
Dallmeyer, A., Claussen, M., Lorenz, S. J., and Shanahan, T. (2020). The end of the African humid period as seen by a transient comprehensive Earth system model simulation of the last 8000 years. Climate of the Past, 16, 117140.Google Scholar
Dalzell, N. A. (1863). Observations on the Influence of Forests, and on the General Principles of Management, as Applicable to Bombay. Bombay: Education Society’s Press.Google Scholar
Dalzell, N. A. (1869). Extracts on Forests and Forestry. Bombay: Education Society’s Press.Google Scholar
Darwin, C. (1859). On the Origin of Species by Means of Natural Selection. London: John Murray.Google Scholar
Davidson, E. A., de Araújo, A. C., Artaxo, P., et al. (2012). The Amazon basin in transition. Nature, 481, 321328.Google Scholar
Davies-Barnard, T., Ridgwell, A., Singarayer, J., and Valdes, P. (2017). Quantifying the influence of the terrestrial biosphere on glacial–interglacial climate dynamics. Climate of the Past, 13, 13811401.Google Scholar
Davies-Barnard, T., Valdes, P. J., Singarayer, J. S., Pacifico, F. M., and Jones, C. D. (2014). Full effects of land use change in the representative concentration pathways. Environmental Research Letters, 9, 114014, DOI: https://doi.org/10.1088/1748-9326/9/11/114014.Google Scholar
Davin, E. L., and de Noblet-Ducoudré, N. (2010). Climatic impact of global-scale deforestation: Radiative versus nonradiative processes. Journal of Climate, 23, 97112.Google Scholar
Davin, E. L., de Noblet-Ducoudré, N., and Friedlingstein, P. (2007). Impact of land cover change on surface climate: Relevance of the radiative forcing concept. Geophysical Research Letters, 34, L13702, DOI: https://doi.org/10.1029/2007GL029678.Google Scholar
Davin, E. L., Rechid, D., Breil, M., et al. (2020). Biogeophysical impacts of forestation in Europe: First results from the LUCAS (Land Use and Climate Across Scales) regional climate model intercomparison. Earth System Dynamics, 11, 183200.Google Scholar
Davis, D. K. (2016a). The Arid Lands: History, Power, Knowledge. Cambridge, MA: Massachusetts Institute of Technology Press.Google Scholar
Davis, K. T., Dobrowski, S. Z., Holden, Z. A., Higuera, P. E., and Abatzoglou, J. T. (2019). Microclimatic buffering in forests of the future: The role of local water balance. Ecography, 42, 111.Google Scholar
Davis, M. (2016b). The coming desert: Kropotkin, Mars and the pulse of Asia. New Left Review, 97, 2343.Google Scholar
Davis, M. B. (1981). Quaternary history and the stability of forest communities. In Forest Succession: Concepts and Application, edited by West, D. C., Shugart, H. H., and Botkin, D. B.. New York: Springer-Verlag, pp. 132153.Google Scholar
Dawson, T. E. (1998). Fog in the California redwood forest: Ecosystem inputs and use by plants. Oecologia, 117, 476485.Google Scholar
De Frenne, P., Lenoir, J., Luoto, M., et al. (2021). Forest microclimates and climate change: Importance, drivers and future research agenda. Global Change Biology, 27, 22792297.Google Scholar
De Frenne, P., Rodríguez-Sánchez, F., Coomes, D. A., et al. (2013). Microclimate moderates plant responses to macroclimate warming. Proceedings of the National Academy of Sciences USA, 110, 1856118565.Google Scholar
De Frenne, P., Zellweger, F., Rodríguez-Sánchez, F., et al. (2019). Global buffering of temperatures under forest canopies. Nature Ecology and Evolution, 3, 744749.Google Scholar
de Noblet, N. I., Prentice, I. C., Joussaume, S., et al. (1996). Possible role of atmosphere–biosphere interactions in triggering the last glaciation. Geophysical Research Letters, 23, 31913194.Google Scholar
de Noblet-Ducoudré, N., Boisier, J.-P., Pitman, A., et al. (2012). Determining robust impacts of land-use-induced land cover changes on surface climate over North America and Eurasia: Results from the first set of LUCID experiments. Journal of Climate, 25, 32613281.Google Scholar
Denys, N. (1672). Description geographique et historique des costes de l’Amerique septentrionale, 2 vols. Paris: Claude Barbin.Google Scholar
Devaraju, N., Bala, G., and Modak, A. (2015). Effects of large-scale deforestation on precipitation in the monsoon regions: Remote versus local effects. Proceedings of the National Academy of Sciences USA, 112, 32573262.Google Scholar
Devaraju, N., de Noblet-Ducoudré, N., Quesada, B., and Bala, G. (2018). Quantifying the relative importance of direct and indirect biophysical effects of deforestation on surface temperature and teleconnections. Journal of Climate, 31, 38113829.Google Scholar
Dhar, A., Parrott, L., and Heckbert, S. (2016). Consequences of mountain pine beetle outbreak on forest ecosystem services in western Canada. Canadian Journal of Forest Research, 46, 987999.Google Scholar
Dickerson-Lange, S. E., Vano, J. A., Gersonde, R., and Lundquist, J. D. (2021). Ranking forest effects on snow storage: A decision tool for forest management. Water Resources Research, 57, e2020WR027926, DOI: https://doi.org/10.1029/2020WR027926.Google Scholar
Dickinson, R. E. (1983). Land surface processes and climate–surface albedos and energy balance. Advances in Geophysics, 25, 305353.Google Scholar
Dickinson, R. E. (1984). Modeling evapotranspiration for three-dimensional global climate models. In Climate Processes and Climate Sensitivity, edited by Hansen, J. E. and Takahashi, T.. Washington, DC: American Geophysical Union, pp. 5872.Google Scholar
Dickinson, R. E., and Henderson-Sellers, A. (1988). Modelling tropical deforestation: A study of GCM land-surface parameterizations. Quarterly Journal of the Royal Meteorological Society, 114, 439462.Google Scholar
Dickinson, R. E., Henderson-Sellers, A., Kennedy, P. J., and Wilson, M. F. (1986). Biosphere–Atmosphere Transfer Scheme (BATS) for the NCAR Community Climate Model, Technical Note NCAR/TN-275+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Dickinson, R. E., Henderson-Sellers, A., and Kennedy, P. J. (1993). Biosphere–Atmosphere Transfer Scheme (BATS) Version 1e as Coupled to the NCAR Community Climate Model, Technical Note NCAR/TN-387+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Dickinson, R. E., Jäger, J., Washington, W. M., and Wolski, R. (1981). Boundary Subroutine for the NCAR Global Climate Model, Technical Note NCAR/TN-173+IA. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Dirmeyer, P. A., and Shukla, J. (1996). The effect on regional and global climate of expansion of the world’s deserts. Quarterly Journal of the Royal Meteorological Society, 122, 451482.Google Scholar
Dirmeyer, P. A., Balsamo, G., Blyth, E. M., Morrison, R., and Cooper, H. M. (2021). Land-atmosphere interactions exacerbated the drought and heatwave over northern Europe during summer 2018. AGU Advances, 2, e2020AV000283, DOI: https://doi.org/10.1029/2020AV000283.Google Scholar
Dodds, G. B. (1969). The stream-flow controversy: A conservation turning point. Journal of American History, 56, 5969.Google Scholar
Douglas, W. O. (1951). Strange Lands and Friendly People. New York: Harper.Google Scholar
Douglass, W. (1749–51). A Summary, Historical and Political, of the First Planting, Progressive Improvements, and Present State of the British Settlements in North-America, 2 vols. Boston: Rogers & Fowle.Google Scholar
Dove, H. W. (1855). Ueber die Vertheilung der Regen in der gemäfsigten Zone. Annalen der Physik und Chemie, 94, 4259.Google Scholar
Dove, H. W. (1855–56). On the distribution of rain in the temperate zone. American Journal of Science and Arts, 20, 397402; 21, 112117.Google Scholar
Drever, C. R., Cook-Patton, S. C., Akhter, F., et al. (2021). Natural climate solutions for Canada. Science Advances, 7, eabd6034, DOI: https://doi.org/10.1126/sciadv.abd6034.Google Scholar
Drori, J. (2018). Around the World in 80 Trees. London: Laurence King.Google Scholar
Dubos, J.-B (1719). Reflexions critiques sur la poesie et sur la peinture, 2 vols. Paris: Jean Mariette.Google Scholar
Duffey, E. (1964). The terrestrial ecology of Ascension Island. Journal of Applied Ecology, 1, 219251.Google Scholar
Duhamel du Monceau, H.-L. (1746). Observations botanico-météorologiques faites à Québec pendant les mois d’Octobre, Novembre & Décembre 1744, & les mois de Janvier, Février, Mars, Avril & Mai 1745. Mémoires de mathématique et de physique, tirés des registres de l’Académie Royale des Sciences, de l’année 1746, pp.8897.Google Scholar
Duhamel du Monceau, H.-L. (1755). Traité des arbres et arbustes qui se cultivent en France en pleine terre, 2 vols. Paris: H. L. Guerin & L. F. Delatour.Google Scholar
Duhamel du Monceau, H.-L. (1758). La physique des arbres, 2 vols. Paris: H. L. Guerin & L. F. Delatour.Google Scholar
Duhamel du Monceau, H.-L. (1760). Des semis et plantations des arbres, et de leur culture. Paris: H. L. Guerin & L. F. Delatour.Google Scholar
Duhamel du Monceau, H.-L. (1764). De l’exploitation des bois, 2 vols. Paris: H. L. Guerin & L. F. Delatour.Google Scholar
Duhamel du Monceau, H.-L. (1767). Du transport, de la conservation et de la force des bois. Paris: L. F. Delatour.Google Scholar
Dunbar, J. (1780). Essays on the History of Mankind in Rude and Cultivated Ages. London: W. Strahan.Google Scholar
Dunbar, W. (1809). Meteorological observations. Transactions of the American Philosophical Society, 6, 4355.Google Scholar
Duveiller, G., Caporaso, L., Abad-Viñas, R., et al. (2020). Local biophysical effects of land use and land cover change: Towards an assessment tool for policy makers. Land Use Policy, 91, 104382, DOI: https://doi.org/10.1016/j.landusepol.2019.104382.Google Scholar
Duveiller, G., Filipponi, F., Ceglar, A., et al. (2021). Revealing the widespread potential of forests to increase low level cloud cover. Nature Communications, 12, 4337, DOI: https://doi.org/10.1038/s41467-021-24551-5.Google Scholar
Duveiller, G., Hooker, J., and Cescatti, A. (2018). The mark of vegetation change on Earth’s surface energy balance. Nature Communications, 9, 679, DOI: https://doi.org/10.1038/s41467-017-02810-8.Google Scholar
Dwight, T. (1821–22). Travels; in New-England and New-York, 4 vols. New Haven, CT: S. Converse.Google Scholar
Ébelmen, J.-J. (1845). Sur les produits de la décomposition des espèces minérales de la famille des silicates. Annales des mines (série 4), 7, 366.Google Scholar
Eberle, J. J., and Greenwood, D. R. (2012). Life at the top of the greenhouse Eocene world: A review of the Eocene flora and vertebrate fauna from Canada’s High Arctic. Geological Society of America Bulletin, 124, 323.Google Scholar
Ebermayer, E. (1873). Die physikalischen Einwirkungen des Waldes auf Luft und Boden und seine klimatologische und hygienische Bedeutung. Aschaffenburg: C. Krebs.Google Scholar
Edburg, S. L., Hicke, J. A., Brooks, P. D., et al. (2012). Cascading impacts of bark beetle-caused tree mortality on coupled biogeophysical and biogeochemical processes. Frontiers in Ecology and the Environment, 10, 416424.Google Scholar
Eder, F., De Roo, F., Rotenberg, E., et al. (2015). Secondary circulations at a solitary forest surrounded by semi-arid shrubland and their impact on eddy-covariance measurements. Agricultural and Forest Meteorology, 211–212, 115127.Google Scholar
Edge, T. J. (1878). The forests of our state: Their value and their influence upon streams, temperature, climate, and rain-fall. In First Annual Report of the Pennsylvania Board of Agriculture for the Year 1877, with an Appendix. Harrisburg: Lane S. Hart, pp. 6177.Google Scholar
Edwards, P. N. (2011). History of climate modeling. WIREs Climate Change, 2, 128139.Google Scholar
Egerton, F. N. (2003). A history of the ecological sciences, part 9. Albertus Magnus: A scholastic naturalist. Bulletin of the Ecological Society of America, 84, 8791.Google Scholar
Egerton, F. N. (2008). A history of the ecological sciences, part 28: Plant growth studies during the 1700s. Bulletin of the Ecological Society of America, 89, 159175.Google Scholar
Egleston, N. H. (1883). Forestry division. In Report of the Commissioner of Agriculture for the Year 1883. Washington, DC: Government Printing Office, pp. 444462.Google Scholar
Egleston, N. H. (1884). Report of chief of the forestry bureau. In Report of the Commissioner of Agriculture for the Year 1884. Washington, DC: Government Printing Office, pp. 137180.Google Scholar
Egleston, N. H. (1885). Report of chief of division of forestry. In Report of the Commissioner of Agriculture, 1885. Washington, DC: Government Printing Office, pp. 183206.Google Scholar
Egleston, N. H. (1896). Arbor Day: Its History and Observance. Washington, DC: Government Printing Office.Google Scholar
Elliott, R. S. (1871a). Report on the industrial resources of western Kansas and eastern Colorado. In Preliminary Report of the United States Geological Survey of Wyoming, and Portions of Contiguous Territories. Washington, DC: Government Printing Office, pp. 442458.Google Scholar
Elliott, R. S. (1871b). Climate of Kansas. In Annual Report of the Board of Regents of the Smithsonian Institution, Showing the Operations, Expenditures, and Condition of the Institution for the Year 1870. Washington, DC: Government Printing Office, pp. 472474.Google Scholar
Elliott, R. S. (1872). Experiments in cultivation on the plains along the line of the Kansas Pacific Railway. In Preliminary Report of the United States Geological Survey of Montana and Portions of Adjacent Territories. Washington, DC: Government Printing Office, pp. 274279.Google Scholar
Elliott, R. S. (1883). Notes Taken in Sixty Years. St. Louis: R. P. Studley.Google Scholar
Ellis, E. C. (2021). Land use and ecological change: A 12,000-year history. Annual Review of Environment and Resources, 46, 133.Google Scholar
Ellison, D., and Ifejika Speranza, C. (2020). From blue to green water and back again: Promoting tree, shrub and forest-based landscape resilience in the Sahel. Science of the Total Environment, 739, 140002, DOI: https://doi.org/10.1016/j.scitotenv.2020.140002.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, 806820.Google Scholar
Ellison, D., Morris, C. E., Locatelli, B., et al. (2017). Trees, forests and water: Cool insights for a hot world. Global Environmental Change, 43, 5161.Google Scholar
Ellsworth, D. S., and Reich, P. B. (1993). Canopy structure and vertical patterns of photosynthesis and related leaf traits in a deciduous forest. Oecologia, 96, 169178.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, 861880.Google Scholar
Emerson, G. B. (1846). A Report on the Trees and Shrubs Growing Naturally in the Forests of Massachusetts. Boston: Dutton & Wentworth.Google Scholar
Emerson, R. W. (1870). Society and Solitude: 12 Chapters. Boston: Fields, Osgood & Co.Google Scholar
Emmons, D. M. (1971). Theories of increased rainfall and the Timber Culture Act of 1873. Forest History, 15(3), 614.Google Scholar
Erb, K.-H., Luyssaert, S., Meyfroidt, P., et al. (2017). Land management: Data availability and process understanding for global change studies. Global Change Biology, 23, 512533.Google Scholar
Evelyn, J. (1664). Sylva, or a Discourse of Forest-Trees, and the Propagation of Timber in His Majesties Dominions. London: John Martyn & James Allestry.Google Scholar
Evelyn, J. (1670). Sylva, or a Discourse of Forest-Trees, and the Propagation of Timber in His Majesties Dominions, 2nd ed. London: John Martyn & James Allestry.Google Scholar
Evelyn, J. (1679). Sylva, or a Discourse of Forest-Trees, and the Propagation of Timber in His Majesties Dominions, 3rd ed. London: John Martyn.Google Scholar
Evelyn, J. (1706). Silva, or a Discourse of Forest-Trees, and the Propagation of Timber in His Majesty’s Dominions, 4th ed. London: Robert Scott; Richard Chiswell; and others.Google Scholar
Evelyn, J. (1776). Silva: Or, a Discourse of Forest-Trees, and the Propagation of Timber in His Majesty’s Dominions, with notes by Hunter, A.. York: A. Ward.Google Scholar
Fabre, J.-A. (1797). Essai sur la théorie des torrens et des rivières. Paris: Bidault.Google Scholar
Fantham, E. (2004). Ovid’s Metamorphoses. New York: Oxford University Press.Google Scholar
FAO (2018). Global Forest Resources Assessment 2020: Terms and Definitions, Forest Resources Assessment Working Paper 188. Rome: Food and Agriculture Organization of the United Nations.Google Scholar
FAO (2020). Global Forest Resources Assessment 2020: Main report. Rome: Food and Agriculture Organization of the United Nations; https://doi.org/10.4060/ca9825en.Google Scholar
Fargione, J., Hill, J., Tilman, D., Polasky, S., and Hawthorne, P. (2008). Land clearing and the biofuel carbon debt. Science, 319, 12351238.Google Scholar
Fargione, J. E., Bassett, S., Boucher, T., et al. (2018). Natural climate solutions for the United States. Science Advances, 4, eaat1869, DOI: https://doi.org/10.1126/sciadv.aat1869.Google Scholar
Fasullo, J. T., Rosenbloom, N., Buchholz, R. R., et al. (2021). Coupled climate responses to recent Australian wildfire and COVID-19 emissions anomalies estimated in CESM2. Geophysical Research Letters, 48, e2021GL093841, DOI: https://doi.org/10.1029/2021GL093841.Google Scholar
Favero, A., Sohngen, B., Huang, Y., and Jin, Y. (2018). Global cost estimates of forest climate mitigation with albedo: A new integrative policy approach. Environmental Research Letters, 13, 125002, DOI: https://doi.org/10.1088/1748-9326/aaeaa2.Google Scholar
Feddema, J. J., Oleson, K. W., Bonan, G. B., et al. (2005). The importance of land-cover change in simulating future climates. Science, 310, 16741678.Google Scholar
Feng, D., Bao, W., Yang, Y., and Fu, M. (2021). How do government policies promote greening? Evidence from China. Land Use Policy, 104, 105389, DOI: https://doi.org/10.1016/j.landusepol.2021.105389.CrossRefGoogle Scholar
Feng, X., Fu, B., Piao, S., et al. (2016). Revegetation in China’s Loess Plateau is approaching sustainable water resource limits. Nature Climate Change, 6, 10191022.Google Scholar
Ferguson, D. K., and Knobloch, E. (1998). A fresh look at the rich assemblage from the Pliocene sink-hole of Willershausen, Germany. Review of Palaeobotany and Palynology, 101, 271286.Google Scholar
Fernández de Oviedo y Valdés, G. (1851–55). Historia general y natural de las Indias, islas y tierra-firme del Mar Océano, edited by Amador de los Rios, José, 4 vols. Madrid: Imprenta de la Real Academia de la Historia.Google Scholar
Fernow, B. E. (1887). Report of chief of forestry division. In Report of the Commissioner of Agriculture, 1886. Washington, DC: Government Printing Office, pp. 149226.Google Scholar
Fernow, B. E. (1888). Report of the chief of forestry division. In Report of the Commissioner of Agriculture, 1887. Washington, DC: Government Printing Office, pp. 605616.Google Scholar
Fernow, B. E. (1889). Report of the chief of forestry division. In Report of the Commissioner of Agriculture, 1888. Washington, DC: Government Printing Office, pp. 597641.Google Scholar
Fernow, B. E. (1891). What Is Forestry? U.S. Department of Agriculture, Forestry Division Bulletin Number 5. Washington, DC: Government Printing Office.Google Scholar
Fernow, B. E. (1893a). Forest influences: Introduction and summary of conclusions. In Forest Influences (U.S. Department of Agriculture, Forestry Division Bulletin Number 7), edited by Fernow, B. E.. Washington, DC: Government Printing Office, pp. 922.Google Scholar
Fernow, B. E. (1893b). Relation of forests to water supplies. In Forest Influences (U.S. Department of Agriculture, Forestry Division Bulletin Number 7), edited by Fernow, B. E.. Washington, DC: Government Printing Office, pp. 123170.Google Scholar
Fernow, B. E. (1894). Forest conditions and forestry problems in the United States. Proceedings of the American Forestry Association, 10, 2936.Google Scholar
Fernow, B. E. (1910). Current Literature: The Influence of Forests on Climate and on Floods by Willis L. Moore. Forestry Quarterly, 8, 7475.Google Scholar
Ferrel, W. (1889). Note on the influence of forests upon rainfall. American Meteorological Journal, 5(10), 433435.Google Scholar
Findell, K. L., Berg, A., Gentine, P., et al. (2017). The impact of anthropogenic land use and land cover change on regional climate extremes. Nature Communications, 8, 989, DOI: https://doi.org/10.1038/s41467-017-01038-w.Google Scholar
Firth, J. C. (1874). On forest culture. Transactions and Proceedings of the New Zealand Institute, 7, 181195.Google Scholar
Fischer, E. M., Seneviratne, S. I., Lüthi, D., and Schär, C. (2007). Contribution of land–atmosphere coupling to recent European summer heat waves. Geophysical Research Letters, 34, L06707, DOI: https://doi.org/10.1029/2006GL029068.CrossRefGoogle Scholar
FitzNigel, R. (1983). Dialogus de Scaccario: The Course of the Exchequer, edited and translated by Johnson, C. with corrections by Carter, F. E. L. and Greenway, D. E.. Oxford: Oxford University Press.Google Scholar
Fleming, J. R. (1998). Historical Perspectives on Climate Change. New York: Oxford University Press.Google Scholar
Foley, J. A., Coe, M. T., Scheffer, M., and Wang, G. (2003). Regime shifts in the Sahara and Sahel: Interactions between ecological and climatic systems in northern Africa. Ecosystems, 6, 524539.CrossRefGoogle Scholar
Foley, J. A., DeFries, R., Asner, G. P., et al. (2005). Global consequences of land use. Science, 309, 570574.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, 5254.Google Scholar
Folland, C. K., Palmer, T. N., and Parker, D. E. (1986). Sahel rainfall and worldwide sea temperatures, 1901–85. Nature, 320, 602607.Google Scholar
Ford, C. (2016). Natural Interests: The Contest over Environment in Modern France. Cambridge, MA: Harvard University Press.Google Scholar
Ford, C. R., Laseter, S. H., Swank, W. T., and Vose, J. M. (2011). Can forest management be used to sustain water-based ecosystem services in the face of climate change? Ecological Applications, 21, 20492067.Google Scholar
Forkel, M., Carvalhais, N., Rödenbeck, C., et al. (2016). Enhanced seasonal CO2 exchange caused by amplified plant productivity in northern ecosystems. Science, 351, 696699.Google Scholar
Forman, R. T. T., and Godron, M. (1986). Landscape Ecology. New York: Wiley.Google Scholar
