Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-9q27g Total loading time: 0 Render date: 2024-07-19T17:26:18.749Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 March 2016

Axel Kleidon
Affiliation:
Max-Planck-Institut für Biogeochemie, Jena
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
Publisher: Cambridge University Press
Print publication year: 2016

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aiello, L. C., and Wheeler, P. 1995. The expensive-tissue hypothesis: the brain and the digestive system in human and primate evolution. Curr. Anthropol., 36, 199–221.Google Scholar
Allen, M. R., and Ingram, W. J. 2002. Constraints on future changes in climate and the hydrologic cycle. Nature, 419, 224–232.Google Scholar
Allman, J. M. 1999. Evolving Brains. New York: Scientific American Library.
Ambaum, M. H. P. 2010. Thermal Physics of the Atmosphere. Chichester: J. Wiley and Sons.
Amend, J. P., and Shock, E. L. 2001. Energetics of overall metabolic reactions of thermophilic and hyperthermophilic Archea and Bacteria. FEMS Microbiol. Rev., 25, 175–243.Google Scholar
Amthor, J. S. 1995. Terrestrial higher-plant response to increasing atmospheric CO2 in relation to the global carbon cycle. Glob. Ch. Biol., 1, 243–274.Google Scholar
Amthor, J. S. 2010. From sunlight to phytomass: on the potential efficiency of converting solar radiation to phyto-energy. New Phytol., 188, 939–959.Google Scholar
Andreas, E. L. 2011. Fallacies of the enthalpy transfer coefficient over the ocean in high winds. J. Atmos. Sci., 68, 1435–1445.Google Scholar
Andresen, B., Berry, R. S., Nitzan, A., and Salamon, P. 1977. Thermodynamics in finite time. I: the step-Carnot cycle. Phys. Rev. A, 15, 2086–2093.Google Scholar
Andresen, B., Salamon, P., and Berry, R. S. 1984. Thermodynamics in finite time. Phys. Today, 37, 62–70.Google Scholar
Aoki, I. 1983. Entropy productions on the Earth and other planets of the solar system. J. Phys. Soc. Japan, 52, 1075–1078.Google Scholar
Archer, C. L., and Caldeira, K. 2009. Global assessment of high-altitude wind power. Energies, 2, 307–319.Google Scholar
Arens, S., and Kleidon, A. 2008. Global sensitivity of weathering rates to atmospheric CO2 under the assumption of saturated river discharge. Mineral. Mag., 72, 301–304.Google Scholar
Arya, S. P. 1998. Introduction to Micrometeorology. San Diego, CA: Academic Press.
Ashton, K. G., Tracy, M. C., and de Queiroz, A. 2000. Is Bergmann's rule valid for mammals? Am. Nat., 156, 390–415.Google Scholar
Atkins, P, and de Paula, J. 2010. Physical Chemistry. 9th edn. Oxford and New York: Oxford University Press.
Ayres, R., and Kneese, A. 1969. Production, consumption, and externalities. Am. Econ. Rev., 59, 282–297.Google Scholar
Ayres, R. U. 1989. Industrial metabolism. Pages 23–49 of: Ausubel, J.H., and Sladovich, H.E. (eds) Technology and Environment. Washington DC: National Academy Press.
Ayres, R. U., and Nair, I. 1984. Thermodynamics and economics. Phys. Today, 37, 62–71.Google Scholar
Ayres, R. U., Ayres, L.W., and Warr, B. 2003. Exergy, power and work in the US economy, 1900–1998. Energy, 28, 219–273.Google Scholar
Backus, G. E. 1975. Gross thermodynamics of heat engines in deep interior of Earth. Proc. Natl. Acad. Sci. USA, 72, 1555–1558.Google Scholar
Bala, G., Duffy, P. B., and Taylor, K. E. 2008. Impact of geoengineering schemes on the global hydrologic cycle. Proc. Natl. Acad. Sci. USA, 105, 7664–7669.Google Scholar
Barber, J. 2009. Photosynthetic energy conversion: natural and artificial. Chem. Soc. Rev., 38, 185–196.Google Scholar
Beer, C., Reichstein, M., Tomelleri, E., Ciais, P., Jung, M., Carvalhais, N., Roedenbeck, C., Arain, M. A., Baldocchi, D., and Bonan, G. B. 2010. Terrestrial gross carbon dioxide uptake: global distribution and covariation with climate. Science, 329, 834–838.Google Scholar
Beerling, D. J., and Royer, D. L. 2002. Reading a CO2 signal from fossil stomata. New Phytol., 153, 387–397.Google Scholar
Bejan, A. 1996. Entropy generation minimization: the new thermodynamics of finite-size devices and finite-time processes. J. Appl. Phys., 79, 1191–1218.Google Scholar
Bejan, A. 1997. Advanced Engineering Thermodynamics. New York: Wiley.
Bejan, A. 2002. Fundamentals of exergy analysis, entropy generation minimization, and the generation of flow architecture. Int. J. Energy Res., 26, 545–565.Google Scholar
Bejan, A. 2007. Constructal theory of pattern formation. Hydrol. Earth Syst. Sci., 11, 753–768.Google Scholar
Bering, E. A., Few, A. A., and Benbrook, J. R. 1998. The global electric circuit. Phys. Today, 24–30.Google Scholar
Betts, A. K., and Ridgway, W. 1989. Climatic equilibrium of the atmospheric convective boundary layer over a tropical ocean. J. Atmos. Sci., 46, 2621–2641.Google Scholar
Betts, A. K., Ball, J. H., Beljaars, A. C. M., Miller, M. J., and Viterbo, P. A. 1996. The land surface-atmosphere interaction: a review based on observational and global modeling perspectives. J. Geophys. Res., 101, 7209–7225.Google Scholar
Betz, A. 1920. Das Maximum der theoretisch möglichen Ausnützung des Windes durch Windmotoren. Z. Gesamte Turbinenwesen, 26, 307–309.Google Scholar
Bister, M., and Emanuel, K. A. 1998. Dissipative heating and hurricane intensity. Meteorol. Atmos. Phys., 65, 233–240.Google Scholar
Bister, M., Renno, N., Pauluis, O., and Emanuel, K. 2011. Comment on Makarieva et al. ‘A critique of some modern applications of the Carnot heat engine concept: the dissipative heat engine cannot exist’. Proc. R. Soc. London A, 467, 1–6.Google Scholar
Bobe, R., and Behrensmeyer, A. K. 2004. The expansion of grassland ecosystems in Africa in relation to mammalian evolution and the origin of the genus Homo. Paleogeogr. Paleocl., 207, 399–420.Google Scholar
Bohren, C. F., and Albrecht, B. A. 1998. Atmospheric Thermodynamics. New York: Oxford University Press.
Bolton, J. R., and Hall, D. O. 1991. The maximum efficiency of photosynthesis. Photochem. Photobiol., 53, 545–548.Google Scholar
Boltzmann, L. 1886. Der zweite Hauptsatz der mechanischenWärmetheorie. Almanach der kaiserlichen Akademie der Wissenschaften, 36, 225–259.Google Scholar
Bonan, G B. 2008. Forests and climate change: forcings, feedbacks, and the climate benefits of forests. Science, 320, 1444–1449.Google Scholar
Bony, S., Bellon, G., Klocke, D., Sherwood, S., Fermepin, S., and Denvil, S. 2013. Robust direct effect of carbon dioixide on tropical circulation and regional precipitation. Nature Geosci., 6, 447–451.Google Scholar
Borucki, W. J., and Chameides, W. L. 1984. Lightning: estimates of rates of energy dissipation and nitrogen fixation. Rev. Geophys., 22, 363–372.Google Scholar
Bowring, S. P. K., Miller, L. M., Ganzeveld, L., and Kleidon, A. 2014. Applying the concept of “energy return on investment” to desert greening of the Sahara/Sahel using a global climate model. Earth Syst. Dynam., 5, 43–53.Google Scholar
Boyce, C. K., Brodribb, T. J., Feild, T. S., and Zwieniecki, M. A. 2009. Angiosperm leaf vein evolution was physiologically and environmentally transformative. Proc. R. Soc. Lond. B, 276, 1771–1776.Google Scholar
BP, . 2014. BP Statistical Review of World Energy 2014. Tech. rept. London, UK: BP P.L.C.
Brodribb, T. J., and Feild, T. S. 2010. Leaf hydraulic evolution led a surge in leaf photosynthetic capacity during early angiosperm diversification. Ecol. Lett., 13, 175– 183.Google Scholar
Brodribb, T. J., Feild, T. S., and Jordan, G. J. 2007. Leaf maximum photosynthetic rate and venation are linked by hydraulics. Plant Physiol., 144, 1890–1898.Google Scholar
Brovkin, V. 2002. Climate-vegetation interaction. J. Phys-Paris, 12, 57–72.Google Scholar
Brunsell, N. A., Schymanski, S. J., and Kleidon, A. 2011. Quantifying the thermodynamic entropy budget of the land surface: is this useful? Earth Syst. Dynam., 2, 87–103.Google Scholar
Brunt, D. 1941. Physical and Dynamical Meteorology. London: Cambridge University Press.
Budyko, M. I. 1974. Climate and Life. Translated from the Original Russian Edition. New York: Academic Press.
Callies, U., and Herbert, F. 1988. Radiative processes and non-equilibrium thermodynamics. Z. Angew. Math. Phys., 39, 242–266.Google Scholar
Campbell, G. S., and Norman, J. M. 1998. An Introduction to Environmental Biophysics. 2nd edn. New York: Springer.
Campbell, I. H., and Taylor, S. R. 1983. No water, no granites - no oceans, no continents. Geophys. Res. Lett., 10, 1061–1064.Google Scholar
Carnot, S. 1824. Reflections on theMotive Power of Fire and on Machines Fitted to Develop that Power. Paris: Bachelier.
