Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-tdptf Total loading time: 0 Render date: 2024-08-06T11:30:05.480Z Has data issue: false hasContentIssue false

9 - Evapotranspiration

from Part III - Coupling Hillslope Geomorphology, Soils, Hydrology, and Ecosystems

Published online by Cambridge University Press:  27 October 2016

T. Andrew Black
Affiliation:
University of British Columbia
Rachhpal S. Jassal
Affiliation:
University of British Columbia
Edward A. Johnson
Affiliation:
University of Calgary
Yvonne E. Martin
Affiliation:
University of Calgary
Get access

Summary

Introduction

The Process of Evapotranspiration

Evapotranspiration (E) is the combined processes of physical evaporation and biological transpiration, by which liquid water from open water, soil, and vegetation surfaces is transformed into vapor and transported into the atmosphere. For terrestrial ecosystems, these three components are referred to as evaporation from bare soil (E s ), transpiration or water use by growing vegetation (E t), and evaporation from wet foliage of intercepted water (E i ). Sublimation, the transformation of solid water (i.e., ice or snow) to water vapor, though generally considered separate from evapotranspiration, is considered here as part of E s or E i as the case may be.

Energy is required to change the state of the molecules of water from liquid to vapor. This energy is known as the latent heat of vaporization of water (λ), and is provided by solar irradiance (R s ) and, to a lesser extent, the ambient temperature of the air (T a ). Also, as the liquid water intermolecular attractions decrease with the increase in temperature, λ decreases slightly with increasing temperature. The driving force to remove water vapor from the evaporating surface is the difference between the water vapor pressure at the evaporating surface and that of the surrounding atmosphere. As evaporation proceeds, the surrounding air becomes gradually saturated and the process will slow down and might stop if the wet air is not transferred to the atmosphere. As the replacement of the saturated air with drier air depends greatly on wind speed, air humidity and wind speed are also important controls of E. In the case of transpiration, liquid water, together with some nutrients, is taken up by the roots and transported through the plant to the leaves. While conversion of liquid water contained in plant tissues to water vapor occurs in the stomatal cavities of the leaf, vapor exchange with the atmosphere occurs through stomata. Thus, energy is required for evapotranspiration (evaporation from wet leaves and transpiration) to proceed from vegetation. In fact, the concept of evapotranspiration can be simplified as “water + energy = evapotranspiration.”