Forry, S. (1842). The Climate of the United States and Its Endemic Influences. New York: J. & H. G. Langley.Google Scholar
Forry, S. (1844). Researches in elucidation of the distribution of heat over the globe, and especially of the climatic features peculiar to the region of the United States. American Journal of Science and Arts, 47, 1850, 221241.Google Scholar
Forster, E. J., Healey, J. R., Dymond, C., and Styles, D. (2021). Commercial afforestation can deliver effective climate change mitigation under multiple decarbonisation pathways. Nature Communications, 12, 3831, DOI: https://doi.org/10.1038/s41467-021-24084-x.Google Scholar
Forzieri, G., Alkama, R., Miralles, D. G., and Cescatti, A. (2017). Satellites reveal contrasting responses of regional climate to the widespread greening of Earth. Science, 356, 11801184.Google Scholar
Fowler, J. (1774). A Summary Account of the Present Flourishing State of the Respectable Colony of Tobago, in the British West Indies. London: A. Grant.Google Scholar
Fowler, M. D., Kooperman, G. J., Randerson, J. T., and Pritchard, M. S. (2019). The effect of plant physiological responses to rising CO2 on global streamflow. Nature Climate Change, 9, 873879.Google Scholar
Fraas, C. (1847). Klima und Pflanzenwelt in der Zeit, ein Beitrag zur Geschichte beider. Landshut: J. G. Wölfle.Google Scholar
Frankenfield, H. C. (1910). The experiment station at Wagon Wheel Gap, Colo. Monthly Weather Review, 38, 14531455.Google Scholar
Franklin, B. (1755). Observations concerning the increase of mankind, peopling of countries, etc. In Observations on the Late and Present Conduct of the French, with Regard to Their Encroachments upon the British Colonies in North America. Together with Remarks on the Importance of These Colonies to Great-Britain, edited by Clarke, W.. Boston: S. Kneeland, pp. Appendix, 115.Google Scholar
Franklin, B. (1905). Observations concerning the increase of mankind, peopling of countries, etc. In The Writings of Benjamin Franklin, vol. 3: 1750–1759, edited by Smyth, A. H.. New York: Macmillan, pp. 6373.Google Scholar
Franklin, B. (1966). To Ezra Stiles, May 29, 1763. In The Papers of Benjamin Franklin, vol. 10. January 1, 1762, through December 31, 1763, edited by Labaree, L. W.. New Haven, CT: Yale University Press, pp. 264267.Google Scholar
Franklin, J. (1784). The Philosophical and Political History of the Thirteen United States of America. London: J. Hinton & W. Adams.Google Scholar
Franks, P. J., Adams, M. A., Amthor, J. S., et al. (2013). Sensitivity of plants to changing atmospheric CO2 concentration: From the geological past to the next century. New Phytologist, 197, 10771094.Google Scholar
Franks, P. J., Royer, D. L., Beerling, D. J., et al. (2014). New constraints on atmospheric CO2 concentration for the Phanerozoic. Geophysical Research Letters, 41, 46854694.CrossRefGoogle Scholar
Fressoz, J.-B. (2015). Losing the Earth knowingly: Six environmental grammars around 1800. In The Anthropocene and the Global Environmental Crisis: Rethinking Modernity in a New Epoch, edited by Hamilton, C., Bonneuil, C., and Gemenne, F.. New York: Routledge, pp. 7083.Google Scholar
Fressoz, J.-B., and Locher, F. (2015). Régénérer la nature, restaurer les climats: François-Antoine Rauch et les Annales Européennes de physique végétale et d’economie publique, 1815-1830. Le Temps des medias, 25(2), 5269.Google Scholar
Fressoz, J.-B., and Locher, F. (2020). Les révoltes du ciel: une histoire du changement climatique (XVe–XXe siècle). Paris: Éditions du Seuil.Google Scholar
Friedlingstein, P., O’Sullivan, M., Jones, M. W., et al. (2020). Global carbon budget 2020. Earth System Science Data, 12, 32693340.Google Scholar
Gallimore, R. G., and Kutzbach, J. E. (1996). Role of orbitally induced changes in tundra area in the onset of glaciation. Nature, 381, 503505.Google Scholar
Galloway, J. N., Aber, J. D., Erisman, J. W., et al. (2003). The nitrogen cascade. BioScience, 53, 341356.Google Scholar
Galloway, J. N., Townsend, A. R., Erisman, J. W., et al. (2008). Transformation of the nitrogen cycle: Recent trends, questions, and potential solutions. Science, 320, 889892.Google Scholar
Galvez, M. E., and Gaillardet, J. (2012). Historical constraints on the origins of the carbon cycle concept. Comptes Rendus Geoscience, 344, 549567.Google Scholar
Galway-Witham, J., and Stringer, C. (2018). How did Homo sapiens evolve? Science, 360, 12961298.Google Scholar
Gannett, H. (1888a). Do forests influence rainfall? Science, 11 (257), 3–5 (January 6, 1888).Google Scholar
Gannett, H. (1888b). Is the rainfall increasing upon the plains? Science, 11 (265), 99100 (March 2, 1888).Google Scholar
Gash, J. H. C., and Nobre, C. A. (1997). Climatic effects of Amazonian deforestation: Some results from ABRACOS. Bulletin of the American Meteorological Society, 78, 823830.Google Scholar
Gash, J. H. C., Nobre, C. A., Roberts, J. M., and Victoria, R. L. (1996). Amazonian Deforestation and Climate. New York: Wiley.Google Scholar
Gates, D. M. (1963). Leaf temperature and energy exchange. Archiv für Meteorologie, Geophysik und Bioklimatologie, 12B, 321336.Google Scholar
Gatti, L. V., Basso, L. S., Miller, J. B., et al. (2021). Amazonia as a carbon source linked to deforestation and climate change. Nature, 595, 388393.Google Scholar
Gaubert, B., Stephens, B. B., Basu, S., et al. (2019). Global atmospheric CO2 inverse models converging on neutral tropical land exchange, but disagreeing on fossil fuel and atmospheric growth rate. Biogeosciences, 16, 117134.CrossRefGoogle ScholarPubMed
Ge, J., Guo, W., Pitman, A. J., et al. (2019). The nonradiative effect dominates local surface temperature change caused by afforestation in China. Journal of Climate, 32, 44454471.Google Scholar
Ge, J., Pitman, A. J., Guo, W., Zan, B., and Fu, C. (2020). Impact of revegetation of the Loess Plateau of China on the regional growing season water balance. Hydrology and Earth System Sciences, 24, 515533.Google Scholar
Gedney, N., Cox, P. M., Betts, R. A., et al. (2006). Detection of a direct carbon dioxide effect in continental river runoff records. Nature, 439, 835838.Google Scholar
Geiger, R. (1927). Das Klima der bodennahen Luftschicht. Braunschweig: Friedr. Vieweg.Google Scholar
Geiger, R. (1942). The Climate of the Layer of Air near the Ground, translated by J. Leighly for restricted official use of the Soil Conservation Service, United States Department of Agriculture. New Philadelphia, OH: Muskingum Climatic Research Center.Google Scholar
Geiger, R. (1950). The Climate near the Ground, translation by M. N. Stewart and others of the second German edition of Das Klima der bodennahen Luftschicht with revisions and enlargements by the author. Cambridge, MA: Harvard University Press.Google Scholar
Geis, D. (1981). Walt Disney’s Treasury of Silly Symphonies. New York: Harry N. Abrams.Google Scholar
Genesio, L., Bassi, R., and Miglietta, F. (2021). Plants with less chlorophyll: A global change perspective. Global Change Biology, 27, 959967.Google Scholar
Gentine, P., Green, J. K., Guérin, M., et al. (2019a). Coupling between the terrestrial carbon and water cycles: A review. Environmental Research Letters, 14, 083003, DOI: https://doi.org/10.1088/1748-9326/ab22d6.CrossRefGoogle Scholar
Gentine, P., Massmann, A., Lintner, B. R., et al. (2019b). Land–atmosphere interactions in the tropics: A review. Hydrology and Earth System Science, 23, 41714197.CrossRefGoogle Scholar
Giannini, A., Biasutti, M., and Verstraete, M. M. (2008). A climate model-based review of drought in the Sahel: Desertification, the re-greening and climate change. Global and Planetary Change, 64, 119128.Google Scholar
Giannini, A., Saravanan, R., and Chang, P. (2003). Oceanic forcing of Sahel rainfall on interannual to interdecadal time scales. Science, 302, 10271030.Google Scholar
Gidden, M. J., Riahi, K., Smith, S. J., et al. (2019). Global emissions pathways under different socioeconomic scenarios for use in CMIP6: A dataset of harmonized emissions trajectories through the end of the century. Geoscientific Model Development, 12, 14431475.Google Scholar
Gilpin, W. (1791). Remarks on Forest Scenery, and Other Woodland Views, 2 vols. London: R. Blamire.Google Scholar
Girardin, C. A. J., Jenkins, S., Seddon, N., et al. (2021). Nature-based solutions can help cool the planet: If we act now. Nature, 593, 191194.Google Scholar
Giraud-Soulavie, J.-L. (1783). Histoire naturelle de la France méridionale. Second partie. Les végétaux, vol. 1. Paris: Quillau; Merigot l’ainé; Merigot jeune; Belin.Google Scholar
Glacken, C. J. (1956). Changing ideas of the habitable world. In Man’s Role in Changing the Face of the Earth, edited by Thomas, W. L., Jr. Chicago: University of Chicago Press, pp. 7092.Google Scholar
Glacken, C. J. (1967). Traces on the Rhodian Shore. Berkeley: University of California Press.Google Scholar
Gleason, K. E., McConnell, J. R., Arienzo, M. M., Chellman, N., and Calvin, W. M. (2019). Four-fold increase in solar forcing on snow in western U.S. burned forests since 1999. Nature Communications, 10, 2026, DOI: https://doi.org/10.1038/s41467-019-09935-y.Google Scholar
Gleason, K. E., Nolin, A. W., and Roth, T. R. (2013). Charred forests increase snowmelt: Effects of burned woody debris and incoming solar radiation on snow ablation. Geophysical Research Letters, 40, 46544661.Google Scholar
Goeking, S. A., and Tarboton, D. G. (2020). Forests and water yield: A synthesis of disturbance effects on streamflow and snowpack in western coniferous forests. Journal of Forestry, 118, 172192.Google Scholar
Golinski, J. (2008). American climate and the civilization of nature. In Science and Empire in the Atlantic World, edited by Delbourgo, J. and Dew, N.. New York: Routledge, pp. 153174.Google Scholar
Golley, F. B. (1993). A History of the Ecosystem Concept in Ecology: More than the Sum of the Parts. New Haven, CT: Yale University Press.Google Scholar
Goss, M., Swain, D. L., Abatzoglou, J. T., et al. (2020). Climate change is increasing the likelihood of extreme autumn wildfire conditions across California. Environmental Research Letters, 15, 094016, DOI: https://doi.org/10.1088/1748-9326/ab83a7.Google Scholar
Gould, S. J. (1989). Church, Humboldt, and Darwin: The tension and harmony of art and science. In Frederic Edwin Church, edited by Kelly, F.. Washington, DC: National Gallery of Art, pp. 94107.Google Scholar
Goulden, M. L., Winston, G. C., McMillan, A. M. S., et al. (2006). An eddy covariance mesonet to measure the effect of forest age on land–atmosphere exchange. Global Change Biology, 12, 21462162.Google Scholar
Grassi, G., House, J., Dentener, F., et al. (2017). The key role of forests in meeting climate targets requires science for credible mitigation. Nature Climate Change, 7, 220226.Google Scholar
Graven, H. D., Keeling, R. F., Piper, S. C., et al. (2013). Enhanced seasonal exchange of CO2 by northern ecosystems since 1960. Science, 341, 10851089.Google Scholar
Greeley, H. (1871). What I Know of Farming: A Series of Brief and Plain Expositions of Practical Agriculture as an Art Based upon Science. New York: G. W. Carleton.Google Scholar
Gregg, J. (1844). Commerce of the Prairies: Or the Journal of a Sante Fé Trader, during Eight Expeditions across the Great Western Prairies, and a Residence of Nearly Nine Years in Northern Mexico, 2 vols. New York: Henry G. Langley.Google Scholar
Grew, N. (1682). The Anatomy of Plants, with an Idea of a Philosophical History of Plants. [London]: W. Rawlins.Google Scholar
Griscom, B. W., Adams, J., Ellis, P. W., et al. (2017). Natural climate solutions. Proceedings of the National Academy of Sciences USA, 114, 1164511650.Google Scholar
Griscom, B. W., Busch, J., Cook-Patton, S. C., et al. (2020). National mitigation potential from natural climate solutions in the tropics. Philosophical Transactions of the Royal Society London B, 375, 20190126, DOI: https://doi.org/10.1098/rstb.2019.0126.Google Scholar
Grove, R. (1989). Scottish missionaries, evangelical discourses and the origins of conservation thinking in southern Africa 1820–1900. Journal of Southern African Studies, 15, 163187.Google Scholar
Grove, R. H. (1995). Green Imperialism: Colonial Expansion, Tropical Island Edens and the Origins of Environmentalism, 1600–1860. Cambridge, UK: Cambridge University Press.Google Scholar
Gu, L., Baldocchi, D. D., Wofsy, S. C., et al. (2003). Response of a deciduous forest to the Mount Pinatubo eruption: Enhanced photosynthesis. Science, 299, 20352038.Google Scholar
Guyot, C. (1898). L’Enseignement forestier en France: l’école de Nancy. Nancy: Crépin-Leblond.Google Scholar
Habenicht, R. E. (1963). John Heywood’s A Dialogue of Proverbs, edited, with introduction, commentary, and indexes. Berkeley: University of California Press.Google Scholar
Hack, J. J., Boville, B. A., Briegleb, B. P., et al. (1993). Description of the NCAR Community Climate Model (CCM2), Technical Note NCAR/TN-382+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Haeckel, E. (1866). Generelle Morphologie der Organismen: allgemeine Grundzuüge der organischen Formen-Wissenschaft, mechanisch begründet durch die von Charles Darwin reformirte Descendenz-Theorie, 2 vols. Berlin: Georg Reimer.Google Scholar
Haesen, S., Lembrechts, J. J., De Frenne, P., et al. (2021). ForestTemp: Sub-canopy microclimate temperatures of European forests. Global Change Biology, 27, 63076319.Google Scholar
Hale, T. (1758–59). A Compleat Body of Husbandry, 2nd ed., 4 vols. London: T. Osborne; T. Trye; S. Crowder.Google Scholar
Hales, S. (1727). Vegetable Staticks: Or, an Account of Some Statical Experiments on the Sap in Vegetables: Being an Essay towards a Natural History of Vegetation. London: W. & J. Innys; T. Woodward.Google Scholar
Halim, M. A., Chen, H. Y. H., and Thomas, S. C. (2019). Stand age and species composition effects on surface albedo in a mixedwood boreal forest. Biogeosciences, 16, 43574375.Google Scholar
Hall, B., Motzkin, G., Foster, D. R., Syfert, M., and Burk, J. (2002). Three hundred years of forest and land-use change in Massachusetts, USA. Journal of Biogeography, 29, 13191335.Google Scholar
Hall, F. G., Betts, A. K., Frolking, S., et al. (2004). The boreal climate. In Vegetation, Water, Humans and the Climate: A New Perspective on an Interactive System, edited by Kabat, P., Claussen, M., Dirmeyer, P. A., et al. Berlin: Springer-Verlag, pp. 93114.Google Scholar
Halley, E. (1686). An historical account of the trade winds, and monsoons, observable in the seas between and near the tropicks, with an attempt to assign the phisical cause of the said winds. Philosophical Transactions, 16(183), 153168.Google Scholar
Halley, E. (1691). An account of the circulation of the watry vapours of the sea, and of the cause of springs, presented to the Royal Society. Philosophical Transactions 17(192), 468473.Google Scholar
Hann, J. (1867). Wald und Regen. Zeitschrift der österreichischen Gesellschaft für Meteorologie, 2, 129136.Google Scholar
Hann, J. (1869). Thatsachen und Bemerkungen über einige schädliche Folgen dr Zerstörung des natürlichen Pflanzenkleides der Erdoberfläche. Zeitschrift der österreichischen Gesellschaft für Meteorologie, 4, 1822.Google Scholar
Hann, J. (1888). Wald und Regen in Indien. Meteorologische Zeitschrift, 5, 235237.Google Scholar
Hann, J. (1897). Handbuch der Klimatologie, 3 vols. Stuttgart: J. Engelhorn.Google Scholar
Hann, J. (1903). Handbook of Climatology, part 1. General Climatology, translated by R. De Courcy Ward. New York: Macmillan.Google Scholar
Hansen, M. C., Potapov, P. V., Moore, R., et al. (2013). High-resolution global maps of 21st-century forest cover change. Science, 342, 850853.Google Scholar
Harrington, M. W. (1876). The Analysis of Plants: Intended for Schools and Colleges and for the Independent Botanical Student. Ann Arbor: Sheehan.Google Scholar
Harrington, M. W. (1877). The tropical ferns collected by Professor Steere in the years 1870–75. Journal of the Linnean Society: Botany, 16(89), 2537.Google Scholar
Harrington, M. W. (1887) Is the rain-fall increasing on the plains? American Meteorological Journal, 4(8), 369373.Google Scholar
Harrington, M. W. (1893). Review of forest meteorological observations: A study preliminary to the discussion of the relation of forests to climate. In Forest Influences (U.S. Department of Agriculture, Forestry Division Bulletin Number 7), edited by Fernow, B. E.. Washington, DC: Government Printing Office, pp. 23122.Google Scholar
Harris, J. (1710). Lexicon Technicum: Or, an Universal English Dictionary of Arts and Sciences, vol. 2. London: Dan. Brown; Tim. Goodwin; and others.Google Scholar
Harrison, R. P. (1992). Forests: The Shadow of Civilization. Chicago: University of Chicago Press.Google Scholar
Hawken, P. (2017). Drawdown: The Most Comprehensive Plan Ever Proposed to Reverse Global Warming. New York: Penguin Books.Google Scholar
Hayden, F. V. (1867). Ferdinand V. Hayden to Joseph S. Wilson, July 1, 1867. In Report of the Commissioner of General Land Office, for the Year 1867. Washington, DC: Government Printing Office, pp. 128181.Google Scholar
Hayden, F. V. (1869). Preliminary Field Report of the United States Geological Survey of Colorado and New Mexico. Washington, DC: Government Printing Office.Google Scholar
Hayes, K. J. (2008). The Road to Monticello: The Life and Mind of Thomas Jefferson. New York: Oxford University Press.Google Scholar
Hearn, M. P. (2000). The Annotated Wizard of Oz. New York: W. W. Norton.Google Scholar
Hemery, G., and Simblet, S. (2014). The New Sylva: A Discourse of Forest and Orchard Trees for the Twenty-first Century. London: Bloomsbury.Google Scholar
Henry, J. (1886). Scientific Writings of Joseph Henry, 2 vols. Washington, DC: Smithsonian Institution.Google Scholar
Henshilwood, C. S., and Marean, C. W. (2003). The origin of modern human behavior: Critique of the models and their test implications. Current Anthropology, 44, 627651.Google Scholar
Herbert, R., Stier, P., and Dagan, G. (2021). Isolating large-scale smoke impacts on cloud and precipitation processes over the Amazon with convection permitting resolution. Journal of Geophysical Research: Atmospheres, 126, e2021JD034615, DOI: https://doi.org/10.1029/2021JD034615.Google Scholar
Herder, J. G. (1800). Outlines of a Philosophy of the History of Man, translated by T. Churchill. London: J. Johnson.Google Scholar
Herschel, J. F. W. (1859). Physical geography. In The Encyclopædia Britannica, or Dictionary of Arts, Sciences, and General Literature, 8th ed., vol. 17. Edinburgh: Adam & Charles Black, pp. 569647.Google Scholar
Herschel, J. F. W. (1861). Physical Geography: From the Encyclopædia Britannica. Edinburgh: Adam & Charles Black.Google Scholar
Hey, R. (1837). The Spirit of the Woods. London: Longman, Rees, Orme, Brown, Green, & Longman.Google Scholar
Heywood, J. (1546). A dialogue conteinyng the nomber in effect of all the prouerbes in the englishe tongue compacte in a matter concernyng two maner of mariages. London: Thomas Berthelet (Text Creation Partnership, Ann Arbor, Michigan; http://name.umdl.umich.edu/A03168.0001.001).Google Scholar
Hibberd, S. (1855). Brambles and Bay Leaves: Essays on the Homely and the Beautiful. London: Longman, Brown, Green, & Longmans.Google Scholar
Hibbert, A. R. (1967). Forest treatment effects on water yield. In Forest Hydrology: Proceedings of a National Science Foundation Advanced Science Seminar Held at The Pennsylvania State University, University Park, Pennsylvania, Aug 29–Sept 10, 1965, edited by Sopper, W. E. and Lull, H. W.. Oxford: Pergamon Press, pp. 527543.Google Scholar
Hicke, J. A., Allen, C. D., Desai, A. R., et al. (2012). Effects of biotic disturbances on forest carbon cycling in the United States and Canada. Global Change Biology, 18, 734.Google Scholar
Higgins, R. (2017). Thoreau and the Language of Trees. Oakland: University of California Press.Google Scholar
Higuera, P. E., and Abatzoglou, J. T. (2021). Record-setting climate enabled the extraordinary 2020 fire season in the western United States. Global Change Biology, 27, 12.Google Scholar
Higuera, P. E., Shuman, B. N., and Wolf, K. D. (2021). Rocky Mountain subalpine forests now burning more than any time in recent millennia. Proceedings of the National Academy of Sciences USA, 118, e2103135118, DOI: https://doi.org/10.1073/pnas.2103135118.Google Scholar
Hinsdale, B. A., and Demmon, I. N. (1906). History of the University of Michigan. Ann Arbor: University of Michigan Press.Google Scholar
Hirsch, A. L., Guillod, B. P., Seneviratne, S. I., et al. (2018). Biogeophysical impacts of land-use change on climate extremes in low-emission scenarios: Results from HAPPI-Land. Earth’s Future, 6, 396409.Google Scholar
Hirschi, M., Seneviratne, S. I., Alexandrov, V., et al. (2011). Observational evidence for soil-moisture impact on hot extremes in southeastern Europe. Nature Geoscience, 4, 1721.Google Scholar
Hoelzmann, P., Jolly, D., Harrison, S. P., et al. (1998). Mid-Holocene land-surface conditions in northern Africa and the Arabian peninsula: A data set for the analysis of biogeophysical feedbacks in the climate system. Global Biogeochemical Cycles, 12, 3551.Google Scholar
Hogan, D. J. (2014). The Wizard of Oz FAQ: All That’s Left to Know about Life According to Oz. Milwaukee, WI: Applause Theatre and Cinema Books.Google Scholar
Hogg, E. H., Price, D. T., and Black, T. A. (2000). Postulated feedbacks of deciduous forest phenology on seasonal climate patterns in the western Canadian interior. Journal of Climate, 13, 42294243.Google Scholar
Hohenstein, A. (1860). Der Wald sammt dessen wichtigem Einfluss auf das Klima der Länder, Wohl der Staaten und Völker, sowie der Gesundheit der Menschen. Vienna: Carl Gerold’s Sohn.Google Scholar
Holdridge, L. R. (1967). Life Zone Ecology. San Jose, Costa Rica: Tropical Science Center.Google Scholar
Holyoke, E. A. (1793). An estimate of the excess of the heat and cold of the American atmosphere beyond the European, in the same parallel of latitude: To which are added, some thoughts on the causes of this excess. Memoirs of the American Academy of Arts and Sciences, 2(1), 6592.Google Scholar