Carpenter, S. R., and Kitchell, J. F. 1984. Plankton community structure and limnetic primary production. Am. Nat., 124, 159–172.Google Scholar
Catling, D. C. 2005. Coupled evolution of Earth's atmosphere and biosphere. Pages 191– 206 of: Kleidon, A., and Lorenz, R. D. (eds), Non-Equilibrium Thermodynamics and the Production of Entropy: Life, Earth, and Beyond. Heidelberg: Springer.
Cawood, P. A., Hawkesworth, C. J., and Dhuime, B. 2013. The continental record and the generation of continental crust. GSA Bulletin, 125, 14–32.Google Scholar
Chaisson, E. J. 1998. The cosmic environment for the growth of complexity. Biosystems, 46, 13–19.Google Scholar
Charru, F., Andreotti, B., and Claudin, P. 2013. Sand ripples and dunes. Annu. Rev. Fluid Mech., 45, 469–493.Google Scholar
Christensen, U. R., and Tilgner, A. 2004. Power requirements of the geodynamo from ohmic losses in numerical and laboratory dynamos. Nature, 429, 169–171.Google Scholar
Ciais, P., Sabine, C., Bala, G., Bopp, L., Brovkin, V., Canadell, J., Chhabra, A., DeFries, R., Galloway, J., Heimann, M., Jones, C., Quéré, C. Le, Myneni, R. B., Piao, S., and Thornton, P. 2013. Carbon and other biogeochemical cycles. In: Stocker, T. F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S. K., Boschung, J., Nauels, A., Xia, Y., Bex, V., and Midgley, P.M. (eds), Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York: Cambridge University Press.
Ciamician, G. 1912. The photochemistry of the future. Science, 36, 385–394.Google Scholar
Cicerone, R. J. 1987. Changes in stratospheric ozone. Science, 237, 35–42.Google Scholar
Clark, T. L., Jenkins, M. A., Coen, J., and Packham, D. 1996. A coupled atmosphere-fire model: convective feedback on fire-line dynamics. J. Appl. Meteor., 35, 875–901.Google Scholar
Clausius, R. 1850. Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich daraus für die Wärmelehre selbst ableiten lassen. Ann. Phys. Chem., 155, 368–397.Google Scholar
Cleveland, C. J., and Ruth, M. 1997. When, where, and by how much do biophysical limits constrain the economic process? A survey of Nicholas Georgescu-Roegen's contribution to ecological economics. Ecol. Econ., 22, 203–223.Google Scholar
Cleveland, C. J., Costanza, R., Hall, C. A. S., and Kaufmann, R. 1984. Energy and the US economy: a biophysical perspective. Science, 225, 890–897.Google Scholar
Cohen, J. E. 1995. Population growth and Earth's human carrying capacity. Science, 269, 341–346.Google Scholar
Corten, G. 2001. Novel views on the extraction of energy from wind-heat generation concentration and terrain. Page 5 of: Proceedings of the 2001 EWEC Conference. available at www.ecn.nl/docs/library/report/2001/rx01054.pdf. Accessed May 2, 2014.
Count, E. W. 1947. Brain and body weight in man. Ann. N. Y. Acad. Sci., 46, 993–1122.Google Scholar
Cross, M, and Hohenberg, P. 1993. Pattern formation outside of equilibrium. Rev. Mod. Phys., 65, 851–1112.Google Scholar
Crutzen, P. J. 2002. Geology of mankind. Nature, 415, 23.Google Scholar
Curzon, F. L., and Ahlborn, B. 1975. Efficiency of a Carnot engine at maximum power output. Am. J. Phys., 43, 22–24.Google Scholar
Daly, H. E. 1992. Is the entropy law relevant to economics of natural resource scarcity? Yes, of course it is! J. Environ. Econ. Manage., 23, 91–95.Google Scholar
Davies, J. H., and Davies, D. R. 2010. Earth's surface heat flux. Solid Earth, 1, 5–24.Google Scholar
de, Arellano, J., Vila-Guerau, Ouwersloot, H. G., Baldocchi, D., and Jacobs, C. M. J. 2014. Shallow cumulus rooted in photosynthesis. Geophys. Res. Lett., 41, 1796–1802.Google Scholar
de Bruin, H. A. R., and Lablans, W. N. 1998. Reference crop evapotranspiration determined with a modified Makkink equation. Hydrol. Process., 12, 1053–1062.Google Scholar
Denmead, O. T., Raupach, M. R., Dunin, F. X., Cleugh, H. A., and Leuning, R. 1996. Boundary layer budgets for regional estimates of scalar fluxes. Glob. Ch. Biol., 2, 255–264.Google Scholar
Desjardins, R. L., Brach, E. J., Alvo, P., and Schuepp, P. H. 1982. Aircraft monitoring of surface carbon dioxide exchange. Science, 216, 733–735.Google Scholar
Dewar, R. C. 2003. Information theory explanation of the fluctuation theorem, Maximum Entropy Production, and self-organized criticality in non-equilibrium stationary states. J. Physics A, 36, 631–641.Google Scholar
Dewar, R. C. 2005a. Maximum entropy production and non-equilibrium statistical mechanics. In: Kleidon, A., and Lorenz, R. D. (eds), Non-Equilibrium Thermodynamics and the Production of Entropy: Life, Earth, and Beyond. Heidelberg: Springer.
Dewar, R. C. 2005b. Maximum entropy production and the fluctuation theorem. J. Physics A, 38, L371–L381.Google Scholar
Dewar, R C. 2010. Maximum entropy production as an inference algorithm that translates physical assumptions into macroscopic predictions: don't shoot the messenger. Entropy, 11, 931–944.Google Scholar
Dewar, R C, Lineweaver, C H, Niven, R K, and Regenauer-Lieb, K. 2014. Beyond the second law: an overview. Pages 3–27 of: Dewar, R C, Lineweaver, C H, Niven, R K, and Regenauer-Lieb, K (eds), Beyond the Second Law: Entropy Production and Nonequilibrium Systems. Heidelberg: Springer.
Dincer, I., and Rosen, M. A. 2005. Thermodynamic aspects of renewables and sustainable development. Renew. Sust. Energ. Rev., 9, 169–189.Google Scholar
Dukes, J. S. 2003. Burning buried sunshine: human consumption of ancient solar energy. Clim. Ch., 61, 31–44.Google Scholar
Duysens, L. N. M. 1958. The path of light energy in photosynthesis. Pages 10–25 of: Brookhaven Symposia in Biology 1: The Photochemical Apparatus, its Structure & Function. Upton, NY, USA: Brookhaven Natl. Lab.
Dyke, J. G., Gans, F., and Kleidon, A. 2011. Towards understanding how surface life can affect interior geological processes: a non-equilibrium thermodynamics approach. Earth Syst. Dynam., 2, 139–160.Google Scholar
Dyson, F J. 1960. Search for artificial stellar sources of infrared radiation. Science, 131, 1667–1668.Google Scholar
Eddington, A. S. 1928. The Nature of the Physical World. New York: Macmillan.
Edlefsen, N. E., and Anderson, A. B. C. 1943. Thermodynamics of soil moisture. Hilgardia, 15, 31–298.Google Scholar
Ehleringer, J. R., and Cerling, T. E. 1995. Atmospheric CO2 and the ratio of intercellular to ambinent CO2 concentrations in plants. Tree Physiol., 15, 105–111.Google Scholar
Elder, J. 1976. The Bowels of the Earth. London: Oxford University Press.
Elimelech, M., and Phillip, W. A. 2011. The future of seawater desalination: energy, technology and the environment. Science, 333, 712–717.Google Scholar
Emanuel, K. 1987. The dependence of hurricane intensity on climate. Nature, 326, 483–485.Google Scholar
Emanuel, K. 2005. Increasing destructiveness of tropical cyclones over the past 30 years. Nature, 436, 686–688.Google Scholar
Emanuel, K. 2007. Environmental factors affecting tropical cyclone power dissipation. J. Clim., 20, 5497–5509.Google Scholar
Emanuel, K. 2013. Downscaling CMIP5 climate models shows increased tropical cyclone activity over the 21st century. Proc. Natl. Acad. Sci. USA, 110, 12219–12224.Google Scholar
Emanuel, K. A. 1986. An air-sea interaction theory for tropical cyclones. Part I: steady-state maintenance. J. Atmos. Sci., 43, 585–604.Google Scholar
Emanuel, K. A. 1999. Thermodynamic control of hurricane intensity. Nature, 401, 665–669.Google Scholar
Ertel, H., and Köhler, H. 1948. On the thermodynamic efficiency of steady atmospheric circulations. Geofisica Pura e Applicata, 13, 102–108.Google Scholar
Essex, C. 1984. Radiation and the irreversible thermodynamics of climate. J. Atmos. Sci., 41, 1985–1991.Google Scholar
Falk, D. 1990. Brain evolution in Homo: the “radiator” theory. Behav. Brain Sci., 13, 333–381.Google Scholar
Farquhar, G. D., and Sharkey, T. D. 1982. Stomatal conductance and photosynthesis. Ann. Rev. Plant. Physiol., 33, 317–345.Google Scholar
Feild, T. S., Brodribb, T. J., Iglesias, A., Chatelet, D. S., Baresch, A., Upchurch, G. R.,
Gomez, B., Mohr, B. A. R., Coiffard, C., Kvacek, J., and Jaramillo, C. 2011. Fossil evidence for Cretaceous escalation in angiosperm leaf vein evolution. Proc. Natl. Acad. Sci. USA, 108, 8363–8366.Google Scholar
Ferrari, R., and Wunsch, C. 2009. Ocean circulation kinetic energy: reservoirs, sources, and sinks. Annu. Rev. Fluid Mech., 41, 253–282.Google Scholar
Feynman, R. P., Leighton, R. B., and Sands, M. 1966. The Feynman Lectures on Physics. Reading, MA, USA: Addison-Wesley.