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

Allen, R. G., Kjaersgaard, J., Garcia, M., Tasumi, M. and Trezza, R. (2008). Fine-tuning components of inverse-calibrated, thermal-based remote sensing models for evapotranspiration. In The Future of Land Imaging: The 17th William T. Pecora Memorial Remote Sensing Symposium, November 16–20, Denver, Colarado. Bethesda, MD: American Society for Photogrammetry and Remote Sensing, 11 p.
Allen, R. G., Pereira, L. S., Howell, T. A. and Jensen, M. E. (2011). Evapotranspiration information reporting: I. Factors governing measurement accuracy. Agricultural Water Management, 98, 899–920.CrossRefGoogle Scholar
Allen, R. G., Pereira, L. S., Raes, D. and Smith, M. (1998). Crop Evapotranspiration: Guidelines for Computing Crop Water Requirements. FAO Irrigation and Drainage Paper No. 56, Rome, Italy.
Allen, R. G., Tasumi, M. and Trezza, R. (2007). Satellite-based energy balance for Mapping Evapotranspiration with Internalized Calibration (METRIC) model. Journal of Irrigation and Drainage Engineering 133, 380–94.CrossRefGoogle Scholar
Amiro, B. D., Barr, A. G., Black, T. A. et al. (2006). Carbon, energy and water fluxes at mature and disturbed forest sites, Saskatchewan, Canada. Agricultural and Forest Meteorology, 136, 237–51.CrossRefGoogle Scholar
Andreadis, K. M. and Lettenmaier, D. P. (2006). Trends in 20th century drought over the continental United States. Geophysical Research Letters, 33, doi:10.1029/2006GL025711.CrossRefGoogle Scholar
Anderson, R. G. and Goulden, M. L. (2009). A mobile platform to constrain regional estimates of evapotranspiration. Agricultural and Forest Meteorology, 149, 771–82.CrossRefGoogle Scholar
Anderson, M. C., Norman, J. M., Kustas, W. P., Li, F., Prueger, J. H. and Mecikalsi, J. R. (2005). Effects of vegetation clumping on two-source model estimates of surface energy fluxes from an agricultural landscape during SMACEX. Journal of Hydrometeorology, 6, 892–909.CrossRefGoogle Scholar
Bai, J., Chen, X., Li, K., Luo, G. and Yu, Q. (2014). Quantifying the contributions of agricultural oasis expansion, management practices and climate change to net primary production and evapotranspiration in croplands in arid northwest China. Journal of Arid Environments, 100, 31–41.CrossRefGoogle Scholar
Baldocchi, D. D. (1992). A Lagrangian random-walk model for simulating water vapor. CO2 and sensible heat flux densities and scalar profiles over and with a soybean canopy. Boundary Layer Meteorology, 61, 113–44.CrossRefGoogle Scholar
Baldocchi, D. D. (2003). Assessing the eddy covariance technique for evaluating carbon dioxide exchange rates of ecosystems: past, present and future. Global Change Biology, 9, 479–92.CrossRefGoogle Scholar
Baldocchi, D. D. (2008). Breathing of the terrestrial biosphere: lessons learned from a global network of carbon dioxide flux measurement systems. Australian Journal of Botany, 56, 1–26.CrossRefGoogle Scholar
Baldocchi, D. D., Falge, E., Gu, L. et al. (2001). FLUXNET: a new tool to study the temporal and spatial variability of ecosystem-scale carbon dioxide, water vapor, and energy flux densities. Bulletin of the American Meteorological Society, 82, 2415–34.2.3.CO;2>CrossRefGoogle Scholar
Baldocchi, D. D., Kelliher, F. M., Black, C. A. and Jarvis, P. G. (2000). Climate and vegetation controls on boreal zone energy exchange. Global Change Biology, 6, 69–83.CrossRefGoogle Scholar
Baldocchi, D. D. and Xu, L. (2007). What limits evaporation from Mediterranean oak woodlands –the supply of moisture in the soil, physiological control by plants or the demand by the atmosphere? Advances in Water Resources, 30, 2113–22.Google Scholar
Ball, J. T., Woodrow, I. E. and Berry, J. A. (1987). A model predicting stomatal conductance and its contribution to control of photosynthesis under different environmental conditions. In Progress in Photosynthesis Research, ed. Biggins, J.. Zoetermeer, Netherlands: Martinus Nijhof, pp. 221–34.CrossRef
Barr, A. G., Black, T. A., Hogg, E. H. et al. (2006). Climatic controls on the carbon and water balances of a boreal aspen forest, 1994– 2003. Global Change Biology, 12, 1–16.Google Scholar
Barr, A. G., van der Kamp, G., Black, T. A., McCaughey, H. and Nesic, Z. (2012). Energy balance closure at the BERMS flux towers in relation to the water balance of the White Gull Creek watershed 1999–2009. Agricultural and Forest Meteorology, 153, 3–13.CrossRefGoogle Scholar
Bastiaanssen, W. G. M., Noordman, E. J. M., Pelgrum, H., Davids, G., Thoreson, B. P. and Allen, R. G. (2005). SEBAL model with remotely sensed data to improve water resources management under actual field conditions. Jouranl of Irrigation and Drainage Engineering, 131, 85–93.CrossRefGoogle Scholar