Holzman, B. (1937). Sources of Moisture for Precipitation in the United States, Technical Bulletin No. 589. Washington, DC: U.S. Department of Agriculture.Google Scholar
Hopcroft, P. O., and Valdes, P. J. (2021). Paleoclimate-conditioning reveals a North Africa land–atmosphere tipping point. Proceedings of the National Academy of Sciences USA, 118, e2108783118, DOI: https://doi.org/10.1073/pnas.2108783118.Google Scholar
Hornbeck, J. W., Pierce, R. S., and Federer, C. A. (1970). Streamflow changes after forest clearing in New England. Water Resources Research, 6, 11241132.Google Scholar
Hough, F. B. (1874). On the duty of governments in the preservation of forests. Proceedings of the American Association for the Advancement of Science, 22B, 110.Google Scholar
Hough, F. B. (1878a). On the preservation of forests and the planting of timber. In Transactions of the New York State Agricultural Society, 1872–1876, vol. 32. [Troy, N.Y.]: Jerome B. Parmenter, pp. 177194, 293.Google Scholar
Hough, F. B. (1878b). Report upon Forestry. Washington, DC: Government Printing Office.Google Scholar
Hough, F. B. (1882). The Elements of Forestry. Cincinnati: Robert Clarke.Google Scholar
Hough, F. B. (1885). Letter from Dr. Franklin B. Hough, in regard to the effect of forests in increasing the amount of rainfall. In Report in Regard to the Range and Ranch Cattle Business of the United States, edited by Nimmo, J., Jr. Washington, DC: Government Printing Office, pp. 130131.Google Scholar
Houghton, R. A. (2005). Aboveground forest biomass and the global carbon balance. Global Change Biology, 11, 945958.Google Scholar
Houghton, R. A. (2013). Keeping management effects separate from environmental effects in terrestrial carbon accounting. Global Change Biology, 19, 26092612.Google Scholar
Hovi, A., Lindberg, E., Lang, M., et al. (2019). Seasonal dynamics of albedo across European boreal forests: Analysis of MODIS albedo and structural metrics from airborne LiDAR. Remote Sensing of Environment, 224, 365381.Google Scholar
Howard, R. A., and Howard, E. S. (1983). Alexander Anderson’s Geography and History of St. Vincent, West Indies. Cambridge, MA: Arnold Arboretum.Google Scholar
Howe, L. C., MacInnis, B., Krosnick, J. A., Markowitz, E. M., and Socolow, R. (2019). Acknowledging uncertainty impacts public acceptance of climate scientists’ predictions. Nature Climate Change, 9, 863867.Google Scholar
Hu, X., Shi, L., Lin, L., and Magliulo, V. (2020). Improving surface roughness lengths estimation using machine learning algorithms. Agricultural and Forest Meteorology, 287, 107956, DOI: https://doi.org/10.1016/j.agrformet.2020.107956.Google Scholar
Hubau, W., Lewis, S. L., Phillips, O. L., et al. (2020). Asynchronous carbon sink saturation in African and Amazonian tropical forests. Nature, 579, 8087.Google Scholar
Hudiburg, T. W., Law, B. E., Moomaw, W. R., Harmon, M. E., and Stenzel, J. E. (2019). Meeting GHG reduction targets requires accounting for all forest sector emissions. Environmental Research Letters, 14, 095005, DOI: https://doi.org/10.1088/1748-9326/ab28bb.Google Scholar
Hulme, M. (2009). On the origin of “the greenhouse effect”: John Tyndall’s 1859 interrogation of nature. Weather, 64, 121123.Google Scholar
Humboldt, A. von (1808). Ansichten der Natur mit wissenschaftlichen Erläuterungen. Tübingen: J. G. Cotta.Google Scholar
Humboldt, A. von (1817). Des lignes isothermes et de la distribution de la chaleur sur le globe. Mémoires de physique et de chimie, de la société d’Arcueil, 3, 462602.Google Scholar
Humboldt, A. von (1820–21). On isothermal lines, and the distribution of heat over the globe. Edinburgh Philosophical Journal, 3, 120, 256274; 4, 2337, 262281; 5, 2839.Google Scholar
Humboldt, A. von (1826). Ansichten der Natur, mit wissenschaftlichen Erläuterungen, 2 vols. Stuttgart & Tübingen: J. G. Cotta.Google Scholar
Humboldt, A. von (1831). Fragmens de géologie et de climatologie asiatiques, 2 vols. Paris: Gide; Pihan Delaforest; Delaunay.Google Scholar
Humboldt, A. von (1843). Asie centrale: recherches sur les chaines de montagnes et la climatologie comparée, 3 vols. Paris: Gide.Google Scholar
Humboldt, A. von (1846–58). Cosmos: Sketch of a Physical Description of the Universe, translated by E. Sabine, 4 vols. London: Longman, Brown, Green & Longmans; John Murray.Google Scholar
Humboldt, A. von (1849). Ansichten der Natur, mit wissenschaftlichen Erläuterungen, 2 vols. Stuttgart & Tübingen: J. G. Cotta.Google Scholar
Humboldt, A. von (1850). Views of Nature: Or Contemplations on the Sublime Phenomena of Creation; with Scientific Illustrations, translated by E. C. Otté and H. G. Bohn. London: Henry G. Bohn.Google Scholar
Humboldt, A. von (2009). Briefe aus Russland 1829, edited by Knobloch, E., Schwarz, I. and Suckow, C., introductory essay by Ette, O.. Berlin: Akademie Verlag.Google Scholar
Humboldt, A. von (2014). Views of Nature, translated by M. W. Person, edited by Jackson, S. T. and Walls, L. D.. Chicago: University of Chicago Press.Google Scholar
Humboldt, A. von, and Bonpland, A. (1814–29). Personal Narrative of Travels to the Equinoctial Regions of the New Continent, during the Years 1799–1804, translated by H. M. Williams, 7 vols. London: Longman, Hurst, Rees, Orme & Brown.Google Scholar
Humboldt, A. von, and Bonpland, A. (2009). Essay on the Geography of Plants, edited with an introduction by Jackson, S. T., translated by Romanowski, S.. Chicago: University of Chicago Press.Google Scholar
Hume, D. (1752). Of the populousness of antient nations. In Political Discourses. Edinburgh: R. Fleming, pp. 155261.Google Scholar
Humphrey, V., Berg, A., Ciais, P., et al. (2021). Soil moisture–atmosphere feedback dominates land carbon uptake variability. Nature, 592, 6569.Google Scholar
Humphrey, V., Zscheischler, J., Ciais, P., et al. (2018). Sensitivity of atmospheric CO2 growth rate to observed changes in terrestrial water storage. Nature, 560, 628631.Google Scholar
Hunt, A. (2016). Reviving Roman Religion: Sacred Trees in the Roman World. Cambridge, UK: Cambridge University Press.Google Scholar
Huntley, B., and Prentice, I. C. (1993). Holocene vegetation and climates of Europe. In Global Climates since the Last Glacial Maximum, edited by Wright, H. E., Jr., Kutzbach, J. E., Webb, T., III, et al. Minneapolis: University of Minnesota Press, pp. 136168.Google Scholar
Hurmekoski, E., Smyth, C. E., Stern, T., Verkerk, P. J., and Asada, R. (2021). Substitution impacts of wood use at the market level: A systematic review. Environmental Research Letters, 16, 123004, DOI: https://doi.org/10.1088/1748-9326/ac386f.Google Scholar
Hursh, C. R. (1948). Local Climate in the Copper Basin of Tennessee as Modified by the Removal of Vegetation, Circular Number 774. Washington, DC: U.S. Department of Agriculture.Google Scholar
Hurtt, G. C., Chini, L., Sahajpal, R., et al. (2020). Harmonization of global land use change and management for the period 850–2100 (LUH2) for CMIP6. Geoscientific Model Development, 13, 54255464.Google Scholar
Hutchison, B. A., Matt, D. R., McMillen, R. T., et al. (1986). The architecture of a deciduous forest canopy in eastern Tennessee, U.S.A. Journal of Ecology, 74, 635646.Google Scholar
Huuskonen, S., Domisch, T., Finér, L., et al. (2021). What is the potential for replacing monocultures with mixed-species stands to enhance ecosystem services in boreal forests in Fennoscandia? Forest Ecology and Management, 479, 118558, DOI: https://doi.org/10.1016/j.foreco.2020.118558.Google Scholar
Iglesias, V., Balch, J. K., and Travis, W. R. (2022). U.S. fires became larger, more frequent, and more widespread in the 2000s. Science Advances, 8, eabc0020, DOI: https://doi.org/10.1126/sciadv.abc0020.Google Scholar
Imlay, G. (1793). A Description of the Western Territory of North America. Dublin: William Jones.Google Scholar
Irvine, P. J., Ridgwell, A., and Lunt, D. J. (2011). Climatic effects of surface albedo geoengineering. Journal of Geophysical Research, 116, D24112, DOI: https://doi.org/10.1029/2011JD016281.Google Scholar
Irving, W. (1848). The Sketchbook of Geoffrey Crayon, Gent. – The Author’s Revised Edition. New York: George P. Putnam.Google Scholar
Jackson, R. (2020). Eunice Foote, John Tyndall and a question of priority. Notes and Records, 74, 105118.Google Scholar
Jackson, R. B., Jobbágy, E. G., Avissar, R., et al. (2005). Trading water for carbon with biological carbon sequestration. Science, 310, 19441947.Google Scholar
Jackson, R. B., Randerson, J. T., Canadell, J. G., et al. (2008). Protecting climate with forests. Environmental Research Letters, 3, 044006, DOI: https://doi.org/10.1088/1748–9326/3/4/044006.Google Scholar
Jahren, A. H. (2007). The Arctic forest of the middle Eocene. Annual Review of Earth and Planetary Sciences, 35, 509540.Google Scholar
James, N. D. G. (1981). A History of English Forestry. Oxford: Basil Blackwell.Google Scholar
James, N. D. G. (1996). A history of forestry and monographic forestry literature in Germany, France, and the United Kingdom. In The Literature of Forestry and Agroforestry, edited by McDonald, P. and Lassoie, J.. Ithaca, NY: Cornell University Press, pp. 1544.Google Scholar
Jamieson, T. F. (1860). The Tweeddale Prize Essay on the Rainfall. Edinburgh: William Blackwood.Google Scholar
Janisch, H. R. (1908). Extracts from the St. Helena Records (Second Edition) and Chronicles of Cape Commanders. Jamestown, St. Helena: Benjamin Grant.Google Scholar
Jefferson, T. (1787). Notes on the State of Virginia. London: John Stockdale.Google Scholar
Jefferson, T. (1903). To Lewis E. Beck, July 16, 1824. In The Writings of Thomas Jefferson, vol. 16, edited by Bergh, A. E.. Washington, DC: Thomas Jefferson Memorial Association of the United States, pp. 7172.Google Scholar
Jefferson, T. (1954). To Jean Baptiste Le Roy, November 13, 1786. In The Papers of Thomas Jefferson, vol. 10. 22 June to 31 December 1786, edited by Boyd, J. P.. Princeton, NJ: Princeton University Press, pp. 524530.Google Scholar
Jefferson, T. (1955). To Buffon, October 1, 1787. In The Papers of Thomas Jefferson, vol. 12. 7 August 1787 to 31 March 1788, edited by Boyd, J. P.. Princeton, NJ: Princeton University Press, pp. 194195.Google Scholar
Jefferson, T. (2006). To Nathaniel Chapman, December 11, 1809. In The Papers of Thomas Jefferson, Retirement Series, vol. 2. 16 November 1809 to 11 August 1810, edited by Looney, J. J.. Princeton, NJ: Princeton University Press, pp. 7072.Google Scholar
Jia, G., Shevliakova, E., Artaxo, P., et al. (2019). Land–climate interactions. In Climate Change and Land: An IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems, edited by Shukla, P. R., Skea, J., Buendia, E. Calvo, et al. Geneva: World Meteorological Organization, pp. 131247.Google Scholar
Jiang, M., Medlyn, B. E., Drake, J. E., et al. (2020). The fate of carbon in a mature forest under carbon dioxide enrichment. Nature, 580, 227231.Google Scholar
Jiang, Y., Wang, G., Liu, W., et al. (2021). Modeled response of South American climate to three decades of deforestation. Journal of Climate, 34, 21892203.Google Scholar
Johnson, E. (1910). Johnson’s Wonder-Working Providence, 1628–1651, edited by Jameson, J. F.. New York: Charles Scribner’s Sons.Google Scholar
Joussaume, S., Taylor, K. E., Braconnot, P., et al. (1999). Monsoon changes for 6000 years ago: Results of 18 simulations from the Paleoclimate Modeling Intercomparison Project (PMIP). Geophysical Research Letters, 26, 859862.Google Scholar
Jouzel, J., Masson-Delmotte, V., Cattani, O., et al. (2007). Orbital and millennial Antarctic climate variability over the past 800,000 years. Science, 317, 793796.Google Scholar
Juang, J.-Y., Katul, G., Siqueira, M., Stoy, P., and Novick, K. (2007). Separating the effects of albedo from eco-physiological changes on surface temperature along a successional chronosequence in the southeastern United States. Geophysical Research Letters, 34, L21408, DOI: https://doi.org/10.1029/2007GL031296.Google Scholar
Kalliokoski, T., Bäck, J., Boy, M., et al. (2020). Mitigation impact of different harvest scenarios of Finnish forests that account for albedo, aerosols, and trade-offs of carbon sequestration and avoided emissions. Frontiers in Forests and Global Change, 3, 562044, DOI: https://doi.org/10.3389/ffgc.2020.562044.Google Scholar
Kalm, P. (1770–71). Travels into North America, translated by J. R. Forster, 3 vols. Warrington: William Eyres; London: T. Lowndes.Google Scholar
Kalm, P. (1937). The America of 1750: Peter Kalm’s Travels in North America; The English Version of 1770, edited and translated by A. B. Benson, 2 vols. New York: Wilson-Erickson.Google Scholar
Kasahara, A., and Washington, W. M. (1967). NCAR global general circulation model of the atmosphere. Monthly Weather Review, 95, 389402.Google Scholar
Kasahara, A., and Washington, W. M. (1971). General circulation experiments with a six-layer NCAR model, including orography, cloudiness and surface temperature calculations. Journal of the Atmospheric Sciences, 28, 657701.Google Scholar
Kedzie, R. C. (1867). The influence of forest trees on agriculture. In Sixth Annual Report of the Secretary of the State Board of Agriculture of the State of Michigan, for the Year 1867. Lansing: John A. Kerr, pp. 465483.Google Scholar
Kedzie, R. C., Woodman, J. J., and Fellows, O. H. (1866). Report of the committee. In Fifth Annual Report of the Secretary of the State Board of Agriculture of the State of Michigan, for the Year 1866. Lansing: John A. Kerr, Appendix pp. 331.Google Scholar
Keeling, R. F., Piper, S. C. and Heimann, M. (1996). Global and hemispheric CO2 sinks deduced from changes in atmospheric O2 concentration. Nature, 381, 218221.Google Scholar
Keenan, T. F., and Williams, C. A. (2018). The terrestrial carbon sink. Annual Review of Environment and Resources, 43, 219 –243.Google Scholar
Keenan, T. F., Luo, X., De Kauwe, M. G., et al. (2021). A constraint on historic growth in global photosynthesis due to increasing CO2. Nature, 600, 253258.Google Scholar
Keever, C. (1950). Causes of succession on old fields of the Piedmont, North Carolina. Ecological Monographs, 20, 229250.Google Scholar
Keever, C. (1983). A retrospective view of old-field succession after 35 years. American Midland Naturalist, 110, 397404.Google Scholar
Keith, H., Vardon, M., Obst, C., et al. (2021). Evaluating nature-based solutions for climate mitigation and conservation requires comprehensive carbon accounting. Science of the Total Environment, 769, 144341, DOI: https://doi.org/10.1016/j.scitotenv.2020.144341.Google Scholar
Kellomäki, S., Väisänen, H., Kirschbaum, M. U. F., Kirsikka-Aho, S., and Peltola, H. (2021). Effects of different management options of Norway spruce on radiative forcing through changes in carbon stocks and albedo. Forestry, 94, 588597.Google Scholar
Kemena, T. P., Matthes, K., Martin, T., Wahl, S., and Oschlies, A. (2018). Atmospheric feedbacks in North Africa from an irrigated, afforested Sahara. Climate Dynamics, 50, 45614581.Google Scholar
Kemp, M. (2008). Looking at the face of the Earth. Nature, 456, 876.Google Scholar
Kerner von Marilaun, A. (1888). Pflanzenleben, vol. 1. Gestalt und Leben der Pflanze. Leipzig: Bibliographischen Instituts.Google Scholar
Kerner von Marilaun, A. (1894). The Natural History of Plants: Their Forms, Growth, Reproduction, and Distribution, vol. 1. Biology and Configuration of Plants, translated by F. W. Oliver. London: Blackie.Google Scholar
Khanna, J., Medvigy, D., Fueglistaler, S., and Walko, R. (2017). Regional dry-season climate changes due to three decades of Amazonian deforestation. Nature Climate Change, 7, 200204.Google Scholar
Kittredge, J. (1948). Forest Influences: The Effects of Woody Vegetation on Climate, Water, and Soil, with Applications to the Conservation of Water and the Control of Floods and Erosion. New York: McGraw-Hill.Google Scholar
Klages, J. P., Salzmann, U., Bickert, T., et al. (2020). Temperate rainforests near the South Pole during peak Cretaceous warmth. Nature, 580, 8186.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. Climatic Change, 44, 471493.Google Scholar
Klein, R. G. (1995). Anatomy, behavior, and modern human origins. Journal of World Prehistory, 9, 167198.Google Scholar
Kooperman, G. J., Chen, Y., Hoffman, F. M., et al. (2018). Forest response to rising CO2 drives zonally asymmetric rainfall change over tropical land. Nature Climate Change, 8, 434440.Google Scholar
Kopenawa, D., and Albert, B. (2013). The Falling Sky: Words of a Yanomami Shaman, translated by N. Elliott and A. Dundy. Cambridge, MA: Harvard University Press.Google Scholar
Koren, I., Kaufman, Y. J., Remer, L. A., and Martins, J. V. (2004). Measurement of the effect of Amazon smoke on inhibition of cloud formation. Science, 303, 13421345.Google Scholar
Koren, I., Martins, J. V., Remer, L. A., and Afargan, H. (2008). Smoke invigoration versus inhibition of clouds over the Amazon. Science, 321, 946949.Google Scholar
Koster, R. D., Dirmeyer, P. A., Guo, Z., et al. (2004). Regions of strong coupling between soil moisture and precipitation. Science, 305, 11381140.Google Scholar
Kotok, E. I. (1940). The forester’s dependence on the science of meteorology. Bulletin of the American Meteorological Society, 21, 383384, 397406.Google Scholar
Kreidenweis, U., Humpenöder, F., Stevanović, M., et al. (2016). Afforestation to mitigate climate change: Impacts on food prices under consideration of albedo effects. Environmental Research Letters, 11, 085001, DOI: https://doi.org/10.1088/1748-9326/11/8/085001.Google Scholar
Kröniger, K., De Roo, F., Brugger, P., et al. (2018). Effect of secondary circulations on the surface–atmosphere exchange of energy at an isolated semi-arid forest. Boundary-Layer Meteorology, 169, 209232.Google Scholar
Kucharski, F., Zeng, N., and Kalnay, E. (2013). A further assessment of vegetation feedback on decadal Sahel rainfall variability. Climate Dynamics, 40, 14531466.Google Scholar
Kulmala, M., Ezhova, E., Kalliokoski, T., et al. (2020). CarbonSink+: Accounting for multiple climate feedbacks from forests. Boreal Environmental Research, 25, 145159.Google Scholar
Kulmala, M., Nieminen, T., Nikandrova, A., et al. (2014). CO2-induced terrestrial climate feedback mechanism: From carbon sink to aerosol source and back. Boreal Environment Research, 19 (suppl. B), 122131.Google Scholar
Kulmala, M., Suni, T., Lehtinen, K. E. J., et al. (2004). A new feedback mechanism linking forests, aerosols, and climate. Atmospheric Chemistry and Physics, 4, 557562.Google Scholar
Kupperman, K. O. (1982). The puzzle of the American climate in the early colonial period. American Historical Review, 87, 12621289.Google Scholar
Kurtén, T., Kulmala, M., Dal Maso, M., et al. (2003). Estimation of different forest-related contributions to the radiative balance using observations in southern Finland. Boreal Environment Research, 8, 275285.Google Scholar
Kurz, W. A., Dymond, C. C., Stinson, G., et al. (2008). Mountain pine beetle and forest carbon feedback to climate change. Nature, 452, 987990.Google Scholar
Kutzbach, J., Bonan, G., Foley, J., and Harrison, S. P. (1996). Vegetation and soil feedbacks on the response of the African monsoon to orbital forcing in the early to middle Holocene. Nature, 384, 623626.Google Scholar
Kutzleb, C. R. (1971). Can forests bring rain to the Plains? Forest History, 15(3), 1421.Google Scholar
Kuusinen, N., Tomppo, E., Shuai, Y., and Berninger, F. (2014). Effects of forest age on albedo in boreal forests estimated from MODIS and Landsat albedo retrievals. Remote Sensing of Environment, 145, 145153.Google Scholar
Laguë, M. M., Bonan, G. B., and Swann, A. L. S. (2019). Separating the impact of individual land surface properties on the terrestrial surface energy budget in both the coupled and uncoupled land-atmosphere system. Journal of Climate, 32, 57255744.Google Scholar
Laguë, M. M., Pietschnig, M., Ragen, S., Smith, T. A., and Battisti, D. S. (2021a). Terrestrial evaporation and global climate: Lessons from northland, a planet with a hemispheric continent. Journal of Climate, 34, 22532276.Google Scholar
Laguë, M. M., and Swann, A. L. S. (2016). Progressive midlatitude afforestation: Impacts on clouds, global energy transport, and precipitation. Journal of Climate, 29, 55615573.Google Scholar
Laguë, M. M., Swann, A. L. S., and Boos, W. R. (2021b). Radiative feedbacks on land surface change and associated tropical precipitation shifts. Journal of Climate, 34, 66516672.Google Scholar
Lamarck, J.-B. (1794). Recherches sur les causes des principaux faits physiques, 2 vols. Paris: Maradan.Google Scholar
Lamarck, J.-B. (1801/02). Annuaire météorologique, pour l’an X de l’ère de la République française, vol. 3. Paris: Maillard.Google Scholar
Lamarck, J.-B. (1802). Hydrogéologie. Paris: Agasse; Maillard.Google Scholar
Lamarck, J.-B. (1820). Système analytique des connaissances positives de l’homme. Paris: A. Belin.Google Scholar
Lamarck, J.-B. (1964). Hydrogeology, translated by Carozzi, A. V.. Urbana, IL: University of Illinois Press.Google Scholar
Lamb, H. H. (1977). Climate: Present, Past and Future, vol. 2. Climatic History and the Future. London: Methuen.Google Scholar
Lamb, H. H. (1995). Climate, History and the Modern World, 2nd ed. London: Routledge.Google Scholar
Lange, M. (2005). Ecological laws: What would they be and why would they matter? Oikos, 110, 394403.Google Scholar
Lapham, I. A., Knapp, J. G., and Crocker, H. (1867). Report on the Disastrous Effects of the Destruction of Forest Trees, Now Going on So Rapidly in the State of Wisconsin. Madison: Atwood & Rublee.Google Scholar
Larsen, J. A. (1980). The Boreal Ecosystem. New York: Academic Press.Google Scholar
Lathière, J., Hewitt, C. N., and Beerling, D. J. (2010). Sensitivity of isoprene emissions from the terrestrial biosphere to 20th century changes in atmospheric CO2 concentration, climate, and land use. Global Biogeochemical Cycles, 24, GB1004, DOI: https://doi.org/10.1029/2009GB003548.Google Scholar
Lawrence, D., and Vandecar, K. (2015). Effects of tropical deforestation on climate and agriculture. Nature Climate Change, 5, 2736.Google Scholar