Field, C. B., Behrenfeld, M. J., Randerson, J. T., and Falkowski, P. 1998. Primary production of the biosphere: integrating terrestrial and oceanic components. Science, 281, 237–240.Google Scholar
Field, C. B., Campbell, J. E., and Lobell, D. B. 2007. Biomass energy: the scale of the potential resource. Trends Ecol. Evol., 23, 65–72.Google Scholar
Fischer-Kowalski, M., and Haberl, H. 1998. Sustainable development: socio-economic metabolism and the colonization of nature. Int. Soc. Sci. J., 50, 573–587.Google Scholar
Foley, J. A., DeFries, R., Asner, G. P., Barford, C., Bonan, G., Carpenter, S. R., Chapin, F. S., Coe, M. T., Daily, G. C., Gibbs, H. K., Helkowski, J. H., Holloway, T., Howard, E. A., Kucharik, C. J., Monfreda, C., Patz, J. A., Prentice, I. C., Ramankutty, N., and Snyder, P. K. 2005. Global consequences of land use. Science, 309, 570–574.Google Scholar
Foley, R. A., and Lee, P. C. 1991. Ecology and energetics of encephalization in hominid evolution. Phil. Trans. R. Soc. B, 334, 223–232.Google Scholar
Forster, P., Ramaswamy, V., Artaxo, P., Berntsen, T., Betts, R., Fahey, D.W., Haywood, J., Lean, J., Lowe, D.C., Myhre, G., Nganga, J., Prinn, R., Raga, G., Schulz, M., and Dorland, R. Van. 2007. Changes in atmospheric constituents and in radiative forcing. In: Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., M., Tignor, and Miller, H.L. (eds), Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York: Cambridge University Press.
Franck, S., von Bloh, W., Müller, C., Bondeau, A., and Sakschweski, B. 2011. Harvesting the Sun: new estimations of the maximum population of planet Earth. Ecol.Mod., 222, 2019–2026.Google Scholar
Gans, F., Miller, L. M., and Kleidon, A. 2012. The problem of the second wind turbine: a note on a common but flawed wind power estimation method. Earth Syst. Dynam., 3, 79–86.Google Scholar
Garratt, J. R. 1992. The Atmospheric Boundary Layer. Cambridge, UK: Cambridge University Press.
Garrett, C., and Cummins, P. 2007. The efficiency of a turbine in a tidal channel. J. Fluid Mech., 588, 243–251.Google Scholar
Garrett, T J. 2009. Are there basic physical constraints on future anthropogenic emissions of carbon dioxide? Clim. Change, 104, 437–455.Google Scholar
Georgescu-Roegen, N. 1971. The Entropy Law and the Economic Process. Cambridge, MA: Harvard University Press.
Glaser, P. E. 1968. Power from the Sun: its future. Science, 162, 857–861.Google Scholar
Gnanadesikan, A., Slater, R. D., Swathi, P. S., and Vallis, G. K. 2005. The energetics of ocean heat transport. J. Clim., 18, 2604–2616.Google Scholar
Goldstein, B., Hiriart, G., Bertani, R., Bromley, C., Gutierrez-Negrin, L., Huenges, E., Muraoka, H., Ragnarsson, A., Tester, J., and Zui, V. 2011. Geothermal energy. Pages 401–436 of: Edenhofer, O., Pichs-Madruga, R., Sokona, Y., Seyboth, K., Matschoss, P., Kadner, S., Zwickel, T., Eickemeier, P., Hansen, G., Schlömer, S., and von Stechow, C. (eds), IPCC Special Report on Renewable Energy Sources and Climate ChangeMitigation. Cambridge, UK and New York: Cambridge University Press.
Goody, R. 2000. Sources and sinks of climate entropy. Q. J. R. Meteorol. Soc., 126, 1953–1970.Google Scholar
Goody, R. 2003. On the mechanical efficiency of deep, tropical convection. J. Atmos. Sci., 60, 2827–2832.Google Scholar
Goody, R. 2007. Maximum entropy production in climate theory. J. Atmos. Sci., 64, 2735–2739.Google Scholar
Goody, R., and Abdou, W. 1996. Reversible and irreversible sources of radiation entropy. Q. J. Roy. Meteorol. Soc., 122(530), 483–494.Google Scholar
Gould, S. J. 1966. Allometry and size in ontogeny and phylogeny. Biol. Rev., 41, 587–640.Google Scholar
Gyftopoulos, E. P., and Beretta, G. P. 1991. Thermodynamics: Foundations and Applications. New York: Macmillan.
Haaf, W., Friedrich, K., Mayr, G., and Schlaich, J. 1983. Solar chimneys. Part I: principle and construction of the pilot plant in Manzanares. Int. J. Solar Energy, 2, 3–20.Google Scholar
Haberl, H, Erb, K H, Krausmann, F, Gaube, V, Bondeau, A, Pluttzar, C, Gingrich, S, Lucht, W, and Fischer-Kowalski, M. 2007. Quantifying and mapping the human appropriation of net primary productivity in earth's terrestrial ecosystems. Proc. Natl. Acad. Sci. USA, 104, 12942–12947.Google Scholar
Haberl, H., Erb, K.-H., and Krausmann, F. 2014. Human appropriation of net primary production: patterns, trends, and planetary boundaries. Annu. Rev. Environ. Resour., 39, 363–391.Google Scholar
Haff, P. K. 2010. Hillslopes, rivers, plows, and trucks: mass transport on Earth's surface by natural and technological processes. Earth Surf. Process. Landforms, 35, 1157–1166.Google Scholar
Haff, P. K. 2013. Technology as a geological phenomenon: implications for human wellbeing. In: Waters, C. N., Zalasiewicz, J. A., Williams, M., Ellis, M. A., and Snelling, A. M. (eds), A Stratigraphic Basis for the Anthropocene. Special Publications, vol. 395. London: Geological Society.
Haken, H. 1975. Cooperative phenomena in systems far from thermal equilibrium and in nonphysical systems. Rev. Mod. Phys., 47, 67–121.Google Scholar
Hall, C., Lindenberger, D., Kümmel, R., Kroeger, T., and Eichhorn, W. 2001. The need to reintegrate the natural sciences with economics. Bioscience, 51, 663–673.Google Scholar
Hammond, K. A., and Diamond, J. 1997. Maximal sustained energy budgets in humans and animals. Nature, 386, 457–462.Google Scholar
Hansen, J., Lacis, A., Rind, D., Russell, G., Stone, P., Fung, I., Ruedy, R., and Lerner, J. 1984. Climate sensitivity: analysis of feedback mechanisms. In: Climate Processes and Climate Sensitivity, Geophysical Monograph 29. Washington, DC: American Geophysical Union.
Hartmann, D. L. 1994. Global Physical Climatology. San Diego: Academic Press.
Held, I. M., and Soden, B. J. 2006. Robust responses of the hydrological cycle to global warming. J. Clim., 19, 5686–5699.Google Scholar
Herrmann, W. A. 2006. Quantifying global exergy resources. Energy, 31, 1685–1702.Google Scholar
Hewitt, J. M., McKenzie, D. P., and Weiss, N. O. 1975. Dissipative heating in convective flows. J. Fluid Mech., 68, 721–738.Google Scholar
Hill, R., and Rich, P. R. 1983. A physical interpretation for the natural photosynthetic process. Proc. Natl. Acad. Sci. USA, 80, 978–982.Google Scholar
Hitchcock, D. R., and Lovelock, J. E. 1967. Life detection by atmospheric analysis. Icarus, 7, 149–159.Google Scholar
Hobbs, P. V. 2000. Introduction to Atmospheric Chemistry. Cambridge, UK: Cambridge University Press.
Hoening, D., Hansen-Goos, H., Airo, A., and Spohn, T. 2014. Biotic vs. abiotic Earth: a model for mantle hydration and continental coverage. Planet. Space Sci., 98, 5–13.Google Scholar
Hoffert, M. I., Caldeira, K., Benford, G., Criswell, D. R., Green, C., Herzog, H., Jain, A. K., Kheshgi, H. S., Lackner, K. S., Lewis, J. S., Lightfoot, H. D., Manheimer, W., Mankins, J. C., Mauel, M. E., Perkins, L. J., Schlesinger, M. E., Volk, T., and Wigley, T. M. L. 2002. Advanced technology paths to global climate stability: energy for a greenhouse planet. Science, 298, 981–987.Google Scholar
Holdaway, R. J., Sparrow, A. D., and Coomes, D. 2010. Trends in entropy production during ecosystem development in the Amazon basin. Phil. Trans. R. Soc. B, 365, 1437–1447.Google Scholar
Holland, H. H. 2006. The oxygenation of the atmosphere and ocean. Phil. Trans. Roy. Soc. London B, 361, 903–915.Google Scholar
Holton, J. R. 1992. An Introduction to Dynamic Meteorology. 3rd edn. San Diego: Academic Press.
Howard, A. D. 1990. Theoretical model of optimal drainage networks. Water Resour. Res., 26, 2107–2117.Google Scholar
Huang, S-S. 1959. Occurrence of life in the universe. Am. Sci., 47, 397–402.Google Scholar
Hubbert, M. King. 1981. The world's evolving energy system. Am. J. Phys., 49, 1007–1029.Google Scholar
Hyde, W. T., Crowley, T. J., Baum, S. K., and Peltier, R. 2000. Neoproterozoic ‘snowball Earth’ simulations with a coupled climate/ice-sheet model. Nature, 405, 425–429.Google Scholar
Imhoff, M. L., Bounoua, L., Ricketts, T., Loucks, C., Harriss, R., and Lawrence, W. T. 2004. Global patterns in human consumption of net primary production. Nature, 429, 870–873.Google Scholar
IPCC. 2011. IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. Prepared byWorking Group III of the Intergovernmental Panel on Climate Change [O., Edenhofer, R., Pichs-Madruga, Y., Sokona, K., Seyboth, P., Matschoss, S., Kadner, T., Zwickel, P., Eickemeier, G., Hansen, S., Schlömer, C. von, Stechow (eds)]. Cambridge, UK and New York: Cambridge University Press.