Beer, C., Ciais, P., Reichstein, M. et al. (2009). Temporal and among-site variability of inherent water use efficiency at the ecosystem level. Global BiogeochemicalCycles 23, GB2018.CrossRefGoogle Scholar
Beer, C., Reichstein, M., Ciais, P., Farquhar, G. D. and Papale, D. (2007). Mean annual GPP of Europe derived from its water balance. Geophysical Research Letters, 34, doi:10.1029/2006GL029006.CrossRef
Ben-Gal, A. and Shani, U. (2003). Water use and yield of tomatoes under limited water and excess boron. Plant and Soil, 256, 170–86.CrossRefGoogle Scholar
Bergengren, J. C., Waliser, D. E. and Yung, Y. L. (2011). Ecological sensitivity: a biospheric view of climate change. Climatic Change, 107, 433–57.CrossRefGoogle Scholar
Betts, R. A., Boucher, O., Collins, M. and Cox, P. (2007). Projected increase in continental runoff due to plant responses to increasing carbon dioxide. Nature, 448, 1037–41.CrossRefGoogle Scholar
Black, T. A. (1979) Evapotranspiration from Douglas-fir stands exposed to soil water deficits. Water Resources Research, 15, 164–70.CrossRefGoogle Scholar
Black, T. A., Gardner, W. R. and Thurtell, G. W. (1969). Prediction of evaporation, drainage, and soil water storage for a bare soil. Soil Science Society of America, Proceedings, 33, 655–60.CrossRefGoogle Scholar
Black, T. A., Thurtell, G. W. and Tanner, C. B. (1968). Hydraulic load cell lysimeter, construction, calibration and tests. Soil Science Society of America, Proceedings, 32, 623–9.CrossRefGoogle Scholar
Blaney, H. F. and Criddle, W. D. (1950). Determining Water Requirements in Irrigated Areas from Climatological Irrigation Data. Technical Paper No. 96. Washington, DC: US Department of Agriculture, Soil Conservation Service.
Blanken, P. D. and Black, T. A. (2004). The canopy conductance of a boreal aspen forest, Prince Albert National Park, Canada. Hydrological Processes, 18, 1561–78.CrossRefGoogle Scholar
Blanken, P. D., Black, T. A., Neumann, H. et al. (2001). The seasonal water and energy exchange above and within a boreal aspen forest. Journal of Hydrology, 245, 118–36.CrossRefGoogle Scholar
Bond-Lamberty, B, Peckham, S. D., Gower, S. T. and Ewer, B. E. (2009). Effects of fire on regional evapotranspiration in the central Canadian boreal forest. Global Change Biology, 15, 1242–54.CrossRefGoogle Scholar
Bowen, I. S. (1926). The ratio of heat losses by conduction and by evaporation from any water surface. Physics Review, 27, 779–87.CrossRefGoogle Scholar
Brazdil, R., Chroma, K., Dobrovolný, P. and Tolasz, R. (2009). Climate fluctuations in the Czech Republic during the period 1961–2005. International Journal of Climatology, 29, 223–42.CrossRefGoogle Scholar
Brix, H. (1981). Mechanisms of response to fertilization. II. Utilization by trees and stands. In Improving Forest Fertilization Decision-Making in British Columbia. Victoria, BC: British Columbia Ministry of Forestry, pp. 76–93.
Brown, M., Black, T. A., Nesic, Z. et al. (2013). Evapotranspiration and canopy characteristics of two lodgepole pine stands following mountain pine beetle attack. Hydrological Processes, 28, doi:10.1002/hyp.9870.CrossRefGoogle Scholar
Bouchet, R. J. (1963). Evapotranspiration reelle et potentielle, signification climatique. IAHS Publication, 62, 134–42.Google Scholar
Brümmer, C., Black, T. A., Jassal, R. S. et al. (2011). How climate and vegetation type influence evapotranspiration and water use efficiency in Canadian forest, peatland and grassland ecosystems. Agricultural and Forest Meteorology, 153, doi:10.1016/j.agrformet.2011.04.008.CrossRefGoogle Scholar
Brutsaert, W. (1982). Evaporation into the Atmosphere: Theory, History, and Applications. Dordrecht, Netherlands: Reidel.CrossRef
Brutsaert, W. (2006). Indications of increasing land surface evaporation during the second half of the 20th century. Geophysical Research Letters, 33, doi:10.1029/2006GL027532.CrossRefGoogle Scholar
Brutsaert, W. and Parlange, M. B. (1998). Hydrologic cycle explains the evaporation paradox. Nature, 396, 30–2.CrossRefGoogle Scholar
Brutsaert, W. and Stricker, H. (1979). An advection-aridity approach to estimate actual regional evaporation. Water Resources Research, 15, 443–50.CrossRefGoogle Scholar
Buckley, T. N. (2005). The control of stomata by water balance. New Phytolologist, 168, 275–91.CrossRefGoogle Scholar
Budyko, M. I. (1974). Climate and Life. New York: Academic Press.
Burba, G. G. and Anderson, D. J. (2010). A Brief Practical Guide to Eddy Covariance Flux Measurements: Principles and Workflow Examples for Scientific and Industrial Applications. Lincoln, NB: LI-COR Biosciences.
Burba, G. G. and Verma, S. B. (2005). Seasonal and interannual variability in evapotranspiration of native tallgrass prairie and cultivated wheat ecosystems. Agricultural and Forest Meteorology, 135, 190–201.CrossRefGoogle Scholar