Lawton, J. H. (1999). Are there general laws in ecology? Oikos, 84, 177192.Google Scholar
Lazarus, M. H., and Pardoe, H. S. (2003). Catalogue of Botanical Prints and Drawings at the National Museums & Galleries of Wales. Cardiff: National Museums & Galleries of Wales.Google Scholar
Lecoy, A. (1879). The forest question in New Zealand. Transactions and Proceedings of the New Zealand Institute, 12, 323.Google Scholar
Lee, D. (1850). Agricultural meteorology. In Report of the Commissioner of Patents, for the Year 1849. Part. II. Agriculture. Washington, DC: Office of Printers to the Senate, pp. 3848.Google Scholar
Lee, X., Goulden, M. L., Hollinger, D. Y., et al. (2011). Observed increase in local cooling effect of deforestation at higher latitudes. Nature, 479, 384387.Google Scholar
Legg, S. (2014). Debating the climatological role of forests in Australia, 1827–1949: A survey of the popular press. In Climate, Science, and Colonization: Histories from Australia and New Zealand, edited by Beattie, J., O’Gorman, E., and Henry, M.. New York: Palgrave Macmillan, pp. 119136.Google Scholar
Legg, S. M. (2018). Views from the Antipodes: The “forest influence” debate in the Australian and New Zealand press, 1827–1956. Australian Geographer, 49, 4160.Google Scholar
Le Jeune, P. (1634). Relation de ce qui s’est passe en la Nouvelle France en l’annee 1633. Paris: Sebastien Cramoisy.Google Scholar
Lejeune, Q., Davin, E. L., Gudmundsson, L., Winckler, J., and Seneviratne, S. I. (2018). Historical deforestation locally increased the intensity of hot days in northern mid-latitudes. Nature Climate Change, 8, 386390.Google Scholar
Lejeune, Q., Seneviratne, S. I., and Davin, E. L. (2017). Historical land-cover change impacts on climate: Comparative assessment of LUCID and CMIP5 multimodel experiments. Journal of Climate, 30, 14391459.Google Scholar
Lembke, J. (2005). Virgil’s Georgics: A New Verse Translation. New Haven, CT: Yale University Press.Google Scholar
Lemordant, L., Gentine, P., Swann, A. S., Cook, B. I., and Scheff, J. (2018). Critical impact of vegetation physiology on the continental hydrologic cycle in response to increasing CO2. Proceedings of the National Academy of Sciences USA, 115, 40934098.Google Scholar
Lenton, T. M., Held, H., Kriegler, E., et al. (2008). Tipping elements in the Earth’s climate system. Proceedings of the National Academy of Sciences USA, 105, 17861793.Google Scholar
Lenton, T. M., Rockström, J., Gaffney, O., et al. (2019). Climate tipping points – too risky to bet against. Nature, 575, 592595.Google Scholar
Leonardi, C., and Stagi, F. (2019). The Architecture of Trees. Hudson, NY: Princeton Architectural Press.Google Scholar
Le Roy, C.-G. (1757). Forêt. In Encyclopédie, ou dictionnaire raisonné des sciences, des arts et des métiers, vol. 7, edited by Diderot, D. and Le Rond, J. d’Alembert. Paris: Briasson; David; Le Breton; Durand, pp. 129132.Google Scholar
Le Roy, J. B. (1954). From Jean Baptiste Le Roy, September 28, 1786. In The Papers of Thomas Jefferson, vol. 10. 22 June to 31 December 1786, edited by Boyd, J. P.. Princeton: Princeton University Press, pp. 410411.Google Scholar
Lescarbot, M. (1609a). Histoire de la Nouvelle France. Paris: Jean Milot.Google Scholar
Lescarbot, M. (1609b). Nova Francia: Or the Description of That Part of New France, Which is One Continent with Virginia, translated by P. E. London: George Bishop.Google Scholar
Leslie, A. B., Beaulieu, J., Holman, G., et al. (2018). An overview of extant conifer evolution from the perspective of the fossil record. American Journal of Botany, 105, 15311544.Google Scholar
Leslie, J. (1819). On heat and climate. Annals of Philosophy, 14, 527.Google Scholar
Leslie, J. (1804). An Experimental Inquiry into the Nature, and Propagation, of Heat. London: J. Mawman.Google Scholar
Leuzinger, S., and Körner, C. (2007). Tree species diversity affects canopy leaf temperatures in a mature temperate forest. Agricultural and Forest Meteorology, 146, 2937.Google Scholar
Levis, S., and Bonan, G. B. (2004). Simulating springtime temperature patterns in the Community Atmosphere Model coupled to the Community Land Model using prognostic leaf area. Journal of Climate, 17, 45314540.Google Scholar
Levis, S., Bonan, G. B., and Bonfils, C. (2004). Soil feedback drives the mid-Holocene North African monsoon northward in fully coupled CCSM2 simulations with a dynamic vegetation model. Climate Dynamics, 23, 791802.Google Scholar
Levis, S., Foley, J. A., and Pollard, D. (1999). CO2, climate, and vegetation feedbacks at the Last Glacial Maximum. Journal of Geophysical Research, 104D, 3119131198.Google Scholar
Lewis, S. L., Brando, P. M., Phillips, O. L., van der Heijden, G. M. F., and Nepstad, D. (2011). The 2010 Amazon drought. Science, 331, 554.Google Scholar
Lewis, S. L., Wheeler, C. E., Mitchard, E. T. A., and Koch, A. (2019). Regenerate natural forests to store carbon. Nature, 568, 2528.Google Scholar
Li, Q., Wei, X., Zhang, M., et al. (2017). Forest cover change and water yield in large forested watersheds: A global synthetic assessment. Ecohydrology, 10, e1838, DOI: https://doi.org/10.1002/eco.1838.Google Scholar
Li, Y., Brando, P. M., Morton, D. C., et al. (2022a). Deforestation-induced climate change reduces carbon storage in remaining tropical forests. Nature Communications, 13, 1964, DOI: https://doi.org/10.1038/s41467-022-29601-0.Google Scholar
Li, Y., de Noblet-Ducoudré, N., Davin, E. L., et al. (2016). The role of spatial scale and background climate in the latitudinal temperature response to deforestation. Earth System Dynamics, 7, 167181.Google Scholar
Li, Y., Kalnay, E., Motesharrei, S., et al. (2018a). Climate model shows large-scale wind and solar farms in the Sahara increase rain and vegetation. Science, 361, 10191022.Google Scholar
Li, Y., Liu, Y., Bohrer, G., et al. (2022b). Impacts of forest loss on local climate across the conterminous United States: Evidence from satellite time-series observations. Science of The Total Environment, 802, 149651, DOI: https://doi.org/10.1016/j.scitotenv.2021.149651.Google Scholar
Li, Y., Piao, S., Li, L. Z. X., et al. (2018b). Divergent hydrological response to large-scale afforestation and vegetation greening in China. Science Advances, 4, eaar4182, DOI: https://doi.org/10.1126/sciadv.aar4182.Google Scholar
Li, Y., Piao, S., Chen, A., Ciais, P., and Li, L. Z. X. (2020). Local and teleconnected temperature effects of afforestation and vegetation greening in China. National Science Review, 7, 897912.Google Scholar
Li, Y., Randerson, J. T., Mahowald, N. M., and Lawrence, P. J. (2021). Deforestation strengthens atmospheric transport of mineral dust and phosphorus from North Africa to the Amazon. Journal of Climate, 34, 60876096.Google Scholar
Li, Y., Zhao, M., Motesharrei, S., et al. (2015). Local cooling and warming effects of forests based on satellite observations. Nature Communications, 6, 6603, DOI: https://doi.org/10.1038/ncomms7603.Google Scholar
Liebig, J. von (1862). Die Chemie in ihrer Anwendung auf Agricultur und Physiologie, 7th ed., 2 vols. Braunschweig: Friedrich Vieweg.Google Scholar
Lihavainen, H., Kerminen, V.-M., Tunved, P., et al. (2009). Observational signature of the direct radiative effect by natural boreal forest aerosols and its relation to the corresponding first indirect effect. Journal of Geophysical Research, 114, D20206, DOI: https://doi.org/10.1029/2009JD012078.Google Scholar
Lihavainen, H., Asmi, E., Aaltonen, V., Makkonen, U., and Kerminen, V.-M. (2015). Direct radiative feedback due to biogenic secondary organic aerosol estimated from boreal forest site observations. Environmental Research Letters, 10, 104005, DOI: https://doi.org/10.1088/1748-9326/10/10/104005.Google Scholar
Likens, G. E., Bormann, F. H., Pierce, R. S., Eaton, J. S., and Johnson, N. M. (1977). Biogeochemistry of a Forested Ecosystem. New York: Springer-Verlag.Google Scholar
Linquist, S., Gregory, T. R., Elliott, T. A., et al. (2016). Yes! There are resilient generalizations (or “laws”) in ecology. Quarterly Review of Biology, 91, 119131.Google Scholar
Lintunen, J., Rautiainen, A., and Uusivuori, J. (2022). Which is more important, carbon or albedo? Optimizing harvest rotations for timber and climate benefits in a changing climate. American Journal of Agricultural Economics, 104, 134160.Google Scholar
Liu, H., and Randerson, J. T. (2008). Interannual variability of surface energy exchange depends on stand age in a boreal forest fire chronosequence. Journal of Geophysical Research, 113, G01006, DOI: https://doi.org/10.1029/2007JG000483.Google Scholar
Liu, H., Randerson, J. T., Lindfors, J., and Chapin, F. S., III (2005). Changes in the surface energy budget after fire in boreal ecosystems of interior Alaska: An annual perspective. Journal of Geophysical Research, 110, D13101, DOI: https://doi.org/10.1029/2004JD005158.Google Scholar
Liu, J., Li, S., Ouyang, Z., Tam, C., and Chen, X. (2008). Ecological and socioeconomic effects of China’s policies for ecosystem services. Proceedings of the National Academy of Sciences USA, 105, 94779482.Google Scholar
Liu, L., Cheng, Y., Wang, S., et al. (2020). Impact of biomass burning aerosols on radiation, clouds, and precipitation over the Amazon: Relative importance of aerosol–cloud and aerosol–radiation interactions. Atmospheric Chemistry and Physics, 20, 1328313301.Google Scholar
Locher, F., and Fressoz, J.-B. (2012). Modernity’s frail climate: A climate history of environmental reflexivity. Critical Inquiry, 38, 579598.Google Scholar
Löffelholz-Colberg, F. F. von (1872). Die Bedeutung und Wichtigkeit des Waldes, Ursachen und Folgen der Entwaldung, die Wiederbewaldung mit Rücksicht auf Pflanzenphysiologie, Klimatologie, Meteorologie, Forststatistik, Forstgeographie und die forstlichen Verhältnisse aller Länder…. Leipzig: Heinrich Schmidt.Google Scholar
Lombardozzi, D., Levis, S., Bonan, G., Hess, P. G., and Sparks, J. P. (2015). The influence of chronic ozone exposure on global carbon and water cycles. Journal of Climate, 28, 292305.Google Scholar
Long, E. (1774). The History of Jamaica: or, General Survey of the Antient and Modern State of That Island, 3 vols. London: T. Lowndes.Google Scholar
Longfellow, H. W. (1863). Tales of a Wayside Inn. Boston: Ticknor & Fields.Google Scholar
Loomis, E. (1868). A Treatise on Meteorology, with a Collection of Meteorological Tables. New York: Harper.Google Scholar
Lorentz, B., and Parade, A. (1837). Cours élémentaire de culture des bois, créé a l’école royale forestière de Nancy. Paris: Huzard; Nancy: George-Grimblot.Google Scholar
Lorentz, B., and Parade, A. (1883). Cours élémentaire de culture des bois, créé a l’école forestière de Nancy, 6th ed. Paris: Octave Doin.Google Scholar
Lorenz, E. N. (1970). Climatic change as a mathematical problem. Journal of Applied Meteorology, 9, 325329.2.0.CO;2>CrossRefGoogle Scholar
Lorenz, J. R. (1877). Über Bedeutung und Vertretung der land- und forstwirthschaftlichen Meteorologie. Vienna: Faesy & Frick.Google Scholar
Lorenz, J. R., and Rothe, C. (1874). Lehrbuch der Klimatologie mit besonderer Rücksicht auf Land- und Forstwirthschaft. Vienna: Wilhelm Braumüller.Google Scholar
Lorenz, R., Pitman, A. J., and Sisson, S. A. (2016). Does Amazonian deforestation cause global effects; can we be sure? Journal of Geophysical Research: Atmospheres, 121, 55675584.Google Scholar
Lorenz von Liburnau, J. R. (1878). Wald, Klima und Wasser. Munich: R. Oldenbourg.Google Scholar
Lorenz von Liburnau, J. R. (1879). Bericht für den zweiten internationalen Meteorologen-Congress über die Frage: Wie können die meteorologischen Institute sich der Land- und Forstwirthschaft förderlich erweisen? Vienna: Carl Fromme.Google Scholar
Lorenz von Liburnau, J. R. (1890). Resultate Forstlich-Meteorologischer Beobachtungen insbesondere in den Jahren 1885–1887, part 1: Untersuchungen über die Temperatur und die Feuchtigkeit der Luft unter, in und über den Baumkronen des Waldes, sowie im Freilande, Mittheilungen vom forstlichen Versuchswesen in Österreich. XII Heft. Vienna: K. & K. Hof-Buchhandlung W. Frick.Google Scholar
Lorey, T. (1888). Handbuch der Forstwissenschaft: forstliche Produktionslehre. I. Tübingen: H. Laupp’schen.Google Scholar
Loskutova, M. (2020). Quantifying scarcity: Deforestation in the Upper Volga region and early debates over climate change in nineteenth-century Russia. European Review of History: Revue européenne d’histoire, 27, 253272.Google Scholar
Loudon, J. C. (1838). Arboretum et Fruticetum Britannicum; or, The Trees and Shrubs of Britain, 8 vols. London: Longman, Orme, Brown, Green, & Longmans.Google Scholar
Lovejoy, T. E., and Nobre, C. (2018). Amazon tipping point. Science Advances, 4, eaat2340, DOI: https://doi.org/10.1126/sciadv.aat2340.CrossRefGoogle ScholarPubMed
Lovell, J. (1826). Meteorological Register for the Years 1822, 1823, 1824 & 1825, from Observations Made by the Surgeons of the Army, at the Military Posts of the United States. Washington, DC: Edward de Krafft.Google Scholar
Lovelock, J. E. (1979). Gaia: A New Look at Life on Earth. Oxford: Oxford University Press.Google Scholar
Lovelock, J. E., and Margulis, L. (1974). Atmospheric homeostasis by and for the biosphere: The gaia hypothesis. Tellus, 26, 210.Google Scholar
Lowell, P. (1906). Mars and its Canals. New York: Macmillan.Google Scholar
Lowell, P. (1908). Mars as the Abode of Life. New York: Macmillan.Google Scholar
Lu, Z., Zhang, Q., Miller, P. A., et al. (2021). Impacts of large-scale Sahara solar farms on global climate and vegetation cover. Geophysical Research Letters, 48, e2020GL090789, DOI: https://doi.org/10.1029/2020GL090789.Google Scholar
Lusardi, J. P. (1973). Appendix A, Barnes’ supplications of 1531 and 1534. In The Yale Edition of the Complete Works of St. Thomas More, vol. 8. The Confutation of Tyndale’s Answer, Part II The Text, Books V–IX, Appendices, edited by Schuster, L. A., Marius, R. C., Lusardi, J. P., and Schoeck, R. J.. New Haven, CT: Yale University Press, pp. 10351061.Google Scholar
Lüthi, D., Le Floch, M., Bereiter, B., et al. (2008). High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature, 453, 379382.Google Scholar
Lutz, D. A., and Howarth, R. B. (2014). Valuing albedo as an ecosystem service: Implications for forest management. Climatic Change, 124, 5363.Google Scholar
Lutz, D. A., Burakowski, E. A., Murphy, M. B., et al. (2016). Trade-offs between three forest ecosystem services across the state of New Hampshire, USA: Timber, carbon, and albedo. Ecological Applications, 26, 146161.Google Scholar
Luyssaert, S., Jammet, M., Stoy, P. C., et al. (2014). Land management and land-cover change have impacts of similar magnitude on surface temperature. Nature Climate Change, 4, 389393.Google Scholar
Luyssaert, S., Marie, G., Valade, A., et al. (2018). Trade-offs in using European forests to meet climate objectives. Nature, 562, 259262.Google Scholar
Lyell, C. (1834). Principles of Geology, 3rd ed., 4 vols. London: John Murray.Google Scholar
Lynch, P. (2008). The origins of computer weather prediction and climate modeling. Journal of Computational Physics, 227, 34313444.Google Scholar
Mabey, R. (2016). The Cabaret of Plants: Forty Thousand Years of Plant Life and the Human Imagination. New York: W. W. Norton.Google Scholar
Mahmood, R., Pielke, R. A., Sr., Hubbard, K. G., et al. (2014). Land cover changes and their biogeophysical effects on climate. International Journal of Climatology, 34, 929953.Google Scholar
Mahrt, L., and Ek, M. (1993). Spatial variability of turbulent fluxes and roughness lengths in HAPEX-MOBILHY. Boundary-Layer Meteorology, 65, 381400.Google Scholar
Makar, P. A., Akingunola, A., Chen, J., et al. (2021). Forest-fire aerosol–weather feedbacks over western North America using a high-resolution, online coupled air-quality model. Atmospheric Chemistry and Physics, 21, 1055710587.Google Scholar
Malhi, Y. (2017). The concept of the Anthropocene. Annual Review of Environment and Resources, 42, 77104.Google Scholar
Malmer, A., Murdiyarso, D., Bruijnzeel, L. A., and Ilstedt, U. (2010). Carbon sequestration in tropical forests and water: A critical look at the basis for commonly used generalizations. Global Change Biology, 15, 599604.Google Scholar
Malone, J. J. (1964). Pine Trees and Politics: The Naval Stores and Forest Policy in Colonial New England, 1691–1775. Seattle: University of Washington Press.Google Scholar
Malte-Brun, C. (1834). A System of Universal Geography, or a Description of all the Parts of the World, with additions and corrections by Percival, J. G., 3 vols. Boston: Samuel Walker.Google Scholar
Manabe, S. (1969). Climate and the ocean circulation. I. The atmospheric circulation and the hydrology of the Earth’s surface. Monthly Weather Review, 97, 739774.Google Scholar
Manabe, S., Smagorinsky, J., and Strickler, R. F. (1965). Simulated climatology of a general circulation model with a hydrologic cycle. Monthly Weather Review, 93, 769798.Google Scholar
Maness, H., Kushner, P. J., and Fung, I. (2013). Summertime climate response to mountain pine beetle disturbance in British Columbia. Nature Geoscience, 6, 6570.Google Scholar
Mann, C. R., and Twiss, G. R. (1910). Physics, rev. ed. Chicago: Scott, Foresman & Company.Google ScholarPubMed
Manning, A. C., and Keeling, R. F. (2006). Global oceanic and land biotic carbon sinks from the Scripps atmospheric oxygen flask sampling network. Tellus, 58B, 95116.Google Scholar
Manwood, J. (1598). A Treatise and Discourse of the Lawes of the Forrest. London: Thomas Wight & Bonham Norton.Google Scholar
Marengo, J. A., Souza, C. M., Jr., Thonicke, K., et al. (2018). Changes in climate and land use over the Amazon region: Current and future variability and trends. Frontiers in Earth Science, 6, 228, DOI: https://doi.org/10.3389/feart.2018.00228.Google Scholar
Marks, P. L. (1974). The role of pin cherry (Prunus pensylvanica L.) in the maintenance of stability in northern hardwood ecosystems. Ecological Monographs, 44, 7388.Google Scholar
Marsh, G. P. (1864). Man and Nature; or, Physical Geography as Modified by Human Action. New York: Charles Scribner.Google Scholar
Martin, S. T., Andreae, M. O., Artaxo, P., et al. (2010). Sources and properties of Amazonian aerosol particles. Reviews of Geophysics, 48, RG2002, DOI: https://doi.org/10.1029/2008RG000280.Google Scholar
Massman, W. J. (1997). An analytical one-dimensional model of momentum transfer by vegetation of arbitrary structure. Boundary-Layer Meteorology, 83, 407421.Google Scholar
Mather, C. (1721). The Christian Philosopher: A Collection of the Best Discoveries in Nature, with Religious Improvements. London: Eman. Matthews.Google Scholar
Mathieu, A. (1855). Description des bois des essences forestières les plus importantes. Nancy: Grimbolt & Veuve Raybois.Google Scholar
Mathieu, A. (1877). Flore forestière: description et histoire des végétaux ligneux qui croissant spontanément en France et des essences importantes de l’Algérie, 3rd ed. Paris: Berger-Levrault & Cie; Vueve Bouchard-Huzard.Google Scholar
Mathieu, A. (1878). Météorologie comparée agricole et forestière: Rapport à M. le sous-secrétaire d’État, président du conseil d’administration des forêts. Exposition universelle de 1878. Ministère de l’Agriculture et du commerce. Administration des forêts. Paris: Imprimerie Nationale.Google Scholar
Matsuda, M., Tadaki, Y., Izuhara, S., Takumi, A., and Ohshima, Y. (1987). Seasonal variations of the physical environment of larch forest. Journal of Agricultural Meteorology, 43, 313.Google Scholar
Matteson, K. (2015). Forests in Revolutionary France: Conservation, Community, and Conflict, 1669–1848. New York: Cambridge University Press.Google Scholar
Matthews, H. D., Weaver, A. J., Meissner, K. J., Gillett, N. P., and Eby, M. (2004). Natural and anthropogenic climate change: Incorporating historical land cover change, vegetation dynamics and the global carbon cycle. Climate Dynamics, 22, 461479.Google Scholar
Matthews, W. H., Kellogg, W. W., and Robinson, G. D. (1971). Man’s Impact on the Climate. Cambridge, MA: Massachusetts Institute of Technology Press.Google Scholar
Matthies, B. D., and Valsta, L. T. (2016). Optimal forest species mixture with carbon storage and albedo effect for climate change mitigation. Ecological Economics, 123, 95105.Google Scholar
Mayr, E. (1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Cambridge, MA: Harvard University Press.Google Scholar
McCurry, M. R., Cantrill, D. J., Smith, P. M., et al. (2022). A Lagerstätte from Australia provides insight into the nature of Miocene mesic ecosystems. Science Advances, 8, eabm1406, DOI: https://doi.org/10.1126/sciadv.abm1406.Google Scholar
McDowell, N., Allen, C. D., Anderson-Teixeira, K., et al. (2018). Drivers and mechanisms of tree mortality in moist tropical forests. New Phytologist, 219, 851869.Google Scholar
McDowell, N., Pockman, W. T., Allen, C. D., et al. (2008). Mechanisms of plant survival and mortality during drought: Why do some plants survive while others succumb to drought? New Phytologist, 178, 719739.Google Scholar
McDowell, N. G., Allen, C. D., Anderson-Teixeira, K., et al. (2020). Pervasive shifts in forest dynamics in a changing world. Science, 368, eaaz9463, DOI: https://doi.org/10.1126/science.aaz9463.Google Scholar
McDowell, N. G., Fisher, R. A., Xu, C., et al. (2013). Evaluating theories of drought-induced vegetation mortality using a multimodel–experiment framework. New Phytologist, 200, 304321.Google Scholar
McGuire, K. J., and Likens, G. E. (2011). Historical roots of forest hydrology and biogeochemistry. In Forest Hydrology and Biogeochemistry: Synthesis of Past Research and Future Directions, edited by Levia, D. F., Carlyle-Moses, D., and Tanaka, T.. Dordrecht: Springer, pp. 326.Google Scholar
McIntosh, R. P. (1985). The Background of Ecology: Concept and Theory. Cambridge, UK: Cambridge University Press.Google Scholar
McLachlan, J. S., Clark, J. S., and Manos, P. S. (2005). Molecular indicators of tree migration capacity under rapid climate change. Ecology, 86, 20882098.Google Scholar