Isaacs, J D, and Schmitt, W R. 1980. Ocean energy: forms and prospects. Science, 207, 265–273.Google Scholar
Jacobsen, M. Z., and Archer, C. L. 2012. Saturation wind power potential and its implications for wind energy. Proc. Natl. Acad. Sci. USA, 109, 15679–15684.Google Scholar
Jerison, H. J. 1955. Brain to body ratios and the evolution of intelligence. Science, 121, 447–449.Google Scholar
Jerison, H. J. 1977. The theory of encephalization. Ann. N. Y. Acad. Sci., 299, 146–160.Google Scholar
Jerison, H. J. 1985. Animal intelligence as encephalization. Phil. Trans. R. Soc. B, 308, 21–35.Google Scholar
Jupp, T., and Schultz, A. 2000. A thermodynamic explanation for black smoker temperatures. Nature, 403, 880–883.Google Scholar
Jupp, T. E., and Cox, P. M. 2010. MEP and planetary climates: insights from a two-box climate model containing atmospheric dynamics. Phil. Trans. R. Soc. B, 365, 1355–1365.Google Scholar
Jupp, T. E., and Schultz, A. 2004. Physical balances in subseafloor hydrothermal convection cells. J. Geophys. Res., 109.Google Scholar
Kabelac, S. 1994. Thermodynamik der Strahlung. Braunschweig and Wiesbaden: Vieweg.
Kagan, B. A., and Sündermann, Jürgen. 1996. Dissipation of tidal energy, paleotides, and evolution of the Earth-Moon system. Adv. Geophys., 38, 179–266.Google Scholar
Kasting, J. F. 1993. Earth's early atmosphere. Science, 259, 920–926.Google Scholar
Kasting, J. F., and Catling, D. 2003. Evolution of a habitable planet. Annu. Rev. Astron. Astrophys., 41, 429–463.Google Scholar
Kasting, J. F., Whitmire, D. P., and Reynolds, R. T. 1993. Habitable zones around main sequence stars. Icarus, 101, 108–128.Google Scholar
Keith, D., DeCarolis, J., Denkenberger, D., Lenschow, D., Malyshev, S. L., Pacala, S., and Rasch, P. 2004. The influence of large-scale wind power on global climate. Proc. Natl. Acad. Sci. USA, 101, 16116–16120.Google Scholar
Klausmeier, C A. 1999. Regular and irregular patterns in semiarid vegetation. Science, 284, 1826–1828.Google Scholar
Kleiber, M. 1932. Body size and metabolism. Hilgardia, 6, 315–353.Google Scholar
Kleidon, A. 2004a. Beyond Gaia: thermodynamics of life and Earth system functioning. Clim. Ch., 66, 271–319.Google Scholar
Kleidon, A. 2004b. Optimized stomatal conductance of vegetated land surfaces and its effects on simulated productivity and climate. Geophys. Res. Lett., 31, L21203.Google Scholar
Kleidon, A. 2006. The climate sensitivity to human appropriation of vegetation productivity and its thermodynamic characterization. Glob. Planet. Ch., 54, 109–127.Google Scholar
Kleidon, A. 2007. Optimized stomatal conductance and the climate sensitivity to carbon dioxide. Geophys. Res. Lett., 34, L14709.Google Scholar
Kleidon, A. 2008. Entropy production by evapotranspiration and its geographic variation. Soil Water Res., 3, S89–S94.Google Scholar
Kleidon, A. 2009a. Climatic constraints on maximum possible levels of human activity and their relation to human evolution and global change. Clim. Ch., 95, 405–431.Google Scholar
Kleidon, A. 2009b. Maximum entropy production and general trends in biospheric evolution. Paleontol. J., 43, 130–135.Google Scholar
Kleidon, A. 2009c. Non-equilibrium thermodynamics and maximum entropy production in the Earth system: applications and implications. Naturwissenschaften, 96, 653–677.Google Scholar
Kleidon, A. 2010. Life, hierarchy, and the thermodynamic machinery of planet Earth. Phys. Life Rev., 7, 424–460.Google Scholar
Kleidon, A. 2012. How does the Earth system generate and maintain thermodynamic disequilibrium and what does it imply for the future of the planet?Phil. Trans. R. Soc. A, 370, 1012–1040.Google Scholar
Kleidon, A., and Fraedrich, K. 2005. Biotic entropy production and global atmospherebiosphere interactions. Pages 173–190 of: Kleidon, A., and Lorenz, R. D. (eds), Non-Equilibrium Thermodynamics and the Production of Entropy: Life, Earth, and Beyond. Heidelberg: Springer.
Kleidon, A., and Heimann, M. 1998. A method of determining rooting depth from a terrestrial biosphere model and its impacts on the global water- and carbon cycle. Global Change Biol., 4, 275–286.Google Scholar
Kleidon, A., and Lorenz, R. D. (eds). 2005. Non-Equilibrium Thermodynamics and the Production of Entropy: Life, Earth, and Beyond. Heidelberg: Springer.
Kleidon, A., and Renner, M. 2013a. A simple explanation for the sensitivity of the hydrologic cycle to climate change. Earth Syst. Dynam., 4, 455–465.Google Scholar
Kleidon, A, and Renner, M. 2013b. Thermodynamic limits of hydrologic cycling within the Earth system: concepts, estimates and implications. Hydrol. Earth Syst. Sci., 17, 2873–2892.Google Scholar
Kleidon, A., Fraedrich, K., and Heimann, M. 2000. A green planet versus a desert world: estimating the maximum effect of vegetation on land surface climate. Clim. Ch., 44, 471–493.Google Scholar
Kleidon, A., Fraedrich, K., Kunz, T., and Lunkeit, F. 2003. The atmospheric circulation and states of maximum entropy production. Geophys. Res. Lett., 30, 2223.Google Scholar
Kleidon, A., Fraedrich, K., Kirk, E., and Lunkeit, F. 2006. Maximum entropy production and the strength of boundary layer exchange in an atmospheric general circulation model. Geophys. Res. Lett., 33, L06706.Google Scholar
Kleidon, A., Schymanski, S., and Stieglitz, M. 2009. Thermodynamics, irreversibility and optimality in land surface hydrology. Pages 107–118 of: Strelcova, K., Matyas, C., Kleidon, A., Lapin, M., Matejka, F., Skvarenina, J., and Holecy, J. (eds), Bioclimatology and Natural Hazards. Heidelberg: Springer.
Kleidon, A., Malhi, Y., and Cox, P. M. 2010. Maximum entropy production in environmental and ecological systems. Phil. Trans. R. Soc. B, 365, 1297–1302.Google Scholar
Kleidon, A, Zehe, E, Ehret, U, and Scherer, U. 2013. Thermodynamics, maximum power, and the dynamics of preferential river flow structures on continents. Hydrol. Earth Syst. Sci., 17, 225–251.Google Scholar
Kleidon, A., Zehe, E., Ehret, U., and Scherer, U. 2014a. Earth system dynamics beyond the second law: maximum power limits, dissipative structures, and planetary interactions. Pages 163–182 of: Dewar, R. C., Lineweaver, C. H., Niven, R. K., and Regenauer- Lieb, K. (eds), Beyond the Second Law: Entropy Production and Non-equilibrium Systems. Berlin and Heidelberg: Springer.
Kleidon, A., Renner, M., and Porada, P. 2014b. Estimates of the climatological land surface energy and water balance derived from maximum convective power. Hydrol. Earth Syst. Sci., 18, 2201–2218.Google Scholar
Kleidon, A., Miller, L., and Gans, F. 2016. Physical limits of solar energy conversion in the Earth system. Top. Curr. Chem., 371, 1–22.Google Scholar
Klein, M. J. 1967. Thermodynamics in Einstein's thought. Science, 157, 509–516.Google Scholar
Klein Goldewijk, K., Beusen, A., and Janssen, P. 2010. Long term dynamic modeling of global population and built-up area in a spatially explicit way, HYDE 3.1. Holocene, 20, 565–573.Google Scholar
Klein Goldewijk, K., Beusen, A., de Vos, M., and van Drecht, G. 2011. The HYDE 3.1 spatially explicit database of human induced land use change over the past 12,000 years. Glob. Ecol. Biogeog., 20, 73–86.Google Scholar
Knox, R. S. 1969. Thermodynamics and the primary processes of photosynthesis. Biophys J, 9, 1351–1362.Google Scholar
Kok, J. F., Parteli, E. J. R., Michaels, T. I., and Karam, D. B. 2012. The physics of wind-blown sand and dust. Rep. Prog. Phys., 75, 106901.Google Scholar
Kondepudi, D., and Prigogine, I. 1998. Modern Thermodynamics: From Heat Engines to Dissipative Structures. Chichester: Wiley.
Köppen, W. 1923. Die Klimate der Erde. Berlin: de Gruyter.
Kump, L. R., Brantley, S. L., and Arthur, M. A. 2000. Chemical weathering, atmospheric CO2, and climate. Annu. Rev. Earth Planet. Sci., 28, 611–667.Google Scholar
Lambert, F. L. 2002. Entropy is simple, qualitatively. J. Chem. Ed., 79, 1241–1246.Google Scholar
Landsberg, P T, and Tonge, G. 1979. Thermodynamics of the conversion of diluted radiation. J. Phys. A, 12, 551–562.Google Scholar
Landsberg, P. T., and Tonge, G. 1980. Thermodynamic energy conversion efficiencies. J. Appl. Phys., 51, R1.Google Scholar
Law, B. E., Falge, E., Gu, L., Baldocchi, D. D., Bakwin, P., Berbigier, P., Davis, K., Dolman, A. J., Falk, M., Fuentes, J. D., Goldstein, A., Granier, A., Grelle, A., Hollinger, D., Janssens, I. A., Jarvis, P., Jensen, N. O., Katul, G., Malhi, Y., Matteucci, G., Meyers, T., Monson, R., Munger, W., Oechel, W., Olson, R., Pilegaard, K., U, K. T. Paw, Thorgeirsson, H., Valentini, R., Verman, S., Vesala, T., Wilson, K., and Wofsy, S. 2002. Environmental controls over carbon dioxide and water vapor exchange of terrestrial vegetation. Agric. For. Meteor., 113, 97–120.Google Scholar
Leff, H. S. 1996. Thermodynamic entropy: the spreading and sharing of energy. Am. J. Phys., 64, 1261–1271.Google Scholar
Leff, H. S. 2007. Entropy, its language, and interpretation. Found. Phys., 37, 1744–1766.Google Scholar
Lejeune, O, and Tlidi, M. 1999. A model for the explanation of vegetation stripes (Tiger Bush). J. Veg. Sci., 10, 201–208.Google Scholar
Lejeune, O., Tlidi, M., and Couteron, P. 2002. Localized vegetation patches: a selforganized response to resource scarcity. Phys. Rev. E, 66, 010901.Google Scholar
Lenardic, A, Moresi, L.-N., Jellinek, A. M., and Manga, M. 2005. Continental insulation, mantle cooling, and the surface area of oceans and continents. Earth Planet. Sci. Lett., 234, 317–333.Google Scholar
Lenton, T. M. 1998. Gaia and natural selection. Nature, 394, 439–447.Google Scholar
Lenton, T. M., Schellnhuber, H. J., and Szathmary, E. 2004. Climbing the co-evolutionary ladder. Nature, 913.Google Scholar
Lenton, T. M., Held, H., Kriegler, E., Hall, J. W., Lucht, W., Rahmstorf, S., and Schellnhuber, H. J. 2008. Tipping elements in the Earth's climate system. Proc. Natl. Acad. Sci. USA, 105, 1786–1793.Google Scholar
Li, L., Ingersoll, A. P., Jiang, X., Feldman, D., and Yung, Y. L. 2007. Lorenz energy cycle of the global atmosphere based on reanalysis datasets. Geophys. Res. Lett., 34, L16813.Google Scholar
Liao, W., Heijungs, R., and Huppes, G. 2012. Thermodynamic analysis of human– environment systems: a review focused on industrial ecology. Ecol. Mod., 228, 76–88.Google Scholar
Lineweaver, C. H., and Chopra, A. 2012. Earth and other Earths: astrophysical, geochemical, geophysical, and biological limits on planet habitability. Annu. Rev. Earth Planet. Sci., 40, 597–623.Google Scholar
Lineweaver, C. H., and Egan, C. A. 2008. Life, gravity and the second law of thermodynamics. Phys. Life Rev., 5, 225–242.Google Scholar
Lobell, D. B., Cassman, K. G., and Field, C. B. 2009. Crop yield gaps: their importance, magnitudes, and causes. Ann. Rev. Environ. Resour., 34, 179–204.Google Scholar
Lorenz, E. N. 1955. Available potential energy and the maintenance of the general circulation. Tellus, 7, 157–167.Google Scholar
Lorenz, E. N. 1960. Generation of available potential energy and the intensity of the general circulation. Pages 86–92 of: Pfeffer, R. C. (ed), Dynamics of Climate. Oxford: Pergamon Press.
Lorenz, R. D. 2002. Planets, life and the production of entropy. Intl. J. Astrobiol., 1, 3–13.Google Scholar
Lorenz, R. D. 2010. The two-box model of climate: limitations and applications to planetary habitability and maximum entropy production studies. Phil. Trans. R. Soc. B, 365, 1349–1354.Google Scholar
Lorenz, R. D., and Renno, N. O. 2002. Work output of planetary atmospheric engines: dissipation in clouds and rain. Geophys. Res. Lett., 29, 1023.Google Scholar
Lorenz, R. D., Lunine, J. I., Withers, P. G., and McKay, C. P. 2001. Titan, Mars and Earth: entropy production by latitudinal heat transport. Geophys. Res. Lett., 28, 415–418.Google Scholar
Lotka, A. J. 1921. Note on the economic conversion factors of energy. Proc. Natl. Acad. Sci. USA, 7, 192–197.Google Scholar
Lotka, A. J. 1922a. Contribution to the energetics of evolution. Proc. Natl. Acad. Sci. USA, 8, 147–151.Google Scholar
Lotka, A. J. 1922b. Natural selection as a physical principle. Proc. Natl. Acad. Sci. USA, 8, 151–154.Google Scholar
Lotka, A. J. 1925. Elements of Physical Biology. Baltimore: Williams and Wilkins.
Lovelock, J. E. 1965. A physical basis for life detection experiments. Nature, 207, 568–570.Google Scholar
Lovelock, J. E. 1972a. Gaia: A New Look at Life on Earth. Oxford: Oxford University Press.
Lovelock, J. E. 1972b. Gaia as seen through the atmosphere. Atmos. Environ., 6, 579–580.Google Scholar
Lovelock, J. E. 1975. Thermodynamics and the recognition of alien biospheres. Proc. Roy. Soc. Lond. B, 189, 167–181.Google Scholar
Lovelock, J. E., and Margulis, L. 1974. Atmospheric homeostasis by and for the biosphere: the Gaia hypothesis. Tellus, 26, 2–10.Google Scholar
Lunine, J. I., and Lorenz, R. D. 2002. A simple prescription for calculating day–night temperature contrasts on synchronously rotating planets. In: 33rd Annual Lunar and Planetary Science Conf., Houston, TX, 11–15 March, 2002.
Lyons, T. J. 2002. Clouds prefer native vegetation. Met. Atmos. Phys., 80, 131–140.Google Scholar
Magnus, G. 1844. Versuche über die Spannkräfte des Wasserdampfs. Ann. Phys. Chem., 61, 225–248.Google Scholar
Makkink, G. F. 1957. Testing the Penman formula by means of lysimeters. J. Inst. Wat. Engrs., 11, 277–288.Google Scholar
Manabe, S., and Moeller, F. 1961. On the radiative equilibrium and heat balance of the atmosphere. Mon. Wea. Rev., 89, 503–532.Google Scholar
Marais, D. J. Des. 2000. When did photosynthesis emerge on Earth?Science, 289, 1703–1705.Google Scholar
Marchetti, C. 1979. 1012: a check on the Earth-carrying capacity for man. Energy, 4, 1107–1117.Google Scholar
Margules, M. 1905. Über die Energie der Stürme. Jahrb. Zentralanst. Meteorol., 40, 1–26.Google Scholar
Martin, R. D. 1981. Relative brain size and basal metabolic rate in terrestrial vertebrates. Nature, 293, 57–60.Google Scholar
Martin, W. F., Sousa, F. L., and Lane, N. 2014. Energy at life's origin. Science, 344, 1092–1093.Google Scholar
Martyushev, L. M., and Seleznev, V. D. 2006. Maximum entropy production principle in physics, chemistry, and biology. Phys. Rep., 426, 1–45.Google Scholar
Masuda, K. 1988. Meridional heat transport by the atmosphere and the ocean: analysis of FGGE data. Tellus A, 40, 285–302.Google Scholar
McDonald, J. E. 1960. Direct absorption of solar radiation by atmospheric water vapor. J. Meteor., 17, 319–328.Google Scholar
McNaughton, K. G., and Spriggs, T. W. 1986. A mixed-layer model for regional evaporation. Bound. Lay. Meteorol., 34, 243–262.Google Scholar
McNaughton, S. J. 1979. Grazing as an optimization process: grass-ungulate relationships in the Serengeti. Am. Nat., 113, 691–703.Google Scholar
Meadows, D. H., Meadows, D. L, Randers, J., and III, W. W. Behrens. 1972. The Limits to Growth. New York: Universe Books.
Metchnik, V. I., Gladwin, M. T., and Stacey, F. D. 1974. Core convection as a power source for the geomagnetic dynamo: a thermodynamic argument. J. Geomag. Geoelectr., 26, 405–415.Google Scholar
Meyer, H.J. 1886. Meyers Konversations-Lexikon: Eine Encyklopädie des allgemeinen Wissens. No. Bd. 4. Leipzig: Bibliographisches Institut.