McLeod, A. R., and Long, S. P. (1999). Free-air carbon dioxide enrichment (FACE) in global change research: A review. Advances in Ecological Research, 28, 156.Google Scholar
Meacham, J. (2012). Thomas Jefferson: The Art of Power. New York: Random House.Google Scholar
Medhurst, J., Parsby, J., Linder, S., et al. (2006). A whole-tree chamber system for examining tree-level physiological responses of field-grown trees to environmental variation and climate change. Plant, Cell and Environment, 29, 18531869.Google Scholar
Medlyn, B. E., Zaehle, S., De Kauwe, M. G., et al. (2015). Using ecosystem experiments to improve vegetation models. Nature Climate Change, 5, 528534.Google Scholar
Medrano, H., Gulías, J., Chaves, M. M., Galmés, J., and Flexas, J. (2012). Photosynthetic water-use efficiency. In Terrestrial Photosynthesis in a Changing Environment: A Molecular, Physiological and Ecological Approach, edited by Flexas, J., Loreto, F., and Medrano, H.. Cambridge, UK: Cambridge University Press, pp. 523536.Google Scholar
Medvigy, D., Walko, R. L., Otte, M. J., and Avissar, R. (2013). Simulated changes in Northwest U.S. climate in response to Amazon deforestation. Journal of Climate, 26, 91159136.Google Scholar
Meier, R., Schwaab, J., Seneviratne, S. I., et al. (2021). Empirical estimate of forestation-induced precipitation changes in Europe. Nature Geoscience, 14, 473478.Google Scholar
Meinshausen, M., Nicholls, Z. R. J., Lewis, J., et al. (2020). The shared socio-economic pathway (SSP) greenhouse gas concentrations and their extensions to 2500. Geoscientific Model Development, 13, 35713605.Google Scholar
Meir, P., Wood, T. E., Galbraith, D. R., et al. (2015). Threshold responses to soil moisture deficit by trees and soil in tropical rain forests: Insights from field experiments. BioScience, 65, 882892.Google Scholar
Meissner, K. J., Weaver, A. J., Matthews, H. D., and Cox, P. M. (2003). The role of land surface dynamics in glacial inception: A study with the UVic Earth System Model. Climate Dynamics, 21, 515537.Google Scholar
Melville, A. D. (1987). Ovid Metamorphoses, translated by A. D. Melville with introduction and notes by Kenney, E. J.. Oxford: Oxford University Press.Google Scholar
Menzies, N. K. (1994). Forest and Land Management in Imperial China. New York: St. Martin’s Press.Google Scholar
Mercado, L. M., Bellouin, N., Sitch, S., et al. (2009). Impact of changes in diffuse radiation on the global land carbon sink. Nature, 458, 10141017.Google Scholar
Meusel, J. G. (1781). Historische Litteratur für das Jahr 1781, vol. 2. Erlangen: Palm.Google Scholar
Michaux, F. A. (1819). The North American Sylva, 2 vols. Paris: C. D’Hautel.Google Scholar
Mikkelson, K. M., Bearup, L. A., Maxwell, R. M., et al. (2013). Bark beetle infestation impacts on nutrient cycling, water quality and interdependent hydrological effects. Biogeochemistry, 115, 121.Google Scholar
Miller, S. D., Goulden, M. L., Hutyra, L. R., et al. (2011). Reduced impact logging minimally alters tropical rainforest carbon and energy exchange. Proceedings of the National Academy of Sciences USA, 108, 1943119435.Google Scholar
Miralles, D. G., Gash, J. H., Holmes, T. R. H., de Jeu, R. A. M., and Dolman, A. J. (2010). Global canopy interception from satellite observations. Journal of Geophysical Research, 115, D16122, DOI: https://doi.org/10.1029/2009JD013530.Google Scholar
Miralles, D. G., Gentine, P., Seneviratne, S. I., and Teuling, A. J. (2019). Land-atmospheric feedbacks during droughts and heatwaves: State of the science and current challenges. Annals of the New York Academy of Sciences, 1436, 1935.Google Scholar
Miralles, D. G., Teuling, A. J., van Heerwaarden, C. C., and Vilà-Guerau, de Arellano, J. (2014). Mega-heatwave temperatures due to combined soil desiccation and atmospheric heat accumulation. Nature Geoscience, 7, 345349.Google Scholar
Mirbel, C.-F. Brisseau de (1815). Élémens de physiologie végétale et de botanique, 3 vols. Paris: Magimel.Google Scholar
Mitchell, J. (1767). The Present State of Great Britain and North America, with Regard to Agriculture, Population, Trade, and Manufactures, Impartially Considered. London: T. Becket & P. A. de Hondt.Google Scholar
Möller, D. (2020). Chemistry of the Climate System, vol. 2: History, Change and Sustainability, 3rd ed. Berlin: De Gruyter.Google Scholar
Moomaw, W. R., Masino, S. A., and Faison, E. K. (2019). Intact forests in the United States: Proforestation mitigates climate change and serves the greatest good. Frontiers in Forests and Global Change, 2, 27, DOI: https://doi.org/10.3389/ffgc.2019.00027.Google Scholar
Moon, D. (2010). The debate over climate change in the steppe region in nineteenth-century Russia. The Russian Review, 69, 251275.Google Scholar
Moon, D. (2013). The Plough that Broke the Steppes: Agriculture and Environment on Russia’s Grasslands, 1700–1914. Oxford: Oxford University Press.Google Scholar
Moore, W. L. (1910). A Report on the Influence of Forests on Climate and on Floods. Washington, DC: Government Printing Office.Google Scholar
More, T. (1973). The Confutation of Tyndale’s Answer, Books V–IX. In The Yale Edition of the Complete Works of St. Thomas More, vol. 8. The Confutation of Tyndale’s Answer, Part II The Text, Books V–IX, Appendices, edited by Schuster, L. A., Marius, R. C., Lusardi, J. P., and Schoeck, R. J.. New Haven, CT: Yale University Press, pp. 5751034.Google Scholar
de Jonnès, Moreau, A. (1825). Premier mémoire en réponse a la question proposée par l’Académie Royale de Bruxelles: Quels sont les changemens que peut occasioner le déboisement de forêts considérables sur les contrées et communes adjacentes…. Brussels: P. J. de Mat.Google Scholar
Morecroft, M. D., Taylor, M. E., and Oliver, H. R. (1998). Air and soil microclimates of deciduous woodland compared to an open site. Agricultural and Forest Meteorology, 90, 141156.Google Scholar
Moret, P., Muriel, P., Jaramillo, R., and Dangles, O. (2019). Humboldt’s Tableau Physique revisited. Proceedings of the National Academy of Sciences USA, 116, 1288912894.Google Scholar
Morice, C. P., Kennedy, J. J., Rayner, N. A., et al. (2021). An updated assessment of near-surface temperature change from 1850: The HadCRUT5 data set. Journal of Geophysical Research: Atmospheres, 126, e2019JD032361, DOI: https://doi.org/10.1029/2019JD032361.Google Scholar
Morueta-Holme, N., Engemann, K., Sandoval-Acuña, P., et al. (2015). Strong upslope shifts in Chimborazo’s vegetation over two centuries since Humboldt. Proceedings of the National Academy of Sciences USA, 112, 1274112745.Google Scholar
Mueller, F. (1867). Australian Vegetation, Indigenous or Introduced, Considered Especially in its Bearings on the Occupation of the Territory, and with a View of Unfolding its Resources. Melbourne: Blundell.Google Scholar
Mueller, N. D., Butler, E. E., McKinnon, K. A., et al. (2016). Cooling of US Midwest summer temperature extremes from cropland intensification. Nature Climate Change, 6, 317322.Google Scholar
Muir, J. (1894). The Mountains of California. New York: Century.Google Scholar
Munns, E. N. (1930). An East African estimate of forest influences on climate and water supply. Forest Worker, 6(1), 2425.Google Scholar
Murchison, R. I., Verneuil, E. de., and Keyserling, A. von (1845). The Geology of Russia in Europe and the Ural Mountains, 2 vols. London: John Murray.Google Scholar
Murray, J. (1831). On raining trees. Magazine of Natural History, and Journal of Zoology, Botany, Mineralogy, Geology, and Meteorology, 4, 3234.Google Scholar
Müttrich, A. (1890). Ueber den Einflusß des Waldes auf die periodischen Veränderungen der Lufttemperatur. Zeitschrift für Forst- und Jagdwesen, 22, 385400, 449458, 513526.Google Scholar
Müttrich, A. (1892). Ueber den Einflusß des Waldes auf die Größe der atmosphärischen Niederschläge. Zeitschrift für Forst- und Jagdwesen, 24, 2742.Google Scholar
Nabuurs, G.-J., Delacote, P., Ellison, D., et al. (2017). By 2050 the mitigation effects of EU forests could nearly double through climate smart forestry. Forests, 8, 484, DOI: https://doi.org/10.3390/f8120484.Google Scholar
Nakai, T., Sumida, A., Daikoku, K., et al. (2008). Parameterisation of aerodynamic roughness over boreal, cool- and warm-temperate forests. Agricultural and Forest Meteorology, 148, 19161925.Google Scholar
Napier, M. (1842). The Encyclopædia Britannica, or Dictionary of Arts, Sciences, and General Literature, 7th ed., 21 vols. Edinburgh: Adam & Charles Black.Google Scholar
Nash, L. K. (1952). Plants and the Atmosphere (Harvard Case Histories in Experimental Science, Case 5). Cambridge, MA: Harvard University Press.Google Scholar
National Research Council (1986). Earth System Science – Overview: A Program for Global Change. Washington, DC: National Academies Press.Google Scholar
Naudts, K., Chen, Y., McGrath, M. J., et al. (2016). Europe’s forest management did not mitigate climate warming. Science, 351, 597600.Google Scholar
Negrón, J. F., and Cain, B. (2019). Mountain pine beetle in Colorado: A story of changing forests. Journal of Forestry, 117, 144151.CrossRefGoogle Scholar
Negrón-Juárez, R., Baker, D. B., Zeng, H., Henkel, T. K., and Chambers, J. Q. (2010). Assessing hurricane‐induced tree mortality in U.S. Gulf Coast forest ecosystems. Journal of Geophysical Research, 115, G04030, DOI: https://doi.org/10.1029/2009JG001221.Google Scholar
Negrón-Juárez, R. I., Holm, J. A., Marra, D. M., et al. (2018). Vulnerability of Amazon forests to storm-driven tree mortality. Environmental Research Letters, 13, 054021, DOI: https://doi.org/10.1088/1748-9326/aabe9f.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. Philosophical Transactions of the Royal Society B, 363, 17371746.Google Scholar
Newbold, T. J. (1839). Notice of river dunes on the banks of the Hogri and Pennaur. Madras Journal of Literature and Science, 9(23), 309310.Google Scholar
Newell, R. E. (1971). The Amazon forest and atmospheric general circulation. In Man’s Impact on the Climate, edited by Matthews, W. H., Kellogg, W. W., and Robinson, G. D.. Cambridge, MA: Massachusetts Institute of Technology Press, pp. 457459.Google Scholar
Nicholson, H. (1676). An extract of a letter &c. from Dublin May the 10th, 1676. Philosophical Transactions, 11(127), 647653.Google Scholar
Nicholson, J. W. (1929). The Influence of Forests on Climate and Water Supply in Kenya, Forestry Department Pamphlet Number 2. Nairobi: East African Standard.Google Scholar
Nicholson, S. E. (2013). The West African Sahel: A review of recent studies on the rainfall regime and its interannual variability. International Scholarly Research Notices: Meteorology, 453521, DOI: https://doi.org/10.1155/2013/453521.Google Scholar
Niinemets, Ü., and Anten, N. P. R. (2009). Packing the photosynthetic machinery: From leaf to canopy. In Photosynthesis in silico: Understanding Complexity from Molecules to Ecosystems, edited by Laisk, A., Nedbal, L., and Govindjee, . Dordrecht: Springer, pp. 363399.Google Scholar
Niinemets, Ü., Keenan, T. F., and Hallik, L. (2015). A worldwide analysis of within-canopy variations in leaf structural, chemical and physiological traits across plant functional types. New Phytologist, 205, 973993.Google Scholar
Nisbet, J. (1893). The Climatic and National-Economic Influence of Forests. London: Eyre & Spottiswoode.Google Scholar
Nisbet, J. (1894). The climatic and national-economic influence of forests. Nature, 49, 302305.Google Scholar
Nisbet, J. (1905). The Forester: A Practical Treatise on British Forestry and Arboriculture for Landowners, Land Agents, and Foresters, 2 vols. Edinburgh: William Blackwood.Google Scholar
Nobre, C. A., and Borma, L. S. (2009). “Tipping points” for the Amazon forest. Current Opinion in Environmental Sustainability, 1, 2836.Google Scholar
Nobre, C. A., Silva Dias, M. A., Culf, A. D., et al. (2004). The Amazonian climate. In Vegetation, Water, Humans and the Climate: A New Perspective on an Interactive System, edited by Kabat, P., Claussen, M., Dirmeyer, P. A., et al. Berlin: Springer, pp. 7992.Google Scholar
Norby, R. J., and Zak, D. R. (2011). Ecological lessons from free-air CO2 enrichment (FACE) experiments. Annual Review of Ecology, Evolution, and Systematics, 42, 181203.Google Scholar
Nördlinger, T. (1885). Der Einfluss des Waldes auf die Luft- und Bodenwärme. Berlin: Paul Parey.Google Scholar
Novick, K. A., and Katul, G. G. (2020). The duality of reforestation impacts on surface and air temperature. Journal of Geophysical Research: Biogeosciences, 125, DOI: https://doi.org/10.1029/2019JG005543.Google Scholar
Oberthaler, E., Pénot, S., Sellink, M., Spronk, R., and Hoppe-Harnoncourt, A. (2018). Bruegel: The Hand of the Master – Exhibition Catalogue of the Kunsthistorisches Museum Vienna. Vienna: KHM-Museumsverband.Google Scholar
O’Halloran, T. L., Law, B. E., Goulden, M. L., et al. (2012). Radiative forcing of natural forest disturbances. Global Change Biology, 18, 555565.Google Scholar
Olearius, A. (1662). The Voyages & Travels of the Ambassadors from the Duke of Holstein, to the Great Duke of Muscovy, and the King of Persia…, rendered into English by John Davies. London: Thomas Dring; John Starkey.Google Scholar
Oleson, K. W., Bonan, G. B., Levis, S., and Vertenstein, M. (2004). Effects of land use change on North American climate: Impact of surface datasets and model biogeophysics. Climate Dynamics, 23, 117132.Google Scholar
O’Neill, B. C., Tebaldi, C., van Vuuren, D. P., et al. (2016). The Scenario Model Intercomparison Project (ScenarioMIP) for CMIP6. Geoscientific Model Development, 9, 34613482.Google Scholar
O’Neill, B. C., Kriegler, E., Ebi, K. L., et al. (2017). The roads ahead: Narratives for shared socioeconomic pathways describing world futures in the 21st century. Global Environmental Change, 42, 169180.Google Scholar
Oosting, H. J. (1942). An ecological analysis of the plant communities of Piedmont, North Carolina. American Midland Naturalist, 28, 1126.Google Scholar
Ornstein, L., Aleinov, I., and Rind, D. (2009). Irrigated afforestation of the Sahara and Australian Outback to end global warming. Climatic Change, 97, 409437.Google Scholar
Oswald, F. L. (1877). The climatic influence of vegetation. – A plea for our forests. Popular Science Monthly, 11(August), 385390.Google Scholar
Otto-Bliesner, B. L., Brady, E. C., Fasullo, F., et al. (2016). Climate variability and change since 850 CE: An ensemble approach with the Community Earth System Model (CESM). Bulletin of the American Meteorological Society, 97, 735754.Google Scholar
Oudin, L., Andréassian, V., Lerat, L., and Michel, C. (2008). Has land cover a significant impact on mean annual streamflow? An international assessment using 1508 catchments. Journal of Hydrology, 357, 303316.Google Scholar
Overpeck, J. T., Webb, R. S., and Webb, T., III (1992). Mapping eastern North American vegetation change of the past 18 ka: No-analogs and the future. Geology, 20, 10711074.Google Scholar
Oyama, M. D., and Nobre, C. A. (2003). A new climate-vegetation equilibrium state for Tropical South America. Geophysical Research Letters, 30, 2199, DOI: https://doi.org/10.1029/2003GL018600.Google Scholar
Paasonen, P., Asmi, A., Petäjä, T., et al. (2013). Warming-induced increase in aerosol number concentration likely to moderate climate change. Nature Geoscience, 6, 438442.Google Scholar
Pakenham, T. (1996). Meetings with Remarkable Trees. London: Weidenfeld & Nicolson.Google Scholar
Pakenham, T. (2002). Remarkable Trees of the World. London: Weidenfeld & Nicolson.Google Scholar
Pan, Y., Birdsey, R. A., Phillips, O. L., and Jackson, R. B. (2013). The structure, distribution, and biomass of the world’s forests. Annual Review of Ecology, Evolution, and Systematics, 44, 593622.Google Scholar
Park, J.-H., Goldstein, A. H., Timkovsky, J., et al. (2013). Active atmosphere-ecosystem exchange of the vast majority of detected volatile organic compounds. Science, 341, 643647.Google Scholar
Parker, D. H. (2008). A Critical Edition of Robert Barnes’ A Supplication Unto the Most Gracyous Prince Kynge Henry The VIII, 1534. Toronto: University of Toronto Press.Google Scholar
Parks, S. A., and Abatzoglou, J. T. (2020). Warmer and drier fire seasons contribute to increases in area burned at high severity in western US forests from 1985 to 2017. Geophysical Research Letters, 47, e2020GL089858, DOI: https://doi.org/10.1029/2020GL089858.Google Scholar
Parsons, L. A., Jung, J., Masuda, Y. J., et al. (2021). Tropical deforestation accelerates local warming and loss of safe outdoor working hours. One Earth, 4, 17301740.Google Scholar
Paso y Troncoso, F. del (1905). Papeles de Nueva España, segunda serie: Geografía y estadística, vol. 4. Madrid: Establicimiento Tipográfico “Sucesores de Rivadeneyra.”Google Scholar
Pastorello, G., Trotta, C., Canfora, E., et al. (2020). The FLUXNET2015 dataset and the ONEFlux processing pipeline for eddy covariance data. Scientific Data, 7, 225, DOI: https://doi.org/10.1038/s41597-020-0534-3.Google Scholar
Pausata, F. S. R., Gaetani, M., Messori, G., et al. (2020). The greening of the Sahara: Past changes and future implications. One Earth, 2, 235250.CrossRefGoogle Scholar
Payne, C. (2017). Silent Witnesses: Trees in British Art, 1760–1870. Bristol: Sansom.Google Scholar
Pelloutier, S. (1740). Histoire des Celtes, et particulierement des Gaulois et des Germains, depuis les tems fabuleux, jusqu’à la prise de Rome par les Gaulois. La Haye: Isaac Beauregard.Google Scholar
Peng, S.-S., Piao, S., Zeng, Z., et al. (2014). Afforestation in China cools local land surface temperature. Proceedings of the National Academy of Sciences USA, 111, 29152919.Google Scholar
Peppercorne, F. S. (1879). Influence of forests on climate and rainfall. Transactions and Proceedings of the New Zealand Institute, 12, 2432.Google Scholar
Perugini, L., Caporaso, L., Marconi, S., et al. (2017). Biophysical effects on temperature and precipitation due to land cover change. Environmental Research Letters, 12, 053002, DOI: https://doi.org/10.1088/1748-9326/aa6b3f.Google Scholar
Petäjä, T., Tabakova, K., Manninen, A., et al. (2022). Influence of biogenic emissions from boreal forests on aerosol–cloud interactions. Nature Geoscience, 15, 4247.Google Scholar
Phillips, O. L., Aragão, L. E. O. C., Lewis, S. L., et al. (2009). Drought sensitivity of the Amazon rainforest. Science, 323, 13441347.Google Scholar
Piao, S., Wang, X., Park, T., et al. (2020). Characteristics, drivers and feedbacks of global greening. Nature Reviews Earth and Environment, 1, 1427.Google Scholar
Pielke, R. A., Lee, T. J., Copeland, J. H., et al. (1997). Use of USGS-provided data to improve weather and climate simulations. Ecological Applications, 7, 321.Google Scholar
Pielke, R. A., Rodriguez, J. H., Eastman, J. L., Walko, R. L., and Stocker, R. A. (1993). Influence of albedo variability in complex terrain on mesoscale systems. Journal of Climate, 6, 17981806.Google Scholar
Pielke, R. A., Sr., Pitman, A., Niyogi, D., et al. (2011). Land use/land cover changes and climate: Modeling analysis and observational evidence. WIREs Climate Change, 2, 828850.Google Scholar
Pigafetta, A. (1906). Magellan’s Voyage around the World, translated by J. A. Robertson, 2 vols. Cleveland, OH: Arthur H. Clark.Google Scholar
Pigafetta, A. (1985). Primer viaje alrededor del mundo, edición de Leoncio Cabrero. Madrid: Historia 16.Google Scholar
Pinchot, G. (1905a). A Primer of Forestry. Part II. – Practical Forestry, Bulletin Number 24, Part II. U.S. Department of Agriculture, Bureau of Forestry. Washington, DC: Government Printing Office.Google Scholar
Pinchot, G. (1905b). The Use of the National Forest Reserves: Regulations and Instructions. Washington, DC: U.S. Department of Agriculture.Google Scholar
Pinchot, G. (2017). Dear Forester, February 1, 1905. In Gifford Pinchot: Selected Writings, edited by Miller, C.. University Park, PA: Pennsylvania State University Press, pp. 39-42.Google Scholar
Piper, R. U. (1855). The Trees of America. Boston: William White.Google Scholar
Pitman, A. J., and Lorenz, R. (2016). Scale dependence of the simulated impact of Amazonian deforestation on regional climate. Environmental Research Letters, 11, 094025, DOI: https://doi.org/10.1088/1748-9326/11/9/094025.Google Scholar
Pliny (1963). Natural History, vol. 8. Books 28–32, translated by W. H. S. Jones (Loeb Classical Library 418). Cambridge, MA: Harvard University Press.Google Scholar
Poivre, P. (1768). Voyages d’un philosophe: ou, observations sur les moeurs et les arts des peuples de l’Afrique, de l’Asie et de l’Amerique. Yverdon: n.p.Google Scholar
Poivre, P. (1797). Oeuvres complettes de P. Poivre, intendant des Isles de France et de Bourbon, correspondant de l’académie des sciences, etc. Paris: Fuchs.Google Scholar
Pongratz, J., Reick, C. H., Raddatz, T., and Claussen, M. (2010). Biogeophysical versus biogeochemical climate response to historical anthropogenic land cover change. Geophysical Research Letters, 37, L08702, DOI: https://doi.org/10.1029/2010GL043010.Google Scholar
Pongratz, J., Reick, C. H., Houghton, R. A., and House, J. I. (2014). Terminology as a key uncertainty in net land use and land cover change carbon flux estimates. Earth System Dynamics, 5, 177195.Google Scholar
Pongratz, J., Schwingshackl, C., Bultan, S., et al. (2021). Land use effects on climate: Current state, recent progress, and emerging topics. Current Climate Change Reports, 7, 99120.Google Scholar
Popp, A., Calvin, K., Fujimori, S., et al. (2017). Land-use futures in the shared socio-economic pathways. Global Environmental Change, 42, 331345.Google Scholar
Pöschl, U., Martin, S. T., Sinha, B., et al. (2010). Rainforest aerosols as biogenic nuclei of clouds and precipitation in the Amazon. Science, 329, 15131516.Google Scholar
Postel, S. L., and Thompson, B. H. Jr. (2005). Watershed protection: Capturing the benefits of nature’s water supply services. Natural Resources Forum, 29, 98108.Google Scholar
Potter, S., Solvik, K., Erb, A., et al. (2020). Climate change decreases the cooling effect from postfire albedo in boreal North America. Global Change Biology, 26, 15921607.Google Scholar