Meysman, F. J. R., and Bruers, S. 2007. A thermodynamic perspective on food webs: quantifying entropy production within detrital-based ecosystems. J. Theor. Biol., 249, 124–139.Google Scholar
Miller, L M, Gans, F, and Kleidon, A. 2011a. Estimating maximum global land surface wind power extractability and associated climatic consequences. Earth Syst. Dynam., 2, 1–12.Google Scholar
Miller, L. M., Gans, F., and Kleidon, A. 2011b. Jet stream wind power as a renewable energy resource: little power, big impacts. Earth Syst. Dynam., 2, 201–212.Google Scholar
Milly, P. C. D. 1994. Climate, soil water storage, and the average annual water balance. Water Resour. Res., 30, 2143–2156.Google Scholar
Monsi, M., and Saeki, T. 1953. Über den Lichtfaktor in den Pflanzengesellschaften und seine Bedeutung für die Stoffproduktion. Jap. J. Bot., 14, 22–52.Google Scholar
Monteith, J. L. 1972. Solar radiation and productivity in tropical ecosystems. J. Appl. Ecol., 9, 747–766.Google Scholar
Monteith, J. L. 1977. Climate and the efficiency of crop production in Britain. Phil. Trans. R. Soc. B, 281, 277–294.Google Scholar
Monteith, J. L. 1978. Reassessment of maximum growth rates for C3 and C4 crops. Expl. Agric., 14, 1–5.Google Scholar
Mortimer, R. G., and Mazo, R. M. 1961. Irreversible thermodynamics of systems containing radiation. Application to photochemical reactions. J. Chem. Phys., 35, 1013–1018.Google Scholar
Mueller, N. D., Gerber, J. S., Johnston, M., Ray, D. K., Ramankutty, N., and Foley, J. A. 2012. Closing yield gaps through nutrient and water management. Nature, 490, 254–257.Google Scholar
Mulligan, J. F., and Hertz, H. G. 1997. An unpublished lecture by Heinrich Hertz: “On the energy balance of the Earth.”Am. J. Phys., 65, 36–45.Google Scholar
Munk, W., and Wunsch, C. 1998. Abyssal recipes II: energetics of tidal and wind mixing. Deep-Sea Res., 45, 1977–2010.Google Scholar
Murphy, D. J., and Hall, C. A. S. 2011a. Energy return on investment, peak oil, and the end of economic growth. Ann. N. Y. Acad. Sci., 1219, 52–72.Google Scholar
Murphy, D. J., and Hall, C. A. S. 2011b. Year in review: EROI or energy return on (energy) invested. Ann. N. Y. Acad. Sci., 1185, 102–118.Google Scholar
Nair, U. S., Wu, Y., Kala, J., Lyons, T. J., Sr., R. A. Pielke, and Hacker, J. M. 2011. The role of land use change on the development and evolution of the west coast trough, convective clouds, and precipitation in southwest Australia. J. Geophys. Res., 116, D07103.Google Scholar
Nepstad, D. C., de Carvalho, C. R., Davidson, E. A., Jipp, P. H., Lefebvre, P. A., Negreiros, H. G., da Silva, E. D., Stone, T. A, Trumbore, S. E., and Vieira, S. 1994. The role of deep roots in the hydrological and carbon cycles of Amazon forests and pastures. Nature, 372, 666–669.Google Scholar
Nepstad, D. C., Stickler, C. M., Soares-Filho, B., and Merry, F. 2008. Interactions among Amazon land use, forests and climate: prospects for a near-term forest tipping point. Phil. Trans. R. Soc. B, 363, 1737–1746.Google Scholar
Nihous, G. C. 2006. A preliminary assessment of ocean thermal energy conversion resources. J. Energy Resour. Technol., 129, 10–17.Google Scholar
Nisbet, E. G., and Sleep, N. H. 2001. The habitat and nature of early life. Nature, 409, 1083–1091.Google Scholar
Novikov, I. I. 1958. The efficiency of atomic power stations (A review). J. Nuclear Energy II, 7, 125–128.Google Scholar
Odum, E. P. 1969. The strategy of ecosystem development. Science, 164, 262–270.Google Scholar
Odum, H. T. 1973. Energy, ecology and economics. Ambio, 2, 220–227.Google Scholar
Odum, H. T. 1988. Self-organization, transformity, and information. Science, 242, 1132–1139.Google Scholar
Odum, H. T., and Pinkerton, R. C. 1955. Time's speed regulator: the optimum efficiency for maximum power output in physical and biological systems. Am. Sci., 43, 331–343.Google Scholar
Oke, T R. 1987. Boundary Layer Climates. London: Methuen.
Oki, T, and Kanae, S. 2006. Global hydrological cycle and world water resources. Science, 313, 1068–1072.Google Scholar
Oort, A. H. 1989. Angular momentum cycle in the atmosphere-ocean-solid Earth system. Bull. Amer. Meteorol. Soc., 70, 1231–1242.Google Scholar
Oort, A. H., Ascher, S. C., Levitus, S., and Peixoto, J. P. 1989. New estimates of the available potential energy in the world ocean. J. Geophys. Res., 94, 3187–3200.Google Scholar
Oort, A. H., Anderon, L. A., and Peixoto, J. P. 1994. Estimates of the energy cycle of the oceans. J. Geophys. Res., 99, 7665–7688.Google Scholar
Ornstein, L, Aleinov, I, and Rind, D. 2009. Irrigated afforestation of the Sahara and Australian Outback to end global warming. Clim. Change, 97, 409–437.Google Scholar
Ostwald, W. 1909. Energetische Grundlagen der Kulturwissenschaften. Leipzig: Klinkhardt.
Ou, H. W. 2001. Possible bounds on the Earth's surface temperature: from the perspective of a conceptual global-mean model. J. Clim., 14, 2976–2988.Google Scholar
Ou, H. W. 2006. Thermal properties of a coupled ocean-atmosphere: a conceptual model. Tellus, 58(3), 404–415.Google Scholar
Ou, H. W. 2007. Hydrological cycle and ocean stratification in a coupled climate system: a theoretical study. Tellus, 59A, 683–694.Google Scholar
Ozawa, H., and Ohmura, A. 1997. Thermodynamics of a global-mean state of the atmosphere: a state of maximum entropy increase. J. Clim., 10, 441–445.Google Scholar
Ozawa, H., Shimokawa, S., and Sakuma, H. 2001. Thermodynamics of fluid turbulence: A unified approach to the maximum transport properties. Phys. Rev. E, 64, 026303.Google Scholar
Ozawa, H., Ohmura, A., Lorenz, R. D., and Pujol, T. 2003. The second law of thermodynamics and the global climate system: a review of the maximum entropy production principle. Rev. Geophys., 41, 1018.Google Scholar
Paltridge, G. W. 1975. Global dynamics and climate: a system of minimum entropy exchange. Q. J. Roy. Meteorol. Soc., 101, 475–484.Google Scholar
Paltridge, G. W. 1978. The steady-state format of global climate. Q. J. Roy. Meteorol. Soc., 104, 927–945.Google Scholar
Paltridge, G. W. 1979. Climate and thermodynamic systems of maximum dissipation. Nature, 279, 630–631.Google Scholar
Pascale, S., Gregory, J. M., Ambaum, M. H. P., and Tailleux, R. 2012. A parametric sensitivity study of entropy production and kinetic energy dissipation using the FAMOUS AOGCM. Clim. Dyn., 38, 1211–1227.Google Scholar
Pascale, S., Ragone, F., Lucarini, V., Wang, Y., and Boschi, R. 2013. Nonequilibrium thermodynamics of circulation regimes in optically thin, dry atmospheres. Planet. Space Sci., 84, 48–65.Google Scholar
Pauluis, O., and Dias, J. 2012. Satellite estimates of precipitation-induced dissipation in the atmosphere. Science, 335, 953–956.Google Scholar
Pauluis, O., and Held, I.M. 2002a. Entropy budget of an atmosphere in radiative convective equilibrium. Part I: maximum work and frictional dissipation. J. Atmos. Sci., 59, 126–139.Google Scholar
Pauluis, O., and Held, I.M. 2002b. Entropy budget of an atmosphere in radiative convective equilibrium. Part II: latent heat transport and moist processes. J. Atmos. Sci., 59, 140–149.Google Scholar
Pauluis, O., Balaji, V., and Held, I. M. 2000. Frictional dissipation in a precipitating atmosphere. J. Atmos. Sci., 57, 987–994.Google Scholar
Peixoto, J. P., and Oort, A. H. 1992. Physics of Climate. New York: American Institute of Physics.
Peixoto, J. P., Oort, A. H., de Almeida, M., and Tome, A. 1991. Entropy budget of the atmosphere. J. Geophys. Res., 96, 10,981–10,988.Google Scholar
Penman, H. L. 1948. Natural evaporation from open water bare soil and grass. Proc. R. Soc. London A, 193, 120–146.Google Scholar
Petela, R. 1964. Exergy of heat radiation. J. Heat Transfer, 86, 187–192.Google Scholar
Petela, R. 2003. Exergy of undiluted thermal radiation. Solar Energy, 74, 469–488.Google Scholar
Pierrehumbert, R. T. 2002. The hydrologic cycle in deep-time climate problems. Nature, 419, 191–198.Google Scholar
Pimentel, D., Hurd, L. E., Bellotti, A. C., Forster, M. J., Oka, I. N., Sholes, O. D., and Whitman, R. J. 1973. Food production and the energy crisis. Science, 182, 443–449.Google Scholar
Planck, M. 1906. Theorie der Wärmestrahlung. Leipzig: Barth.
Porada, P., Kleidon, A., and Schymanski, S. J. 2011. Entropy production of soil hydrological processes and its maximization. Earth Syst. Dynam., 2, 179–190.Google Scholar
Press, W. H. 1976. Theoretical maximum for energy from direct and diffuse sunlight. Nature, 264, 734–735.Google Scholar
Price, C., Penner, J., and Prather, M. 1997. NOx from lightning. 1: global distribution based on lightning physics. J. Geophys. Res., 102, 5929–5941.Google Scholar
Priestley, C. H. B., and Taylor, R. J. 1972. On the assessment of surface heat flux and evaporation using large-scale parameters. Mon. Wea. Rev., 100, 81–92.Google Scholar
Prigogine, I. 1947. Etude Thermodynamique des Phenomenes Irreversibles. Liege Belgium: Desoer.
Prigogine, I. 1962. Introduction to Non-equilibrium Thermodynamics. New York: Wiley Interscience.
Prigogine, I. 1978. Time, structure, and fluctuations. Science, 201, 777–785.Google Scholar
Prigogine, I., Nicolis, G., and Babloyantz, A. 1972. Thermodynamics of evolution. Phys. Today, 25, 23–28.Google Scholar
Rabinowitch, E., and Govindjee, . 1969. Photosynthesis. New York: Wiley.
Radmer, R., and Kok, B. 1977. Photosynthesis: limited yields, unlimited dreams. Bioscience, 27, 599–605.Google Scholar
Ramanathan, V., Cess, R. D., Harrison, E. F., Minnis, P., Barkstrom, B. R., Ahmad, E., and Hartmann, D. 1989. Cloud-radiative forcing and climate: results from the Earth radiation budget experiment. Science, 243, 57–63.Google Scholar
Ramankutty, N., Evan, A. T., Monfreda, C., and Foley, J. A. 2008. Farming the planet: 1. Geographic distribution of global agricultural lands in the year 2000. Global Biogeochem. Cy., 22, GB1003.Google Scholar
Raupach, M. R. 1998. Influences of local feedbacks on land-air exchanges of energy and carbon. Glob. Ch. Biol., 4, 477–494.Google Scholar
Renno, N O, and Ingersoll, A P. 1996. Natural convection as a heat engine: a theory for CAPE. J. Atmos. Sci., 53, 572–585.Google Scholar
Retallack, G. J. 2001. Cenozoic expansion of grasslands and climatic cooling. J. Geology, 109, 407–426.Google Scholar
Retallack, G. J. 2007. Coevolution of life and Earth. Pages 295–230 of: Schubert, G. (ed), Treatise of Geophysics. Amsterdam: Elsevier.
Rietkerk, M., Dekker, S., Wassen, M., Verkroost, A., and Bjerkens, M. 2004. A putative mechanism for bog patterning. Am. Nat., 163, 699–708.Google Scholar
Rinaldo, A., Rodriguez-Iturbe, I., Rigon, R., Bras, R. L., Ijjasz-Vasquez, E., and Marani, A. 1992. Minimum energy and fractal structures of drainage networks. Water Resour. Res., 28, 2183–2195.Google Scholar
Rinaldo, A., Rodriguez-Iturbe, I., and Rigon, R. 1998. Channel networks. Ann. Rev. Earth Planet. Sci., 26, 289–327.Google Scholar
Rochetin, N., Lintner, B. R., Findell, K. L., Sobel, A. H., and Gentine, P. 2014. Radiative-convective equilibrium over a land surface. J. Clim., 27, 8611–8629.Google Scholar
Rockström, J., Steffen, W., Noone, K., Persson, A., Chapin, F. S., Lambin, E. F., Lenton, T. M., Scheffer, M., Folke, C., Schellnhuber, H. J., Nykvist, B., de Wit, C. A., Hughes, T., van der Leeuw, S., Rodhe, H., Sörlin, S., Snyder, P. K., Constanza, R., Svedin, U., Falkenmark, M., Karlberg, L., Corell, R. W., Fabry, V. J., Hansen, J., Walker, B., Liverman, D., Richardson, K., Crutzen, P., and Foley, J. A. 2009. A safe operating space for humanity. Nature, 461(472–475).Google Scholar
Rodgers, C. D. 1976. Minimum entropy exchange principle: reply. Q. J. Roy. Meteorol. Soc., 102, 455–457.Google Scholar
Rodriguez-Iturbe, I., and Rinaldo, A. 1997. Fractal River Basins: Chance and Self- Organization. Cambridge UK: Cambridge University Press.
Rogner, H. H., Barthel, F., Cabrera, M., Faaij, A., Girouc, M., Hall, D., Kagramanian, V., Kononov, S., Lefevre, T., Moreira, R., Nötstaller, R., Odell, P., and Taylor, M. 2000. Energy resources. In: World Energy Assessment: Energy and the Challenge of Sustainability. New York, USA: United Nations Development Programme, United Nations Department of Economic and Social Affairs, and World Energy Council.
Rojstaczer, S., Sterling, S. M., and Moore, N. J. 2001. Human appropriation of photosynthesis products. Science, 294, 2549–2552.Google Scholar
Romps, D. M., Seeley, J. T., Vallaro, D., and Molinari, J. 2014. Projected increase in lightning strikes in the United States due to global warming. Science, 346, 851–854.Google Scholar
Romps, David M., and Charn, Alexander B. 2015. Sticky thermals: evidence for a dominant balance between buoyancy and drag in cloud updrafts. J. Atmos. Sci., 72, 2890–2901.Google Scholar
Rosen, M. A., and Scott, D. S. 2003. Entropy production and exergy destruction: Part I - hierarchy of Earth's major constituencies. Int. J. Hydrogen Energ., 28, 1307–1313.Google Scholar
Rosen, P. 1954. Entropy of radiation. Phys. Rev., 96, 555.Google Scholar
Rosenzweig, M. L. 1968. Net primary productivity of terrestrial communities: prediction from climatological data. Am. Nat., 102, 67–74.Google Scholar
Rosing, M. T., Bird, D. K., Sleep, N. H., Glassley, W., and Albarede, F. 2006. The rise of continents: an essay on the geologic consequences of photosynthesis. Paleogeogr. Paleocl., 232, 99–113.Google Scholar
Ross, R. T. 1966. Thermodynamic limitations on the conversion of radiant energy into work. J. Chem. Phys., 45, 1–7.Google Scholar
Rotenberg, E., and Yakir, D. 2010. Contribution of semi-arid forests to the climate system. Science, 327, 451–454.Google Scholar
Rotenberg, E., and Yakir, D. 2011. Distinct patterns of changes in surface energy budget associated with forestation in the semiarid region. Glob. Ch. Biol., 17, 1536–1548.Google Scholar
Rubin, D. M., and Hunter, R. E. 1987. Bedform alignment in directionally varying flows. Science, 237, 276–278.Google Scholar
Ruimy, A., Jarvis, P. G., Baldocchi, D. D., and Saugier, B. 1995. CO2 fluxes over plant canopies and solar radiation: a review. Adv. Ecol. Res., 26, 1–68.Google Scholar
Russell, M. J., and Hall, A. J. 1997. The emergence of life from iron monosulphide bubbles at a submarine hydrothermal redox and pH front. J. Geol. Soc. London, 154, 377–402.Google Scholar
Sagan, C., and Mullen, G. 1972. Earth and Mars: evolution of atmospheres and surface temperatures. Science, 177, 52–56.Google Scholar
Schlaich, J., Bergermann, R., Schiel, W., and Weinrebe, G. 2005. Design of commercial solar updraft tower systems: utilization of solar induced convective flows for power generation. J. Solar Energy Eng., 127, 117–124.Google Scholar
Schlesinger, W H. 1997. Biogeochemistry: An Analysis of Global Change. San Diego: Academic Press.
Schmidt, W. 1915. Strahlung und Verdunstung an freien Wasserflächen; ein Beitrag zum Wärmehaushalt des Weltmeeres und zum Wasserhaushalt der Erde. Ann. d. Hydrogr. u. maritimen Meteorol., 43, 111–178.Google Scholar
Schneider, E. D., and Kay, J. J. 1994a. Complexity and thermodynamics: towards a new ecology. Futures, 26, 626–647.Google Scholar
Schneider, E. D., and Kay, J. J. 1994b. Life as a manifestation of the second law of thermodynamics. Math. Comput. Modeling, 19, 25–48.Google Scholar
Schrenk, M. O., Brazelton, W. J., and Lang, S. Q. 2013. Serpentinization carbon, and deep life. Rev. Mineral. Geochem., 75, 575–606.Google Scholar
Schrödinger, E. 1944. What is Life? The Physical Aspect of the Living Cell. Cambridge, UK: Cambridge University Press.
Schumann, U., and Huntrieser, H. 2007. The global lightning-induced nitrogen oxides source. Atmos. Chem. Phys., 7, 3823–3907.Google Scholar
Schwartzman, D. W., and Middendorf, G. 2000. Biospheric cooling and the emergence of intelligence. Pages 425–429 of: Lemarchand, G., and Meech, K. (eds), A New Era in Bioastronomy. ASP Conference Series, Vol. 213.Google Scholar
Schwartzman, D. W., and Volk, T. 1989. Biotic enhancement of weathering and the habitability of Earth. Nature, 340, 457–460.Google Scholar
Schymanski, S. J., Kleidon, A., Stieglitz, M., and Narula, J. 2010. Maximum entropy production allows a simple representation of heterogeneity in semiarid ecosystems. Phil. Trans. R. Soc. B, 365(1449–1455).Google Scholar
Seager, S. 2013. Exoplanet habitability. Science, 340, 577–581.Google Scholar
Shock, E. L., and Holland, M. E. 2007. Quantitative habitability. Astrobiology, 7, 839–851.Google Scholar
Shukla, J., and Mintz, Y. 1982. The influence of land-surface-evapotranspiration on the Earth's climate. Science, 247, 1322–1325.Google Scholar
Shutts, G. J. 1981. Maximum entropy production states in quasi-geostrophic dynamical models. Q. J. Roy. Meteorol. Soc., 107(453), 503–520.Google Scholar
Simoncini, E., Russell, M. J., and Kleidon, A. 2011. Modeling free energy availability from Hadean hydrothermal systems to first metabolism. Orig. Life Evol. Biosph., 41, 529–532.Google Scholar
Simoncini, E., Virgo, N., and Kleidon, A. 2013. Quantifying drivers of chemical disequilibrium: theory and application to methane in the Earth's atmosphere. Earth Syst. Dynam., 4, 317–331.Google Scholar
Smil, V. 1999. Energies: An Illustrated Guide to the Biosphere and Civilization. Cambridge, MA, USA: MIT Press.