Poyatos, R., Granda, V., Molowny-Horas, R., et al. (2016). SAPFLUXNET: Towards a global database of sap flow measurements. Tree Physiology, 36, 11491455.Google Scholar
Prentice, I. C., Jolly, D., and BIOME6000 (2000). Mid-Holocene and glacial-maximum vegetation geography of the northern continents and Africa. Journal of Biogeography, 27, 507519.Google Scholar
Prévost, A.-F., l’abbé (1744). Voyages du capitaine Robert Lade en differentes parties de l’Afrique, de l’Asie et de l’Amerique, 2 vols. Paris: Didot.Google Scholar
Priestley, J. (1772). Observations on different kinds of air. Philosophical Transactions, 62, 147264.Google Scholar
Pross, J., Contreras, L., Bijl, P. K., et al. (2012). Persistent near-tropical warmth on the Antarctic continent during the early Eocene epoch. Nature, 488, 7377.Google Scholar
Pukkala, T. (2018). Effect of species composition on ecosystem services in European boreal forest. Journal of Forestry Research, 29, 261272.Google Scholar
Purkyně, E. (1875). Etwas über die Waldfrage, Wasserfrage und Sumpffrage. Oesterreichische Monatsschrift für Forstwesen, 25, 479525.Google Scholar
Purkyně, E. (1876). Ueber die Wald- und Wasserfrage. Oesterreichische Monatsschrift für Forstwesen, 26, 136151, 161178, 179204, 209251, 267291, 327349, 405426, 473498.Google Scholar
Purkyně, E. (1877). Ueber die Wald- und Wasserfrage. Oesterreichische Monatsschrift für Forstwesen, 27, 102143.Google Scholar
Quesada, B., Arneth, A., and de Noblet-Ducoudré, N. (2017). Atmospheric, radiative, and hydrologic effects of future land use and land cover changes: A global and multimodel climate picture. Journal of Geophysical Research: Atmospheres, 122, 51135131.Google Scholar
R. C. (1912). Forests and rainfall. Nature, 89, 662664.Google Scholar
R. D. (1908). Les forêts de la planète Mars. Revue des eaux et forêts, 47, 404406.Google Scholar
Rackham, O. (1986). The History of the Countryside. London: J. M. Dent.Google Scholar
Rae, J. W. B., Zhang, Y. G., Liu, X., et al. (2021). Atmospheric CO2 over the past 66 million years from marine archives. Annual Review of Earth and Planetary Sciences, 49, 609641.Google Scholar
Rajan, S. R. (2006). Modernizing Nature: Forestry and Imperial Eco-Development 1800–1950. Oxford: Oxford University Press.Google Scholar
Ramsay, D. (1809). The History of South-Carolina, from Its First Settlement in 1670, to the Year 1808, 2 vols. Charleston, SC: David Longworth.Google Scholar
Randerson, J. T., Liu, H., Flanner, M. G., et al. (2006). The impact of boreal forest fire on climate warming. Science, 314, 11301132.Google Scholar
Rap, A., Scott, C. E., Reddington, C. L., et al. (2018). Enhanced global primary production by biogenic aerosol via diffuse radiation fertilization. Nature Geoscience, 11, 640644.Google Scholar
Rauch, F.-A. (1792). Plan nourricier, ou recherches sur les moyens à mettre en usage pour assurer à jamais le pain au peuple français…. Didot jeune: Paris.Google Scholar
Rauch, F. A. (1802). Harmonie hydro-végétale et météorologique, ou recherches sur les moyens de recréer avec nos forêts la force des températures et la régularité des saisons, par des plantations raisonnées, 2 vols. Paris: Levrault.Google Scholar
Rauch, F.-A. (1818). Régénération de la nature végétale, ou recherches sur les moyens de recréer, dans tous les climats, les anciennes températures et l’ordre primitif des saisons, par des plantations raisonnées, 2 vols. Paris: Didot l’Ainé.Google Scholar
Raupach, M. R. (1988). Canopy transport processes. In Flow and Transport in the Natural Environment: Advances and Applications, edited by Steffen, W. L. and Denmead, O. T.. Berlin: Springer-Verlag, pp. 95127.Google Scholar
Raupach, M. R. (1994). Simplified expressions for vegetation roughness length and zero-plane displacement as functions of canopy height and area index. Boundary-Layer Meteorology, 71, 211216.Google Scholar
Raynor, G. S. (1971). Wind and temperature structure in a coniferous forest and a contiguous field. Forest Science, 17, 351363.Google Scholar
Reifsnyder, W. E. (1973). Forest meteorology in the seventies. Bulletin of the American Meteorological Society, 54, 326330.Google Scholar
Ren, J., Adam, J. C., Hicke, J. A., et al. (2021). How does water yield respond to mountain pine beetle infestation in a semiarid forest? Hydrology and Earth System Sciences, 25, 46814699.Google Scholar
Renou, E. (1866). Théorie de la pluie. Annuaire de la société météorologique de France, 14, 89106.Google Scholar
Restrepo-Coupe, N., Albert, L. P., Longo, M., et al. (2021). Understanding water and energy fluxes in the Amazonia: Lessons from an observation-model intercomparison. Global Change Biology, 27, 18021819.Google Scholar
Riahi, K., van Vuuren, D. P., Kriegler, E., et al. (2017). The shared socioeconomic pathways and their energy, land use, and greenhouse gas emissions implications: An overview. Global Environmental Change, 42, 153168.Google Scholar
Ribbentrop, B. (1900). Forestry in British India. Calcutta: Office of the Superintendent of Government Printing.Google Scholar
Richardson, L. F. (1922). Weather Prediction by Numerical Processes. Cambridge, UK: Cambridge University Press.Google Scholar
Ridgway, R. (1872). Notes on the vegetation of the lower Wabash Valley. American Naturalist, 6, 658665.Google Scholar
Ridgwell, A., Singarayer, J. S., Hetherington, A. M., and Valdes, P. J. (2009). Tackling regional climate change by leaf albedo bio-geoengineering. Current Biology, 19, 146150.Google Scholar
Roberts, J., Cabral, O. M. R., and De Aguiar, L. F. (1990). Stomatal and boundary-layer conductances in an Amazonian terra firme rain forest. Journal of Applied Ecology, 27, 336353.Google Scholar
Roberts, P., Boivin, N., Lee-Thorp, J., Petraglia, M., and Stock, J. (2016). Tropical forests and the genus Homo. Evolutionary Anthropology, 25, 306317.Google Scholar
Robertson, W. (1777). The History of America, 2 vols. London: W. Strahan.Google Scholar
Robinson, N. A. (2014). The Charter of the Forest: Evolving human rights in nature. In Magna Carta and the Rule of Law, edited by Magraw, D. B., Martinez, A., and Brownell, R. E. II. Chicago: American Bar Association, pp. 311377.Google Scholar
Rodman, K. C., Veblen, T. T., Battaglia, M. A., et al. (2020). A changing climate is snuffing out post-fire recovery in montane forests. Global Ecology and Biogeography, 29, 20392051.Google Scholar
Rodrigues, L. (2007). Dr. Alexander Gibson and the emergence of conservationism and desiccationism in Bombay: 1838 to 1860. In Proceedings of the Indian History Congress, 67th Session, Calicut 2006–07. Delhi: Indian History Congress, pp. 655665.Google Scholar
Rodríguez-Fonseca, B., Mohino, E., Mechoso, C. R., et al. (2015). Variability and predictability of West African droughts: A review on the role of sea surface temperature anomalies. Journal of Climate, 28, 40344060.Google Scholar
Roe, S., Streck, C., Obersteiner, M., et al. (2019). Contribution of the land sector to a 1.5 °C world. Nature Climate Change, 9, 817828.Google Scholar
Roe, S., Streck, C., Beach, R., et al. (2021). Land-based measures to mitigate climate change: Potential and feasibility by country. Global Change Biology, 27, 60256058.Google Scholar
Rogelj, J., Popp, A., Calvin, K. V., et al. (2018a). Scenarios towards limiting global mean temperature increase below 1.5 °C. Nature Climate Change, 8, 325332.Google Scholar
Rogelj, J., Shindell, D., Jiang, K., et al. (2018b). Mitigation pathways compatible with 1.5°C in the context of sustainable development. In Global Warming of 1.5°C: An IPCC Special Report on the Impacts of Global Warming of 1.5°C Above Pre-Industrial Levels and Related Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the Global Response to the Threat of Climate Change, Sustainable Development, and Efforts to Eradicate Poverty, edited by Masson-Delmotte, V., Zhai, P., Pörtner, H.-O., et al. Geneva: World Meteorological Organization, pp. 93174.Google Scholar
Rogers, B. M., Randerson, J. T., and Bonan, G. B. (2013). High-latitude cooling associated with landscape changes from North American boreal forest fires. Biogeosciences, 10, 699718.Google Scholar
Rogers, B. M., Soja, A. J., Goulden, M. L., and Randerson, J. T. (2015). Influence of tree species on continental differences in boreal fires and climate feedbacks. Nature Geoscience, 8, 228234.Google Scholar
Rogers, H. (1873). Report on the effects of the cutting down of forests on the climate and health of the Mauritius. Transactions of the Botanical Society of Edinburgh, 11, 115118.Google Scholar
Rohatyn, S., Rotenberg, E., Ramati, E., et al. (2018). Differential impacts of land use and precipitation on “ecosystem water yield.” Water Resources Research, 54, 54575470.Google Scholar
Rooke, H. (1790). Descriptions and Sketches of Some Remarkable Oaks, in the Park at Welbeck, in the County of Nottingham, a Seat of His Grace The Duke of Portland. London: J. Nichols.Google Scholar
Rotenberg, E., and Yakir, D. (2010). Contribution of semi-arid forests to the climate system. Science, 327, 451454.Google Scholar
Rotenberg, E., and Yakir, D. (2011). Distinct patterns of changes in surface energy budget associated with forestation in the semiarid region. Global Change Biology, 17, 15361548.Google Scholar
Rothwell, G. W., Mapes, G., and Mapes, R. H. (1997). Late Paleozoic conifers of North America: Structure, diversity and occurrences. Review of Palaeobotany and Palynology, 95, 95113.Google Scholar
Rothwell, G. W., Mapes, G., Stockey, R. A., and Hilton, J. (2012). The seed cone Eathiestrobus gen. nov.: Fossil evidence for a Jurassic origin of Pinaceae. American Journal of Botany, 99, 708720.Google Scholar
Royer, D. L., Osborne, C. P., and Beerling, D. J. (2003). Carbon loss by deciduous trees in a CO2-rich ancient polar environment. Nature, 424, 6062.Google Scholar
Rugendas, J. M. (1827). Voyage pittoresque dans le Brésil. Paris: Engelmann & Cie.Google Scholar
Rush, B. (1786). An enquiry into the cause of the increase of bilious and intermitting fevers in Pennsylvania, with hints for preventing them. Transactions of the American Philosophical Society, 2, 206212.Google Scholar
Rutkow, E. (2012). American Canopy: Trees, Forests, and the Making of a Nation. New York: Scribner.Google Scholar
Ryan, M. G. (2013). Three decades of research at Flakaliden advancing whole-tree physiology, forest ecosystem and global change research. Tree Physiology, 33, 11231131.Google Scholar
Saberwal, V. K. (1998). Science and the desiccationist discourse of the 20th century. Environment and History, 4, 309343.Google Scholar
Sagan, C., Toon, O. B., and Pollack, J. B. (1979). Anthropogenic albedo changes and the Earth’s climate. Science, 206, 13631368.Google Scholar
Saint, J. (2021). Ariadne. New York: Flatiron Books.Google Scholar
Sallon, S., Solowey, E., Cohen, Y., et al. (2008). Germination, genetics, and growth of an ancient date seed. Science, 320, 1464.Google Scholar
Sallon, S., Cherif, E., Chabrillange, N., et al. (2020). Origins and insights into the historic Judean date palm based on genetic analysis of germinated ancient seeds and morphometric studies. Science Advances, 6, eaax0384, DOI: https://doi.org/10.1126/sciadv.aax0384.Google Scholar
Santopuoli, G., Temperli, C., Alberdi, I., et al. (2021). Pan-European sustainable forest management indicators for assessing Climate-Smart Forestry in Europe. Canadian Journal of Forest Research, 51, 17411750.Google Scholar
Sargent, C.S. (1882). The protection of forests. North American Review, 135, 386401.Google Scholar
Sargent, C. S., Abbot, H. L., Agassiz, A., et al. (1897). Report of the Committee Appointed by the National Academy of Sciences upon the Inauguration of a Forest Policy For the Forested Lands of the United States to the Secretary of the Interior, May 1, 1897. Washington, DC: Government Printing Office.Google Scholar
Saunders, C. J. (1993). The Forest of Medieval Romance: Avernus, Broceliande, Arden. Cambridge, UK: D. S. Brewer.Google Scholar
SCEP (1970). Man’s Impact on the Global Environment – Assessment and Recommendations for Actions: Report of the Study of Critical Environmental Problems (SCEP). Cambridge, MA: Massachusetts Institute of Technology Press.Google Scholar
Schacht, H. (1853). Der Baum: Studien über Bau und Leben der höheren Gewächse. Berlin: G. W. F. Müller.Google Scholar
Scheffers, B. R., Phillips, B. L., Laurance, W. F., et al. (2013). Increasing arboreality with altitude: A novel biogeographic dimension. Proceedings of the Royal Society B, 280, 20131581, DOI: https://doi.org/10.1098/rspb.2013.1581.Google Scholar
Schiff, A. L. (1962). Fire and Water: Scientific Heresy in the Forest Service. Cambridge, MA: Harvard University Press.Google Scholar
Schimel, D., Schneider, F. D., Bloom, A., et al. (2019). Flux towers in the sky: Global ecology from space. New Phytologist, 224, 570584.Google Scholar
Schleiden, M. J. (1848a). Die Pflanze und ihr Leben. Leipzig: Wilhelm Engelmann.Google Scholar
Schleiden, M. J. (1848b). The Plant; A Biography, translated by A. Henfrey. London: Hippolyte Bailliere.Google Scholar
Schlich, W. (1889). A Manual of Forestry, vol. 1. The Utility of Forests, and Fundamental Principles of Sylviculture. London: Bradbury, Agnew & Co.Google Scholar
Schlich, W. (1910). Forests and forestry. In The Encyclopædia Britannica: A Dictionary of Arts, Sciences, Literature and General Information, 11th ed., vol. 10. New York: Encyclopædia Britannica Company, pp. 645651.Google Scholar
Schmidt, M. (2019). Gilgamesh: The Life of a Poem. Princeton: Princeton University Press.Google Scholar
Schneider, S. H., and Dickinson, R. E. (1974). Climate modeling. Reviews of Geophysics and Space Physics, 12, 447493.Google Scholar
Schofield, P. F. (1875). Forests and rainfall. Popular Science Monthly, 8(November), 111112.Google Scholar
Schöpf, J. D. (1875). The Climate and Diseases of America, translated by Chadwick, J. R.. Boston: H. O. Houghton.Google Scholar
Schott, C. A. (1872). Tables and Results of the Precipitation, in Rain and Snow, in the United States: And at Some Stations in Adjacent Parts of North America, and in Central and South America. Washington, DC: Smithsonian Institution.Google Scholar
Schott, C. A. (1876). Tables, Distribution, and Variations of the Atmospheric Temperature in the United States, and Some Adjacent Parts of America. Washington, DC: Smithsonian Institution.Google Scholar
Schubert, S. D., Stewart, R. E., Wang, H., et al. (2016). Global meteorological drought: A synthesis of current understanding with a focus on SST drivers of precipitation deficits. Journal of Climate, 29, 39894019.Google Scholar
Schubert, S. D., Suarez, M. J., Pegion, P. J., Koster, R. D., and Bacmeister, J. T. (2004). On the cause of the 1930s Dust Bowl. Science, 303, 18551859.Google Scholar
Schultz, N. M., Lawrence, P. J., and Lee, X. (2017). Global satellite data highlights the diurnal asymmetry of the surface temperature response to deforestation. Journal of Geophysical Research: Biogeosciences, 122, 903917.Google Scholar
Schulze, E. D., Sierra, C. A., Egenolf, V., et al. (2020). The climate change mitigation effect of bioenergy from sustainably managed forests in Central Europe. GCB Bioenergy, 12, 186197.Google Scholar
Schuster, L. A. (1973). Thomas More’s polemical career, 1523–1533. In The Yale Edition of the Complete Works of St. Thomas More, vol. 8. The Confutation of Tyndale’s Answer, Part III Introduction, Commentary, Glossary, Index, edited by Schuster, L. A., Marius, R. C., Lusardi, J. P., and Schoeck, R. J.. New Haven, CT: Yale University Press, pp. 11351268.Google Scholar
Schwaab, J., Davin, E. L., Bebi, P., et al. (2020). Increasing the broad-leaved tree fraction in European forests mitigates hot temperature extremes. Scientific Reports, 10, 14153, DOI: https://doi.org/10.1038/s41598-020-71055-1.Google Scholar
Schwartz, M. D., and Karl, T. R. (1990). Spring phenology: Nature’s experiment to detect the effect of “green-up” on surface maximum temperatures. Monthly Weather Review, 118, 883890.Google Scholar
Schwingshackl, C., Davin, E. L., Hirschi, M., et al. (2019). Regional climate model projections underestimate future warming due to missing plant physiological CO2 response. Environmental Research Letters, 14, 114019, DOI: https://doi.org/10.1088/1748-9326/ab4949.Google Scholar
Scott, C. E., Monks, S. A., Spracklen, D. V., et al. (2018). Impact on short-lived climate forcers increases projected warming due to deforestation. Nature Communications, 9, 157, DOI: https://doi.org/10.1038/s41467-017-02412-4.Google Scholar
Scott, C. E., Rap, A., Spracklen, D. V., et al. (2014). The direct and indirect radiative effects of biogenic secondary organic aerosol. Atmospheric Chemistry and Physics, 14, 447470.Google Scholar
Seager, R., and Hoerling, M. (2014). Atmosphere and ocean origins of North American droughts. Journal of Climate, 27, 45814606.Google Scholar
Searchinger, T. D., Hamburg, S. P., Melillo, J., et al. (2009). Fixing a critical climate accounting error. Science, 326, 527528.Google Scholar
Segal, M., Avissar, R., McCumber, M. C., and Pielke, R. A. (1988). Evaluation of vegetation effects on the generation and modification of mesoscale circulations. Journal of the Atmospheric Sciences, 45, 22682292.Google Scholar
Sellers, P. J., Bounoua, L., Collatz, G. J., et al. (1996a). Comparison of radiative and physiological effects of doubled atmospheric CO2 on climate. Science, 271, 14021406.Google Scholar
Sellers, P. J., Mintz, Y., Sud, Y. C., and Dalcher, A. (1986). A simple biosphere model (SiB) for use within general circulation models. Journal of the Atmospheric Sciences, 43, 505531.Google Scholar
Sellers, P. J., Randall, D. A., Collatz, G. J., et al. (1996b). A revised land surface parameterization (SiB2) for atmospheric GCMs. Part I: Model formulation. Journal of Climate, 9, 676705.Google Scholar
Sellink, M. (2019). Leading the eye and staging the composition: Some remarks on Pieter Bruegel the Elder’s compositional techniques. In Bruegel: The Hand of the Master. The 450th Anniversary Edition – Essays in Context, edited by Hoppe-Harnoncourt, A., Oberthaler, E., Pénot, S., Sellink, M., and Spronk, R.. Vienna: KHM-Museumsverband, pp. 336352.Google Scholar
Seneviratne, S. I., Corti, T., Davin, E. L., et al. (2010). Investigating soil moisture-climate interactions in a changing climate: A review. Earth-Science Reviews, 99, 125161.Google Scholar
Seneviratne, S. I., Phipps, S. J., Pitman, A. J., et al. (2018). Land radiative management as contributor to regional-scale climate adaptation and mitigation. Nature Geoscience, 11, 8896.Google Scholar
Senf, C., Buras, A., Zang, C. S., Rammig, A., and Seidl, R. (2020). Excess forest mortality is consistently linked to drought across Europe. Nature Communications, 11, 6200, DOI: https://doi.org/10.1038/s41467-020-19924-1.Google Scholar
Senf, C., Pflugmacher, D., Zhiqiang, Y., et al. (2018). Canopy mortality has doubled in Europe’s temperate forests over the last three decades. Nature Communications, 9, 4978, DOI: https://doi.org/10.1038/s41467-018-07539-6.Google Scholar
Seth, A., and Giorgi, F. (1996). Three-dimensional model study of organized mesoscale circulations induced by vegetation. Journal of Geophysical Research, 101D, 73717391.Google Scholar
Shapiro, A. (2014). A grand experiment: USDA Forest Service experimental forests and ranges. In USDA Forest Service Experimental Forests and Ranges: Research for the Long Term, edited by Hayes, D. C., Stout, S. L., Crawford, R. H., and Hoover, A. P.. New York: Springer, pp. 323.Google Scholar
Shaw, R. H., and Pereira, A. R. (1982). Aerodynamic roughness of a plant canopy: A numerical experiment. Agricultural Meteorology, 26, 5165.Google Scholar
Sheil, D. (2014). How plants water our planet: Advances and imperatives. Trends in Plant Science, 19, 209211.Google Scholar
Sheil, D., and Murdiyarso, D. (2009). How forests attract rain: An examination of a new hypothesis. Bioscience, 59, 341347.Google Scholar
Shugart, H. H. (1984). A Theory of Forest Dynamics: The Ecological Implications of Forest Succession Models. New York: Springer-Verlag.Google Scholar
Shugart, H. H. (1987). Dynamic ecosystem consequences of tree birth and death patterns. BioScience, 37, 596602.Google Scholar
Shugart, H. H. (1998). Terrestrial Ecosystems in Changing Environments. Cambridge, UK: Cambridge University Press.Google Scholar
Shugart, H. H., Leemans, R., and Bonan, G. B. (1992). A Systems Analysis of the Global Boreal Forest. Cambridge, UK: Cambridge University Press.Google Scholar
Shukla, J., and Mintz, Y. (1982). Influence of land-surface evapotranspiration on the Earth’s climate. Science, 215, 14981501.Google Scholar
Shukla, P. R., Skea, J., Slade, R., et al. (2019). Technical summary. In Climate Change and Land: An IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems edited by Shukla, P. R., Skea, J., Buendia, E. Calvo, et al. Geneva: World Meteorological Organization, pp. 3774.Google Scholar
Simmons, C. T., and Matthews, H. D. (2016). Assessing the implications of human land-use change for the transient climate response to cumulative carbon emissions. Environmental Research Letters, 11, 035001, DOI: https://doi.org/10.1088/1748-9326/11/3/035001.Google Scholar
Sitch, S., Cox, P. M., Collins, W. J., and Huntingford, C. (2007). Indirect radiative forcing of climate change through ozone effects on the land-carbon sink. Nature, 448, 791794.Google Scholar
Skea, R. (2013). Vincent’s Trees: Paintings and Drawings by Van Gogh. New York: Thames & Hudson.Google Scholar
Skea, R. (2015). Monet’s Trees: Paintings and Drawings by Claude Monet. New York: Thames & Hudson.Google Scholar
Skinner, C. B., Poulsen, C. J., and Mankin, J. S. (2018). Amplification of heat extremes by plant CO2 physiological forcing. Nature Communications, 9, 1094, DOI: https://doi.org/10.1038/s41467-018-03472-w.Google Scholar
Slinski, K. M., Hogue, T. S., Porter, A. T., and McCray, J. E. (2016). Recent bark beetle outbreaks have little impact on streamflow in the Western United States. Environmental Research Letters, 11, 074010, DOI: https://doi.org/10.1088/1748-9326/11/7/074010.Google Scholar