Smil, V. 2000. Energy in the twentieth century: resources, conversions, costs, uses, and consequences. Annu. Rev. Energy Environ., 25, 21–51.Google Scholar
Smith, W. K., Cleveland, C. C., Reed, S. C., and Running, S. W. 2014. Agricultural conversion without external water and nutrient inputs reduces terrestrial vegetation productivity. Geophys. Res. Lett., 41, 449–455.Google Scholar
Sperry, J. S. 2000. Hydraulic constraints on plant gas exchange. Agric. Forest Meteorol., 104, 13–23.Google Scholar
Spohn, T. 2012. Planets and life. Spatium, 30, 3–15.Google Scholar
Stacey, F. D. 1967. Convecting mantle as a thermodynamic engine. Nature, 214, 476–477.Google Scholar
Stephens, G. L., and O'Brien, D. M. 1993. Entropy and climate. 1. ERBE observations of the entropy production of the Earth. Q. J. Roy. Meteorol. Soc., 119(509), 121–152.Google Scholar
Stephens, G. L., Li, J., Wild, M., Clayson, C. A., Loeb, N., Kato, S., L'Ecuyer, T., Stackhouse, P. W., Lebsock, M., and Andrews, T. 2012. An update on Earth's energy balance in light of the latest global observations. Nature Geosci., 691–696.Google Scholar
Stevens, B. 2005. Atmospheric moist convection. Annu. Rev. Earth Planet. Sci., 33, 605–643.Google Scholar
Stevens, B., and Bony, S. 2013. What are climate models missing?Science, 340, 1053–1054.Google Scholar
Stevenson, D. J. 1983. Planetary magnetic fields. Rep. Prog. Phys., 46, 555–620.Google Scholar
Stevenson, D. J., Spohn, T., and Schubert, G. 1983. Magnetism and thermal evolution of the terrestrial planets. Icarus, 54, 466–489.Google Scholar
Stone, P. H. 1978. Constraints on dynamical transports of energy on a spherical planet. Dynamics of Atmospheres and Oceans, 2, 123–139.Google Scholar
Stull, R B. 1989. An Introduction to Boundary Layer Meteorology. Boston: Kluwer Academic Press.
Tailleux, R. 2010. Entropy versus APE production: on the buoyancy input in the oceans energy cycle. Geophys. Res. Lett., 37, L22603.Google Scholar
Tailleux, R. 2013. Available potential energy and exergy in stratified fluids. Annu. Rev. Fluid Mech., 45, 35–58.Google Scholar
Tamura, S. T. 1905. Doctor Margules on the energy of storms. Mon.Wea. Rev., 33, 519–521.Google Scholar
Taylor, S. R., and McLennan, S. M. 1996. The evolution of continental crust. Sci. Am., 274, 76–81.Google Scholar
Tennekes, H. 1973. A model for the dynamics of the inversion above a convective boundary layer. J. Atmos. Sci., 30, 558–566.Google Scholar
Thompson, A. M. 1992. The oxidizing capacity of the Earth's atmosphere: probable past and future changes. Science, 256, 1157–1165.Google Scholar
Thomson, W. 1881. On the sources of energy in nature available to man for the production of mechanical effect. Science, 2, 475–478.Google Scholar
Trenberth, K. E., and Caron, J. M. 2001. Estimates of meridional atmosphere and ocean heat transports. J. Clim., 14, 3433–3443.Google Scholar
Ulanowicz, R. E., and Hannon, B. M. 1987. Life and the production of entropy. Proc. R. Soc. Lond. B, 232, 181–192.Google Scholar
Unrean, P., and Srienc, F. 2011. Metabolic networks evolve towards states of maximum entropy production. Metabolic Engineering, 13, 666–673.Google Scholar
Vallino, J. J. 2010. Ecosystem biogeochemistry considered as a distributed metabolic network ordered by maximum entropy production. Phil. Trans. R. Soc. B, 365, 1417–1427.Google Scholar
Vance, E. 2009. High hopes. Nature, 460, 564–566.Google Scholar
Vanyo, J. P., and Paltridge, G. W. 1981. A model for energy dissipation at the mantle-core boundary. Geophys. J., 66(3), 677–690.Google Scholar
Verhoogen, J. 1980. Energetics of the Earth. Washington DC: National Academy of Sciences.
Vitousek, P. M., Ehrlich, P. R., Ehrlich, A. H., and Matson, P. A. 1986. Human appropriation of the products of photosynthesis. Bioscience, 36, 368–373.Google Scholar
Vitousek, P. M., Mooney, H. A., Lubchenco, J., and Melillo, J.M. 1997. Human domination of Earth's ecosystems. Science, 277, 494–499.Google Scholar
Volk, T. 1998. Gaia's Body: Toward a Physiology of Earth. New York: Springer.
Volk, T, and Pauluis, O. 2010. It is not the entropy you produce, rather, how you produce it. Phil. Trans. R. Soc. B, 365, 1317–1322.Google Scholar
von Storch, J.-S., Eden, C., Fast, I., Haak, H., Hernandez-Deckers, D., Maier-Reimer, E., Marotzke, J., and Stammer, D. 2012. An estimate of the Lorenz energy cycle for the world ocean based on the 1/10” STORM/NCEP simulation. J. Phys. Oceanogr., 42, 2185–2205.Google Scholar
Walker, D. 1992. Energy, Plants and Man. Sausalito, CA, USA: University Science Books.
Walker, J. C., Hays, P. B., and Kasting, J. F. 1981. A negative feedback mechanism for the long-term stabilization of Earth's surface temperature. J. Geophys. Res., 86, 9776–9782.Google Scholar
Wallace, J. M., and Hobbs, P. V. 1977. Atmospheric Science: An Introductory Survey. New York: Academic Press.
Wang, J., Salvucci, G. D., and Bras, R. L. 2004. An extremum principle of evaporation. Water Resour. Res., 40, W09303.Google Scholar
Wang, J., Bras, R. L., Lerdau, M., and Salvucci, G. D. 2007. A maximum hypothesis of transpiration. J. Geophys. Res., 112, G03010.Google Scholar
Werner, B. T. 1999. Complexity in natural landform patterns. Science, 284, 102–104.Google Scholar
West, G. B., Brown, J. H., and Enquist, B. J. 1997. A general model for the origin of allometric scaling laws in biology. Science, 276, 122–126.Google Scholar
West, G. B., Brown, J. H., and Enquist, B. J. 1999. A general model for the structure and allometry of plant vascular systems. Nature, 400, 664–667.Google Scholar
Wheeler, P. E. 1991a. The influence of bipedalism on the energy and water budgets of early hominids. J. Human Evol., 21, 117–136.Google Scholar
Wheeler, P. E. 1991b. The thermoregulatory advantages of hominid bipedalism in open equatorial environments: the contribution of increased convective heat loss and cutaneous evaporative cooling. J. Human Evol., 21, 107–115.Google Scholar
Wheeler, R. M., Mackowiak, C. L., Stutte, G. W., Yorio, N. C., Ruffe, L. M., Sager, J. C., Prince, R. P., and Knott, W. M. 2008. Crop productivities and radiation use efficiencies for bioregenerative life support. Adv. Space Res., 41, 706–713.Google Scholar
Wildt, R. 1956. Radiative transfer and thermodynamics. Astrophys. J., 123, 107–116.Google Scholar
Wildt, R. 1972. Thermodynamics of the gray atmosphere. IV. Entropy transfer and production. Astrophys. J., 174, 69–77.Google Scholar
Williamson, S. C., Detling, J. K., Dodd, J. L., and Dyer, M. I. 1989. Experimental evaluation of the grazing optimization hypothesis. J. Range Manage., 42, 149–152.Google Scholar
Wofsy, S. C., Harriss, R. C., and Kaplan, W. A. 1988. Carbon dioxide in the atmosphere over the Amazon basin. J. Geophys. Res., 93, 1377–1387.Google Scholar
Wong, S. C., Cowan, I. R., and Farquhar, G. D. 1979. Stomatal conductance correlates with photosynthetic capacity. Nature, 282, 424–426.Google Scholar
Woodward, F. I. 1987. Stomatal numbers are sensitive to increases in CO2 from pre-industrial levels. Nature, 327, 617–618.Google Scholar
World Commission on Environment and Development. 1987. Our Common Future. Oxford: Oxford University Press.
Wu, W, and Liu, Y. 2010. Radiation entropy flux and entropy production of the Earth system. Rev. Geophys., 48, RG2003.Google Scholar
Xu, C.-Y., and Singh, V. P. 2000. Evaluation and generalization of radiation-based methods for calculating evaporation. Hydrol. Process., 14, 339–349.Google Scholar
Xu, K.-M., Arakawa, A., and Krueger, S. K. 1992. The macroscopic behavior of cumulus ensembles simulated by a cumulus ensemble model. J. Atmos. Sci., 49, 2402–2420.Google Scholar
Yen, J. D. L., Paganin, D. M., Thomson, J. R., and Macnally, R. 2014. Thermodynamic extremization principles and their relevance to ecology. Austral Ecology, 39, 619–632.Google Scholar
Zehe, E., Ehret, U., Blume, T., Kleidon, A., Scherer, U., and Westhoff, M. 2013. A thermodynamic approach to link self-organization preferential flow and rainfall– runoff behaviour. Hydrol. Earth Syst. Sci., 17, 4297–4322.Google Scholar
Zener, C. 1973. Solar sea power. Phys. Today, 26, 48–53.Google Scholar
Zhou, X., Wang, F., and Ochieng, R. M. 2010. A review of solar chimney power technology. Renew. Sust. Energ. Rev., 14, 2315–2338.Google Scholar
Zhu, X.-G., Long, S. P., and Ort, D. R. 2008. What is the maximum efficiency with which photosynthesis can convert solar energy into biomass?Curr. Opin. Biotechnol., 19, 153–159.Google Scholar
Zotin, A. I. 1984. Bioenergetic trends of evolutionary progress of organisms. Pages 451–458 of: Lamprecht, I., and Zotin, A. I. (eds), Thermodynamics and Regulation of Biological Processes. Berlin New York: de Gruyter.

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
  • Axel Kleidon, Max-Planck-Institut für Biogeochemie, Jena
  • Book: Thermodynamic Foundations of the Earth System
  • Online publication: 05 March 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139342742.015
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
  • Axel Kleidon, Max-Planck-Institut für Biogeochemie, Jena
  • Book: Thermodynamic Foundations of the Earth System
  • Online publication: 05 March 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139342742.015
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
  • Axel Kleidon, Max-Planck-Institut für Biogeochemie, Jena
  • Book: Thermodynamic Foundations of the Earth System
  • Online publication: 05 March 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139342742.015
Available formats
×