Sloan, P. R. (1981). Buffon’s preface to the Vegetable Staticks of Stephen Hales (1735). In From Natural History to the History of Nature: Readings from Buffon and His Critics, edited by Lyon, J. and Sloan, P. R.. Notre Dame, IN: University of Notre Dame Press, pp. 3540.Google Scholar
SMIC (1971). Inadvertent Climate Modification: Report of the Study of Man’s Impact on Climate (SMIC). Cambridge, MA: Massachusetts Institute of Technology Press.Google Scholar
Smith, M. B. (1906). The First Forty Years of Washington Society, edited by Hunt, G.. New York: Charles Scribner’s Sons.Google Scholar
Smith, M. N., Stark, S. C., Taylor, T. C., et al. (2019a). Seasonal and drought-related changes in leaf area profiles depend on height and light environment in an Amazon forest. New Phytologist, 222, 12841297.Google Scholar
Smith, P., Adams, J., Beerling, D. J., et al. (2019b). Impacts of land-based greenhouse gas removal options on ecosystem services and the United Nations sustainable development goals. Annual Review of Environment and Resources, 44, 255286.Google Scholar
Smith, P., Arneth, A., Barnes, D. K. A., et al. (2022). How do we best synergize climate mitigation actions to co-benefit biodiversity? Global Change Biology, 28, 25552577.Google Scholar
Smith, P., Davis, S. J., Creutzig, F., et al. (2016). Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6, 4250.Google Scholar
Snyder, P. K. (2010). The influence of tropical deforestation on the Northern Hemisphere climate by atmospheric teleconnections. Earth Interactions, 14(4), 134.Google Scholar
Snyder, P. K., Delire, C., and Foley, J. A. (2004). Evaluating the influence of different vegetation biomes on the global climate. Climate Dynamics, 23, 279302.Google Scholar
Solomon, A. M., West, D. C., and Solomon, J. A. (1981). Simulating the role of climate change and species immigration in forest succession. In Forest Succession: Concepts and Application, edited by West, D. C., Shugart, H. H., and Botkin, D. B.. New York: Springer-Verlag, pp. 154177.Google Scholar
Soltis, D., Soltis, P., Endress, P., et al. (2018). Phylogeny and Evolution of the Angiosperms, revised and updated edition. Chicago: University of Chicago Press.Google Scholar
Song, X.-P., Hansen, M. C., Stehman, S. V., et al. (2018). Global land change from 1982 to 2016. Nature, 560, 639643.Google Scholar
Sonntag, S., Pongratz, J., Reick, C. H., and Schmidt, H. (2016). Reforestation in a high-CO2 world: Higher mitigation potential than expected, lower adaptation potential than hoped for. Geophysical Research Letters, 43, 65466553, DOI: https://doi.org/10.1002/2016GL068824.Google Scholar
Spencer, A. R. T., Mapes, G., Bateman, R. M., Hilton, J., and Rothwell, G. W. (2015). Middle Jurassic evidence for the origin of Cupressaceae: A paleobotanical context for the roles of regulatory genetics and development in the evolution of conifer seed cones. American Journal of Botany, 102, 942961.Google Scholar
Spracklen, D. V., and Garcia-Carreras, L. (2015). The impact of Amazonian deforestation on Amazon basin rainfall. Geophysical Research Letters, 42, 95469552.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, 282285.Google Scholar
Spracklen, D. V., Baker, J. C. A., Garcia-Carreras, L., and Marsham, J. H. (2018). The effects of tropical vegetation on rainfall. Annual Review of Environment and Resources, 43, 193218.Google Scholar
Spracklen, D. V., Bonn, B., and Carslaw, K. S. (2008). Boreal forests, aerosols and the impacts on clouds and climate. Philosophical Transactions of the Royal Society A, 366, 46134626.Google Scholar
Staal, A., Fetzer, I., Wang-Erlandsson, L., et al. (2020). Hysteresis of tropical forests in the 21st century. Nature Communications, 11, 4978, DOI: https://doi.org/10.1038/s41467-020-18728-7.Google Scholar
Staal, A., Tuinenburg, O. A., Bosmans, J. H. C., et al. (2018). Forest-rainfall cascades buffer against drought across the Amazon. Nature Climate Change, 8, 539543.Google Scholar
Stafford, F. (2016). The Long, Long Life of Trees. New Haven, CT: Yale University Press.Google Scholar
Starr, F., Jr. (1866). American forests; their destruction and preservation. In Report of the Commissioner of Agriculture for the Year 1865. Washington, DC: Government Printing Office, pp. 210234.Google Scholar
Staudt, K., Serafimovich, A., Siebicke, L., Pyles, R. D., and Falge, E. (2011). Vertical structure of evapotranspiration at a forest site (a case study). Agricultural and Forest Meteorology, 151, 709729.Google Scholar
Steffen, W., Persson, A., Deutsch, L., et al. (2011). The Anthropocene: From global change to planetary stewardship. Ambio, 40, 739761.Google Scholar
Steffen, W., Richardson, K., Rockström, J., et al. (2020). The emergence and evolution of Earth System Science. Nature Reviews: Earth and Environment, 1, 5463.Google Scholar
Steffen, W., Rockström, J., Richardson, K., et al. (2018). Trajectories of the Earth system in the Anthropocene. Proceedings of the National Academy of Sciences USA, 115, 82528259.Google Scholar
Stegner, W. (1990). It all began with conservation. Smithsonian, 21(1), 3543.Google Scholar
Stehr, N., and von Storch, H. (2000). Eduard Brückner: The Sources and Consequences of Climate Change and Climate Variability in Historical Times. Dordrecht: Kluwer Academic Publishers.Google Scholar
Stein, W. E., Berry, C. M., Hernick, L. V., and Mannolini, F. (2012). Surprisingly complex community discovered in the mid-Devonian fossil forest at Gilboa. Nature, 483, 7881.Google Scholar
Stein, W. E., Berry, C. M., Morris, J. L., et al. (2020). Mid-Devonian Archaeopteris roots signal revolutionary change in earliest fossil forests. Current Biology, 30, 421–431.e2.Google Scholar
Stephens, L., Fuller, D., Boivin, N., et al. (2019). Archaeological assessment reveals Earth’s early transformation through land use. Science, 365, 897902.Google Scholar
Still, C., Powell, R., Aubrecht, D., et al. (2019). Thermal imaging in plant and ecosystem ecology: Applications and challenges. Ecosphere, 10, e02768, DOI: https://doi.org/10.1002/ecs2.2768.Google Scholar
Still, C. J., Rastogi, B., Page, G. F. M., et al. (2021). Imaging canopy temperature: Shedding (thermal) light on ecosystem processes. New Phytologist, 230, 17461753.Google Scholar
Stoy, P. C., Katul, G. G., Siqueira, M. B., et al. (2006). Separating the effects of climate and vegetation on evapotranspiration along a successional chronosequence in the southeastern US. Global Change Biology, 12, 21152135.Google Scholar
Strassberg, R. E. (1994). Inscribed Landscapes: Travel Writing from Imperial China. Berkeley, University of California Press.Google Scholar
Strassburg, B. B. N., Iribarrem, A., Beyer, H. L., et al. (2020). Global priority areas for ecosystem restoration. Nature, 586, 724729.Google Scholar
Stringer, C. (2016). The origin and evolution of Homo sapiens. Philosophical Transactions of the Royal Society B, 371, 20150237, DOI: https://doi.org/10.1098/rstb.2015.0237.Google Scholar
Stringer, C., and Galway-Witham, J. (2017). On the origin of our species. Nature, 546, 212214.Google Scholar
Strutt, J. G. (1822). Sylva Britannica; or, Portraits of Forest Trees, Distinguished for Their Antiquity, Magnitude, or Beauty, folio ed. London: Colnaghi.Google Scholar
Strutt, J. G. (1830). Sylva Britannica; or, Portraits of Forest Trees, Distinguished for Their Antiquity, Magnitude, or Beauty. London: J. G. Strutt.Google Scholar
Stubbe, H. (1667). Observations made by a curious and learned person, sailing from England, to the Caribe-Islands. Philosophical Transactions, 2(27), 494502.Google Scholar
Stuenzi, S. M., and Schaepman-Strub, G. (2020). Vegetation trajectories and shortwave radiative forcing following boreal forest disturbance in eastern Siberia. Journal of Geophysical Research: Biogeosciences, 125, e2019JG005395, DOI: https://doi.org/10.1029/2019JG005395.Google Scholar
Suess, E. (1875). Die Entstehung der Alpen. Vienna: Wilhelm Braumüller.Google Scholar
Surell, A. (1841). Étude sur les torrents des hautes-alpes. Paris: Carilian-Goeury & V. Dalmont.Google Scholar
Swank, W. T., and Douglass, J. E. (1974). Streamflow greatly reduced by converting deciduous hardwood stands to pine. Science, 185, 857859.Google Scholar
Swank, W. T., and Miner, N. H. (1968). Conversion of hardwood-covered watersheds to white pine reduces water yield. Water Resources Research, 4, 947954.Google Scholar
Swank, W. T., Swift, L. W., Jr., and Douglass, J. E. (1988). Streamflow changes associated with forest cutting, species conversions, and natural disturbances. In Forest Hydrology and Ecology at Coweeta, edited by Swank, W. T. and Crossley, D. A., Jr. New York: Springer-Verlag, pp. 297312.Google Scholar
Swann, A. L., Fung, I. Y., Levis, S., Bonan, G. B., and Doney, S. C. (2010). Changes in Arctic vegetation amplify high-latitude warming through the greenhouse effect. Proceedings of the National Academy of Sciences USA, 107, 12951300.Google Scholar
Swann, A. L. S., Fung, I. Y., and Chiang, J. C. H. (2012). Mid-latitude afforestation shifts general circulation and tropical precipitation. Proceedings of the National Academy of Sciences USA, 109, 712716.Google Scholar
Swann, A. L. S., Fung, I. Y., Liu, Y., and Chiang, J. C. H. (2014). Remote vegetation feedbacks and the mid-Holocene green Sahara. Journal of Climate, 27, 48574870.Google Scholar
Swann, A. L. S., Laguë, M. M., Garcia, E. S., et al. (2018). Continental-scale consequences of tree die-offs in North America: Identifying where forest loss matters most. Environmental Research Letters, 13, 055014, DOI: https://doi.org/10.1088/1748-9326/aaba0f.Google Scholar
Swann, A. L. S., Longo, M., Knox, R. G., Lee, E., and Moorcroft, P. R. (2015). Future deforestation in the Amazon and consequences for South American climate. Agricultural and Forest Meteorology, 214-215, 1224.Google Scholar
Tagesson, T., Schurgers, G., Horion, S., et al. (2020). Recent divergence in the contributions of tropical and boreal forests to the terrestrial carbon sink. Nature Ecology and Evolution, 4, 202209.Google Scholar
Tansley, A. G. (1935). The use and abuse of vegetational concepts and terms. Ecology, 16, 284307.Google Scholar
Taylor, C. M., Parker, D. J., and Harris, P. P. (2007). An observational case study of mesoscale atmospheric circulations induced by soil moisture. Geophysical Research Letters, 34, L15801, DOI: https://doi.org/10.1029/2007GL030572.Google Scholar
Ter-Mikaelian, M. T., Colombo, S. J., and Chen, J. (2015). The burning question: Does forest bioenergy reduce carbon emissions? A review of common misconceptions about forest carbon accounting. Journal of Forestry, 113, 5768.Google Scholar
Teuling, A. J. (2018). A forest evapotranspiration paradox investigated using lysimeter data. Vadose Zone Journal, 17, 170031, DOI: https://doi.org/10.2136/vzj2017.01.0031.Google Scholar
Teuling, A. J., Seneviratne, S. I., Stöckli, R., et al. (2010). Contrasting response of European forest and grassland energy exchange to heatwaves. Nature Geoscience, 3, 722727.Google Scholar
Teuling, A. J., Taylor, C. M., Meirink, J. F., et al. (2017). Observational evidence for cloud cover enhancement over western European forests. Nature Communications, 8, 14065, DOI: https://doi.org/10.1038/ncomms14065.Google Scholar
Theophrastus (1990). De causis plantarum, vol. 3. Books 5–6, edited and translated by B. Einarson and G. K. K. Link (Loeb Classical Library 475). Cambridge, MA: Harvard University Press.Google Scholar
Thomas, C. K., Law, B. E., Irvine, J., et al. (2009). Seasonal hydrology explains interannual and seasonal variation in carbon and water exchange in a semiarid mature ponderosa pine forest in central Oregon. Journal of Geophysical Research, 114, G04006, DOI: https://doi.org/10.1029/2009JG001010.Google Scholar
Thomas, W. L., Jr. (1956). Introductory: About the symposium, about the people, about the theme. In Man’s Role in Changing the Face of the Earth, edited by Thomas, W. L., Jr. Chicago: University of Chicago Press, pp. xxixxxviii.Google Scholar
Thompson, J. R., Carpenter, D. N., Cogbill, C. V., and Foster, D. R. (2013). Four centuries of change in northeastern United States forests. PLoS ONE, 8(9), e72540, DOI: https://doi.org/10.1371/journal.pone.0072540.Google Scholar
Thompson, K. (1980). Forests and climate change in America: Some early views. Climatic Change, 3, 4764.Google Scholar
Thompson, K. (1981). The question of climatic stability in America before 1900. Climatic Change, 3, 227241.Google Scholar
Thompson, M. P., Adams, D., and Sessions, J. (2009). Radiative forcing and the optimal rotation age. Ecological Economics, 68, 27132720.Google Scholar
Thompson, O. E., and Pinker, R. T. (1975). Wind and temperature profile characteristics in a tropical evergreen forest in Thailand. Tellus, 27, 562573.Google Scholar
Thoreau, H. D. (1863a). The succession of forest trees. In Excursions. Boston: Ticknor & Fields, pp. 135160.Google Scholar
Thoreau, H. D. (1863b). Autumnal tints. In Excursions. Boston: Ticknor & Fields, pp. 215265.Google Scholar
Thoreau, H. D. (1906). The Writings of Henry David Thoreau: Journal, edited by Torrey, B., 14 vols. Boston: Houghton Mifflin.Google Scholar
Thoreau, H. D. (2009). The Maine Woods: A Fully Annotated Edition, edited by Cramer, J. S.. New Haven, CT: Yale University Press.Google Scholar
Thornhill, G. D., Ryder, C. L., Highwood, E. J., Shaffrey, L. C., and Johnson, B. T. (2018). The effect of South American biomass burning aerosol emissions on the regional climate. Atmospheric Chemistry and Physics, 18, 53215342.Google Scholar
Thornthwaite, C. W. (1956). Modification of rural microclimates. In Man’s Role in Changing the Face of the Earth, edited by Thomas, W. L., Jr. Chicago: University of Chicago Press, pp. 567583.Google Scholar
Thornton, P. E., Doney, S. C., Lindsay, K., et al. (2009). Carbon–nitrogen interactions regulate climate–carbon cycle feedbacks: Results from an atmosphere–ocean general circulation model. Biogeosciences, 6, 20992120.Google Scholar
Thwaites, R. G. (1897a). The Jesuit Relations and Allied Documents: Travels and Explorations of the Jesuit Missionaries in New France, 1610–1791, vol. 3. Acadia: 1611–1616. Cleveland: Burrows.Google Scholar
Thwaites, R. G. (1897b). The Jesuit Relations and Allied Documents: Travels and Explorations of the Jesuit Missionaries in New France, 1610–1791, vol. 5. Quebec: 1632–1633. Cleveland: Burrows.Google Scholar
Tierney, J. E., Pausata, F. S. R., deMenocal, P. B. (2017). Rainfall regimes of the Green Sahara. Science Advances, 3, e1601503, DOI: https://doi.org/10.1126/sciadv.1601503.Google Scholar
Tilley, M. P. (1950). A Dictionary of the Proverbs in England in the Sixteenth and Seventeenth Centuries. Ann Arbor: University of Michigan Press.Google Scholar
Tilmann, J. P. (1971) An Appraisal of the Geographical Works of Albertus Magnus and His Contributions to Geographical Thought (Michigan Geographical Publication No. 4). Ann Arbor: Department of Geography, University of Michigan.Google Scholar
Todd, D., and Abbe, C. (1891). Additional results of the United States scientific expedition to West Africa. Nature, 43, 563565.Google Scholar
Totman, C. (1989). The Green Archipelago: Forestry in Preindustrial Japan. Berkeley: University of California Press.Google Scholar
Travassos-Britto, B., Pardini, R., El-Hani, C. N., and Prado, P. I. (2021). Towards a pragmatic view of theories in ecology. Oikos, 130, 821830.Google Scholar
Travers, W. T. L. (1870). On the changes effected in the natural features of a new country by the introduction of civilized races. Part III. Transactions and Proceedings of the New Zealand Institute, 3, 326336.Google Scholar
Tristram, H. B. (1868). The Natural History of the Bible: Being a Review of the Physical Geography, Geology, and Meteorology of the Holy Land, 2nd ed. London: Society for Promoting Christian Knowledge.Google Scholar
Tunved, P., Hansson, H.-C., Kerminen, V.-M., et al. (2006). High natural aerosol loading over boreal forests. Science, 312, 261263.Google Scholar
Turner, O. (1849). Pioneer History of the Holland Purchase of Western New York. Buffalo, NY: Jewett, Thomas; George H. Derby.Google Scholar
Twohy, C. H., Toohey, D. W., Levin, E. J. T., et al. (2021). Biomass burning smoke and its influence on clouds over the western U.S. Geophysical Research Letters, 48, e2021GL094224, DOI: https://doi.org/10.1029/2021GL094224.Google Scholar
Tyndall, J. (1863). On radiation through the earth’s atmosphere. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, series 4, 25, 200206.Google Scholar
Unger, N. (2013). Isoprene emission variability through the twentieth century. Journal of Geophysical Research: Atmospheres, 118, 1360613613.Google Scholar
Unger, N. (2014). Human land-use-driven reduction of forest volatiles cools global climate. Nature Climate Change, 4, 907910.Google Scholar
Uno, K. T., Polissar, P. J., Jackson, K. E., and deMenocal, P. B. (2016). Neogene biomarker record of vegetation change in eastern Africa. Proceedings of the National Academy of Sciences USA, 113, 63556363.Google Scholar
Urbanski, S., Barford, C., Wofsy, S., et al. (2007). Factors controlling CO2 exchange on timescales from hourly to decadal at Harvard Forest. Journal of Geophysical Research, 112, G02020, DOI: https://doi.org/10.1029/2006JG000293.Google Scholar
Ure, A. (1821). A Dictionary of Chemistry. London: Thomas & George Underwood; J. Highley & Son; and others.Google Scholar
Van Bueren, G. (2015). More Magna than Magna Carta: Magna Carta’s sister – the Charter of the Forest. In Magna Carta and Its Modern Legacy, edited by Hazell, R. and Melton, J.. New York: Cambridge University Press, pp. 194211.Google Scholar
Van Cleve, K., and Viereck, L. A. (1981). Forest succession in relation to nutrient cycling in the boreal forest of Alaska. In Forest Succession: Concepts and Application, edited by West, D. C., Shugart, H. H., and Botkin, D. B.. New York: Springer-Verlag, pp. 185211.Google Scholar
Van Cleve, K., Chapin, F. S., III, Flanagan, P. W., Viereck, L. A., and Dyrness, C. T. (1986). Forest Ecosystems in the Alaskan Taiga: A Synthesis of Structure and Function. New York: Springer-Verlag.Google Scholar
Van Cleve, K., Dyrness, C. T., Viereck, L. A., et al. (1983). Taiga ecosystems in interior Alaska. BioScience, 33, 3944.Google Scholar
van der Bles, A. M., van der Linden, S., Freeman, A. L. J., and Spiegelhalter, D. J. (2020). The effects of communicating uncertainty on public trust in facts and numbers. Proceedings of the National Academy of Sciences USA, 117, 76727683.Google Scholar
Vanderhoof, M. K., and Williams, C. A. (2015). Persistence of MODIS evapotranspiration impacts from mountain pine beetle outbreaks in lodgepole pine forests, south-central Rocky Mountains. Agricultural and Forest Meteorology, 200, 7891.Google Scholar
Vanderhoof, M., Williams, C. A., Ghimire, B., and Rogan, J. (2013). Impact of mountain pine beetle outbreaks on forest albedo and radiative forcing, as derived from Moderate Resolution Imaging Spectroradiometer, Rocky Mountains, USA. Journal of Geophysical Research: Biogeosciences, 118, 14611471.Google Scholar
Vanderhoof, M., Williams, C. A., Shuai, Y., et al. (2014). Albedo-induced radiative forcing from mountain pine beetle outbreaks in forests, south-central Rocky Mountains: Magnitude, persistence, and relation to outbreak severity. Biogeosciences, 11, 563575.Google Scholar
van der Werf, G. R., Randerson, J. T., Giglio, L., et al. (2017). Global fire emissions estimates during 1997–2016. Earth System Science Data, 9, 697720.Google Scholar
Vargas Zeppetello, L. R., Cook-Patton, S. C., Parsons, L. A., et al. (2022). Consistent cooling benefits of silvopasture in the tropics. Nature Communications, 13, 708, DOI: https://doi.org/10.1038/s41467-022-28388-4.Google Scholar
Vargas Zeppetello, L. R., Parsons, L. A., Spector, J. T., et al. (2020). Large scale tropical deforestation drives extreme warming. Environmental Research Letters, 15, 084012, DOI: https://doi.org/10.1088/1748-9326/ab96d2.Google Scholar
Venäläinen, A., Lehtonen, I., Laapas, M., et al. (2020). Climate change induces multiple risks to boreal forests and forestry in Finland: A literature review. Global Change Biology, 26, 41784196.Google Scholar
Verkerk, P. J., Costanza, R., Hetemäki, L., et al. (2020). Climate-Smart Forestry: The missing link. Forest Policy and Economics, 115, 102164, DOI: https://doi.org/10.1016/j.forpol.2020.102164.Google Scholar
Vernadsky, V. I. (1998). The Biosphere, forward by L. Margulis and colleagues; introduction by J. Grinevald; translated by D. B. Langmuir; revised and annotated by M. A. S. McMenamin. New York: Springer.Google Scholar
Vincent, P. (1637). A True Relation of the Late Battell Fought in New England, between the English, and the Salvages: With the Present State of Things There. London: Nathanael Butter & John Bellamie (Text Creation Partnership, Ann Arbor, Michigan; http://name.umdl.umich.edu/A14439.0001.001).Google Scholar
Vitruvius (1914). The Ten Books on Architecture, translated by M. H. Morgan. Cambridge, MA: Harvard University Press.Google Scholar
Voeikov, A. (1885a). Der Einfluss der Wälder auf das Klima. Petermanns geographische Mitteilungen, 31, 8187.Google Scholar
Voeikov, A. (1885b). Die Regenverhältnisse des malayischen Archipels. Zeitschrift der österreichischen Gesellschaft für Meteorologie, 20, 113138, 201211, 250263.Google Scholar
Voeikov, A. (1886). On the influence of forests upon climate. Quarterly Journal of the Royal Meteorological Society, 12, 2638.Google Scholar
Voeikov, A. (1887). Die Klimate der Erde, 2 vols. Jena, Germany: Hermann Costenoble.Google Scholar
Voeikov, A. (1878). Einfluss der Wälder und der Irrigation auf das Klima. Zeitschrift der Österreichischen Gesellschaft für Meteorologie, 13, 4748.Google Scholar
Voeikov, A. (1888). Klimatologische Zeit- und Streitfragen. Meteorologische Zeitschrift, 5, 1721, 191195.Google Scholar
Voeikov, A. (1901). De l’influence de l’homme sur la terre. Annales de géographie, 10, 97114, 193215.Google Scholar
Vogel, B. (2011). The letter from Dublin: Climate change, colonialism, and the Royal Society in the seventeenth century. Osiris, 26, 111128.Google Scholar
Volney, C.-F. (1804). View of the Climate and Soil of the United States of America. London: J. Johnson.Google Scholar
von Arx, G., Dobbertin, M., and Rebetez, M. (2012). Spatio-temporal effects of forest canopy on understory microclimate in a long-term experiment in Switzerland. Agricultural and Forest Meteorology, 166–167, 14455.Google Scholar
von Arx, G., Pannatier, E. G., Thimonier, A., and Rebetez, M. (2013). Microclimate in forests with varying leaf area index and soil moisture: Potential implications for seedling establishment in a changing climate. Journal of Ecology, 101, 12011213.Google Scholar
von Randow, C., Manzi, A. O., Kruijt, B., et al. (2004). Comparative measurements and seasonal variations in energy and carbon exchange over forest and pasture in South West Amazonia. Theoretical and Applied Climatology, 78, 526.Google Scholar
Walker, A. P., De Kauwe, M. G., Bastos, A., et al. (2021). Integrating the evidence for a terrestrial carbon sink caused by increasing atmospheric CO2. New Phytologist, 229, 24132445.Google Scholar
Walker, X. J., Baltzer, J. L., Cumming, S. G., et al. (2019). Increasing wildfires threaten historic carbon sink of boreal forest soils. Nature, 572, 520523.Google Scholar
Wang, G., and Eltahir, E. A. B. (2000a). Biosphere–atmosphere interactions over West Africa. II: Multiple climate equilibria. Quarterly Journal of the Royal Meteorological Society, 126, 12611280.Google Scholar
Wang, G., and Eltahir, E. A. B. (2000b). Ecosystem dynamics and the Sahel drought. Geophysical Research Letters, 27, 795798.Google Scholar
Wang, G., and Eltahir, E. A. B. (2000c). Role of vegetation dynamics in enhancing the low-frequency variability of the Sahel rainfall. Water Resources Research, 36, 10131021.Google Scholar
Wang, G., Eltahir, E. A. B., Foley, J. A., Pollard, D., and Levis, S. (2004). Decadal variability of rainfall in the Sahel: Results from the coupled GENESIS–IBIS atmosphere–biosphere model. Climate Dynamics, 22, 625637.Google Scholar
Wang, G., Yu, M., and Xue, Y. (2016). Modeling the potential contribution of land cover changes to the late twentieth century Sahel drought using a regional climate model: Impact of lateral boundary conditions. Climate Dynamics, 47, 34573477.Google Scholar
Wang, J., Chagnon, F. J. F., Williams, E. R., et al. (2009). Impact of deforestation in the Amazon basin on cloud climatology. Proceedings of the National Academy of Sciences USA, 106, 36703674.Google Scholar
Wang, J. A., Baccini, A., Farina, M., Randerson, J. T., and Friedl, M. A. (2021). Disturbance suppresses the aboveground carbon sink in North American boreal forests. Nature Climate Change, 11, 435441.Google Scholar
Warde, P. (2006). Fear of wood shortage and the reality of the woodland in Europe, c.1450–1850. History Workshop Journal, 62 (Autumn), 2857.Google Scholar
Warde, P. (2018). The Invention of Sustainability: Nature and Destiny, c. 1500–1870. Cambridge, UK: Cambridge University Press.Google Scholar
Watson, J. E. M., Evans, T., Venter, O., et al. (2018). The exceptional value of intact forest ecosystems. Nature Ecology and Evolution, 2, 599610.Google Scholar
Watson, W. C. (1866). Forests – their influence, uses and reproduction. In Transactions of the New York State Agricultural Society for the Year 1865, vol. 25. Albany, NY: Cornelius Wendell, pp. 288303.Google Scholar
Watts, F. (1872). Report of the Commissioner of Agriculture for the Year 1871. Washington, DC: Government Printing Office.Google Scholar
Watts, F. (1874). Report of the Commissioner of Agriculture for the Year 1872. Washington, DC: Government Printing Office.Google Scholar
Webb, T., III, Bartlein, P. J., Harrison, S. P., and Anderson, K. H. (1993). Vegetation, lake levels, and climate in eastern North America for the past 18,000 years. In Global Climates since the Last Glacial Maximum, edited by Wright, H. E., Jr., Kutzbach, J. E., Webb, T., III, et al. Minneapolis: University of Minnesota Press, pp. 415467.Google Scholar
Webster, N. (1790). A Collection of Essays and Fugitiv Writings: On Moral, Historical, Political and Literary Subjects. Boston: I. Thomas & E. T. Andrews.Google Scholar
Webster, N. (1799). On the effects of evergreens on climate. Transactions of the Society for the Promotion of Agriculture, Arts, and Manufactures, Instituted in the State of New York, 1(4), 5152.Google Scholar
Webster, N. (1809). Experiments respecting dew, intended to ascertain whether dew is the descent of vapour during the night, or the perspiration of the earth, or of plants; or whether it is not the effect of condensation. Memoirs of the American Academy of Arts and Sciences, 3(1), 95103.Google Scholar
Webster, N. (1810). A dissertation on the supposed change in the temperature of winter. Memoirs of the Connecticut Academy of Arts and Sciences, 1(1), 168.Google Scholar
Webster, W. H. B. (1834). Narrative of a Voyage to the Southern Atlantic Ocean, in the Years 1828, 29, 30, Performed in H. M. Sloop Chanticleer, 2 vols. London: Richard Bentley.Google Scholar
Weld, I. (1799). Travels through the States of North America, and the Provinces of Upper and Lower Canada, during the Years 1795, 1796, and 1797. London: John Stockdale.Google Scholar
Weppelmann, S. (2019). Introduction: Bruegel between 2019 and 2069. In Bruegel: The Hand of the Master. The 450th Anniversary Edition – Essays in Context, edited by Hoppe-Harnoncourt, A., Oberthaler, E., Pénot, S., Sellink, M., and Spronk, R.. Vienna: KHM-Museumsverband, pp. 89.Google Scholar
Werth, D., and Avissar, R. (2002). The local and global effects of Amazon deforestation. Journal of Geophysical Research, 107, 8087, DOI: https://doi.org/10.1029/2001JD000717.Google Scholar
West, D. C., Shugart, H. H., and Botkin, D. B. (1981). Forest Succession: Concepts and Application. New York: Springer-Verlag.Google Scholar
Wex, G. (1873). Ueber die Wasserabnahme in den Quellen, Flüssen und Strömen. Zeitschrift des oesterreichischen Ingenieur- und Architekten-Vereins, 25, 2330, 6376, 101119.Google Scholar
Wex, G. (1880). Second Treatise on the Decrease of Water in Springs, Creeks, and Rivers, Contemporaneously with an Increase in Height of Floods in Cultivated Countries, translated by G. Weitzel. Washington, DC: Government Printing Office.Google Scholar
Wex, G. (1881). First Treatise on the Decrease of Water in Springs, Creeks, and Rivers, Contemporaneously with an Increase in Height of Floods in Cultivated Countries, translated by Weitzel, G.. Washington, DC: Government Printing Office.Google Scholar
Whitbourne, R. (1620). A Discourse and Discovery of New-Found-Land. London: William Barret.Google Scholar
White, G. (1789). The Natural History and Antiquities of Selborne, in the County of Southampton. London: T. Bensley.Google Scholar
Whitney, G. G. (1994). From Coastal Wilderness to Fruited Plain: A History of Environmental Change in Temperate North America, 1500 to the Present. Cambridge, UK: Cambridge University Press.Google Scholar
Whittaker, R. H. (1956). Vegetation of the Great Smoky Mountains. Ecological Monographs, 26, 180.Google Scholar
Whittaker, R. H. (1975). Communities and Ecosystems, 2nd ed. New York: MacMillan.Google Scholar
Whittaker, R. H., Bormann, F. H., Likens, G. E., and Siccama, T. G. (1974). The Hubbard Brook Ecosystem Study: Forest biomass and production. Ecological Monographs, 44, 233252.Google Scholar
Wieder, W. R., Cleveland, C. C., Smith, W. K., and Todd-Brown, K. (2015). Future productivity and carbon storage limited by terrestrial nutrient availability. Nature Geoscience, 8, 441444.Google Scholar
Wilber, C. D. (1881). The Great Valleys and Prairies of Nebraska and the Northwest. Omaha: Daily Republican Print.Google Scholar
Williams, A. P., Abatzoglou, J. T., Gershunov, A., et al. (2019). Observed impacts of anthropogenic climate change on wildfire in California. Earth’s Future, 7, 892910.Google Scholar
Williams, C. A., Gu, H., and Jiao, T. (2021). Climate impacts of US forest loss span net warming to net cooling. Science Advances, 7, eaax8859, DOI: https://doi.org/10.1126/sciadv.aax8859.Google Scholar
Williams, C. A., Reichstein, M., Buchmann, N., et al. (2012). Climate and vegetation controls on the surface water balance: Synthesis of evapotranspiration measured across a global network of flux towers. Water Resources Research, 48, W06523, DOI: https://doi.org/10.1029/2011WR011586.Google Scholar
Williams, C. J., Johnson, A. H., LePage, B. A., Vann, D. R., and Sweda, T. (2003). Reconstruction of Tertiary Metasequoia forests. II. Structure, biomass, and productivity of Eocene floodplain forests in the Canadian Arctic. Paleobiology, 29, 271292.Google Scholar
Williams, E., Rosenfeld, D., Madden, N., et al. (2002). Contrasting convective regimes over the Amazon: Implications for cloud electrification. Journal of Geophysical Research, 107D, 8082, DOI: https://doi.org/10.1029/2001JD000380.Google Scholar
Williams, J. W., Shuman, B. N., Webb, T., III, Bartlein, P. J., and Leduc, P. L. (2004). Late-Quaternary vegetation dynamics in North America: Scaling from taxa to biomes. Ecological Monographs, 74, 309334.Google Scholar
Williams, M. (1989). Americans and Their Forests: A Historical Geography. Cambridge, UK: Cambridge University Press.Google Scholar
Williams, M. (2003). Deforesting the Earth: From Prehistory to Global Crisis. Chicago: University of Chicago Press.Google Scholar
Williams, S. (1786). Experiments on evaporation, and meteorological observations made at Bradford in New-England, in 1772. Transactions of the American Philosophical Society, 2, 118141.Google Scholar
Williams, S. (1794). The Natural and Civil History of Vermont. Walpole, NH: Isaiah Thomas & David Carlisle.Google Scholar
Williamson, H. (1771). An attempt to account for the change of climate, which has been observed in the middle colonies in North-America. Transactions of the American Philosophical Society, 1, 272280.Google Scholar
Williamson, H. (1773). Dans lequel on tâche de rendre raison du changement de climat qu’on a observé dans les colonies situées dans l’intérieur des terres de l’Amérique septentrionale. Observations sur la physique, sur l’histoire naturelle et sur les arts, 1, 430436.Google Scholar
Williamson, H. (1811). Observations on the Climate in Different Parts of America, Compared with the Climate in Corresponding Parts of the Other Continent. New York: T. & J. Swords.Google Scholar
Willis, E. P., and Hooke, W. H. (2006). Cleveland Abbe and American meteorology, 1871–1901. Bulletin of the American Meteorological Society, 87, 315326.Google Scholar
Wilson, H. M. (1898). The relation of forestation to water-supply. The Engineering Magazine 14(5), 807816.Google Scholar
Wilson, J. F. (1865a). Water supply in the basin of the River Orange, or ’Gariep, South Africa. Journal of the Royal Geographical Society, 35, 106129.Google Scholar
Wilson, J. F. (1865b). On the progressing desiccation of the basin of the Orange River in Southern Africa. Proceedings of the Royal Geographical Society, 9, 106109.Google Scholar
Wilson, J. P., Montañez, I. P., White, J. D., et al. (2017). Dynamic Carboniferous tropical forests: New views of plant function and potential for physiological forcing of climate. New Phytologist, 215, 13331353.Google Scholar
Wilson, J. S. (1867). Report of the Commissioner of General Land Office, for the Year 1867. Washington, DC: Government Printing Office.Google Scholar
Wilson, J. S. (1868). Report of the Commissioner of General Land Office for the Year 1868. Washington, DC: Government Printing Office.Google Scholar
Winckler, J., Lejeune, Q., Reick, C. H., and Pongratz, J. (2019a). Nonlocal effects dominate the global mean surface temperature response to the biogeophysical effects of deforestation. Geophysical Research Letters, 46, 745755.Google Scholar
Winckler, J., Reick, C. H., Bright, R. M., and Pongratz, J. (2019b). Importance of surface roughness for the local biogeophysical effects of deforestation. Journal of Geophysical Research: Atmospheres, 124, 86058618.Google Scholar
Winckler, J., Reick, C. H., Luyssaert, S., et al. (2019c). Different response of surface temperature and air temperature to deforestation in climate models. Earth System Dynamics, 10, 473484.Google Scholar
Windisch, M. G., Davin, E. L., and Seneviratne, S. I. (2021). Prioritizing forestation based on biogeochemical and local biogeophysical impacts. Nature Climate Change, 11, 867871.Google Scholar
Windisch, M. G., Humpenöder, F., Lejeune, Q., et al. (2022). Accounting for local temperature effect substantially alters afforestation patterns. Environmental Research Letters, 17, 024030, DOI: https://doi.org/10.1088/1748-9326/ac4f0e.Google Scholar
Wood, T. (1875). Should the farmers of America oppose further the destruction of our forests? In Report of the Transactions of the Pennsylvania State Agricultural Society, for the Years 1874–75, vol. 10. Harrisburg: B. F. Meyers, pp. 153155.Google Scholar
Woodward, J. (1699). Some thoughts and experiments concerning vegetation. Philosophical Transactions, 21(253), 193227.Google Scholar
Worden, S., Fu, R., Chakraborty, S., Liu, J., and Worden, J. (2021). Where does moisture come from over the Congo Basin? Journal of Geophysical Research: Biogeosciences, 126, e2020JG006024, DOI: https://doi.org/10.1029/2020JG006024.Google Scholar
Wright, I. J., Reich, P. B., Westoby, M., et al. (2004). The worldwide leaf economics spectrum. Nature, 428, 821827.Google Scholar
Wright, J. S., Fu, R., Worden, J. R., et al. (2017). Rainforest-initiated wet season onset over the southern Amazon. Proceedings of the National Academy of Sciences USA, 114, 84818486.Google Scholar
Wulf, A. (2015). The Invention of Nature: Alexander von Humboldt’s New World. New York: Alfred A. Knopf.Google Scholar
Xu, L., Saatchi, S. S., Yang, Y., et al. (2021). Changes in global terrestrial live biomass over the 21st century. Science Advances, 7, eabe9829, DOI: https://doi.org/10.1126/sciadv.abe9829.Google Scholar
Xu, X., Riley, W. J., Koven, C. D., Jia, G., and Zhang, X. (2020). Earlier leaf-out warms air in the north. Nature Climate Change, 10, 370375.Google Scholar
Xue, Y. (2006). Interactions and feedbacks between climate and dryland vegetations. In Dryland Ecohydrology, edited by D’Odorico, P. and Porporato, A.. Dordrecht: Springer, pp. 85105.Google Scholar
Xue, Y., and Shukla, J. (1993). The influence of land surface properties on Sahel climate. Part I: Desertification. Journal of Climate, 6, 22322245.Google Scholar
Xue, Y., and Shukla, J. (1996). The influence of land surface properties on Sahel climate. Part II: Afforestation. Journal of Climate, 9, 32603275.Google Scholar
Xue, Y., Hutjes, R. W. A., Harding, R. J., et al. (2004). The Sahelian climate. In Vegetation, Water, Humans and the Climate: A New Perspective on an Interactive System, edited by Kabat, P., Claussen, M., Dirmeyer, P. A., et al. Berlin: Springer, pp. 5977.Google Scholar
Yang, Y., Saatchi, S. S., Xu, L., et al. (2018). Post-drought decline of the Amazon carbon sink. Nature Communications, 9, 3172, DOI: https://doi.org/10.1038/s41467-018-05668-6.Google Scholar
Yli-Juuti, T., Mielonen, T., Heikkinen, L., et al. (2021). Significance of the organic aerosol driven climate feedback in the boreal area. Nature Communications, 12, 5637, DOI: https://doi.org/10.1038/s41467-021-25850-7.Google Scholar
Yosef, G., Walko, R., Avisar, R., et al. (2018). Large-scale semi-arid afforestation can enhance precipitation and carbon sequestration potential. Scientific Reports, 8, 996, DOI: https://doi.org/10.1038/s41598-018-19265-6.Google Scholar
Young, C. R. (1979). The Royal Forests of Medieval England. Philadelphia: University of Pennsylvania Press.Google Scholar
Yousefpour, R., Augustynczik, A. L. D., Reyer, C. P. O., et al. (2018). Realizing mitigation efficiency of European commercial forests by climate smart forestry. Scientific Reports, 8, 345, DOI: https://doi.org/10.1038/s41598-017-18778-w.Google Scholar
Yu, L., Liu, Y., Liu, T., and Yan, F. (2020). Impact of recent vegetation greening on temperature and precipitation over China. Agricultural and Forest Meteorology, 295, 108197, DOI: https://doi.org/10.1016/j.agrformet.2020.108197.Google Scholar
Zaehle, S., Friend, A. D., Friedlingstein, P., et al. (2010). Carbon and nitrogen cycle dynamics in the O-CN land surface model: 2. Role of the nitrogen cycle in the historical terrestrial carbon balance. Global Biogeochemical Cycles, 24, GB1006, DOI: https://doi.org/10.1029/2009GB003522.Google Scholar
Zaehle, S., Medlyn, B. E., De Kauwe, M. G., et al. (2014). Evaluation of 11 terrestrial carbon–nitrogen cycle models against observations from two temperate Free-Air CO2 Enrichment studies. New Phytologist, 202, 803822.Google Scholar
Zaitchik, B. F., Macalady, A. K., Bonneau, L. R., and Smith, R. B. (2006). Europe’s 2003 heat wave: A satellite view of impacts and land-atmosphere feedbacks. International Journal of Climatology, 26, 743769.Google Scholar
Zarakas, C. M., Swann, A. L. S., Laguë, M. M., Armour, K. C., and Randerson, J. T. (2020). Plant physiology increases the magnitude and spread of the transient climate response to CO2 in CMIP6 earth system models. Journal of Climate, 33, 85618577.Google Scholar
Zemp, D. C., Schleussner, C.-F., Barbosa, H. M. J., et al. (2017). Self-amplified Amazon forest loss due to vegetation-atmosphere feedbacks. Nature Communications, 8, 14681, DOI: https://doi.org/10.1038/ncomms14681.Google Scholar
Zeng, N., and Yoon, J. (2009). Expansion of the world’s deserts due to vegetation-albedo feedback under global warming. Geophysical Research Letters, 36, L17401, DOI: https://doi.org/10.1029/2009GL039699.Google Scholar
Zeng, N., Neelin, J. D., Lau, K.-M., and Tucker, C. J. (1999). Enhancement of interdecadal climate variability in the Sahel by vegetation interaction. Science, 286, 15371540.Google Scholar
Zeng, Z., Piao, S., Li, L. Z. X., et al. (2017). Climate mitigation from vegetation biophysical feedbacks during the past three decades. Nature Climate Change, 7, 432436.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, 701708.Google Scholar
Zhang, M., Lee, X., Yu, G., et al. (2014). Response of surface air temperature to small-scale land clearing across latitudes. Environmental Research Letters, 9, 034002, DOI: https://doi.org/10.1088/1748-9326/9/3/034002.Google Scholar
Zhang, M., Liu, N., Harper, R., et al. (2017). A global review on hydrological responses to forest change across multiple spatial scales: Importance of scale, climate, forest type and hydrological regime. Journal of Hydrology, 546, 4459.Google Scholar
Zhang, Q., Barnes, M., Benson, M., et al. (2020a). Reforestation and surface cooling in temperate zones: Mechanisms and implications. Global Change Biology, 26, 33843401.Google Scholar
Zhang, X., Du, J., Zhang, L., et al. (2020b). Impact of afforestation on surface ozone in the North China Plain during the three-decade period. Agricultural and Forest Meteorology, 287, 107979, DOI: https://doi.org/10.1016/j.agrformet.2020.107979.Google Scholar
Zhang, X., Huang, T., Zhang, L., et al. (2016a). Three-North Shelter Forest Program contribution to long-term increasing trends of biogenic isoprene emissions in northern China. Atmospheric Chemistry and Physics, 16, 69496960.Google Scholar
Zhang, Y., Peng, C., Li, W., et al. (2016b). Multiple afforestation programs accelerate the greenness in the “Three North” region of China from 1982 to 2013. Ecological Indicators, 61, 404412.Google Scholar
Zhang, Z., Li, X., and Liu, H. (2022). Biophysical feedback of forest canopy height on land surface temperature over contiguous United States. Environmental Research Letters, 17, 034002, DOI: https://doi.org/10.1088/1748-9326/ac4657.Google Scholar
Zhang, Z., Zhang, F., Wang, L., Lin, A., and Zhao, L. (2021). Biophysical climate impact of forests with different age classes in mid- and high-latitude North America. Forest Ecology and Management, 494, 119327, DOI: https://doi.org/10.1016/j.foreco.2021.119327.Google Scholar
Zhao, K., and Jackson, R. B. (2014). Biophysical forcings of land-use changes from potential forestry activities in North America. Ecological Monographs, 84, 329353.Google Scholar
Zhu, Z., Piao, S., Myneni, R. B., et al. (2016). Greening of the Earth and its drivers. Nature Climate Change, 6, 791795.Google Scholar
Zhuang, Y., Fu, R., Santer, B. D., Dickinson, R. E., and Hall, A. (2021). Quantifying contributions of natural variability and anthropogenic forcings on increased fire weather risk over the western United States. Proceedings of the National Academy of Sciences USA, 118, e2111875118, DOI: https://doi.org/10.1073/pnas.2111875118.Google Scholar
Zilberstein, A. (2016). A Temperate Empire: Making Climate Change in Early America. Oxford: Oxford University Press.Google Scholar
Zipes, J. (1988). The Brothers Grimm: From Enchanted Forests to the Modern World. New York: Routledge.Google Scholar
Zon, R. (1912). Forests and water in the light of scientific investigation. In Final Report of the National Waterways Commission. Washington, DC: Government Printing Office, pp. 203302.Google Scholar
Zon, R. (1913). The relation of forests in the Atlantic Plain to the humidity of the central states and prairie region. Proceedings of the Society of American Foresters, 8, 139153.Google Scholar
Zötl, G. von (1831). Handbuch der Forstwirthschaft im Hochgebirge. Vienna: Carl Gerold.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
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Seeing the Forest for the Trees
  • Online publication: 02 February 2023
  • Chapter DOI: https://doi.org/10.1017/9781108601559.021
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
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Seeing the Forest for the Trees
  • Online publication: 02 February 2023
  • Chapter DOI: https://doi.org/10.1017/9781108601559.021
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
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Seeing the Forest for the Trees
  • Online publication: 02 February 2023
  • Chapter DOI: https://doi.org/10.1017/9781108601559.021
Available formats
×