Businger, J. A. (1975). Aerodynamics of vegetative surfaces. In Heat and Mass Transfer in the Biosphere: Part 1. Transfer Processes in the Plant Environment, ed. deVries, D. A. and Afgan, N. H.. New York: John Willey, pp. 139–65.
Camillo, P. J., Gurney, R. J. and Schmugge, T. J. (1983). A soil and atmospheric boundary layer model for evapotranspiration and soil moisture studies. Water Resources Research, 19, 371–80.CrossRefGoogle Scholar
Carlson, T. (2007). An overview of the “triangle method” for estimating surface evapotranspiration and soil moisture from satellite imagery. Sensors, 7, 1612–29.CrossRefGoogle Scholar
Cao, M. K. and Woodward, F. I. (1998). Dynamic responses of terrestrial ecosystem carbon cycling to global climate change. Nature, 393, 249–52.CrossRefGoogle Scholar
Chen, J., Chen, B., Black, T.A. et al. (2013). Comparison of terrestrial evapotranspiration estimates using the mass transfer and Penman–Monteith equations in land surface models. Journal of Geophysical Research: Biogeosciences, 118, 1715–31.CrossRefGoogle Scholar
Choudhury, B. J., Ahmed, N. U., Idaho, S. B., Reginato, R. J. and Daughtry, C. S. T. (1994). Relations between evaporation coefficients and vegetation indices: studies by model simulation. Remote Sensing of Environment, 50, 1–17.CrossRefGoogle Scholar
Choudhury, B. J. and DiGirolamo, N. E. (1998). A biophysical process-based estimate of global land surface evaporation using satellite and ancillary data. I. Model description and comparison with observations. Journal of Hydrology, 205, 164–85.CrossRefGoogle Scholar
Choudhury, B. J., DiGirolamo, N. E., Susskind, J. et al. (1998). A biophysical process-based estimate of global land surface evaporation using satellite and ancillary data. II. Regional and global patterns of seasonal and annual variations. Journal of Hydrology, 205, 186–204.CrossRefGoogle Scholar
Chung, S. O. and Horton, R. (1987). Soil heat and water flow with a partial surface mulch. Water Resources Research, 12, 2175–86.CrossRefGoogle Scholar
Clark, K., Skowronski, N., Gallagher, M., Renninger, H. and Schäfer, K. (2012). Effects of invasive insects and fire on forest energy exchange and evapotranspiration in the New Jersey pinelands. Agricultural and Forest Meteorology, 166, 50–61.CrossRefGoogle Scholar
Classen, A. T., Hart, S. C., Whitham, T. G., Cobb, N. S. and Koch, G. W. (2005). Insect infestations linked to shifts in microclimate: important climate change implications. Soil Science Society of America Journal, 69, 2049–57.CrossRefGoogle Scholar
Clausnitzer, F., Köstner, B., Schwärzel, K. and Bernhofer, C. (2011). Relationships between canopy transpiration, atmospheric conditions and soil water availability—analyses of long-term sap-flow measurements in an old Norway spruce forest at the Ore Mountains/Germany. Agricultural and Forest Meteorology, 151, 1023–34.CrossRefGoogle Scholar
Cleugh, H. A., Leuning, R., Mu, Q. and Running, S. W. (2007). Regional evaporation estimates from flux tower and MODIS satellite data. Remote Sensing of Environment, 106, 285–304.CrossRefGoogle Scholar
Costa, M. H., Botta, A. and Cardille, J. A. (2003). Effect of large-scale changes in land cover on the discharge of the Tocantins river, southwestern Amazonia. Journal of Hydrology, 283, 206–17.CrossRefGoogle Scholar
Crago, R. and Brutsaert, W. (1996). Daytime evaporation and the self-preservation of the evaporative fraction and the Bowen ratio. Journal of Hydrology, 178, 241–55.CrossRefGoogle Scholar
Dai, A., Karl, T. R., Sun, B. and Trenberth, K. E. (2006). Recent trends in cloudiness over the United States: a tale of monitoring inadequacies. Bulletin of the American Meteorological Society, 87, 597–606.CrossRefGoogle Scholar
Dai, Y., Zeng, X., Dickinson, R. E., Baker, I. and Bonan, G. B. (2001). Common land model: technical documentation and user's guide. Bulletin of the American Meteorological Society, 84, 1113–23.Google Scholar
Dale, V. H., Joyce, L. A., McNulty, S. et al. (2001). Climate change and forest disturbances. BioScience, 51, 723–34.CrossRefGoogle Scholar
Dalton, J. (1802). Experimental essays on the constitution of mixed gases: on the force of steam or vapour from water or other liquids in different temperatures, both in a Torricelli vacuum and in air; on evaporation; and on expansion of gases by heat. Proceedings of the Manchester Literary and Philosophical Society, 5, 536–602.Google Scholar
David, T. S., Ferreira, M. I., Cohen, S., Pereira, J. S. and David, J. S. (2004). Constraints on transpiration from an evergreen oak tree in southern Portugal, Agricultural and Forest Meteorology, 122, 193–205.Google Scholar
Dingman, S. L. (1992). Physical Hydrology. Upper Saddle River, NJ: Prentice Hall.
Dirmeyer, P. A., Gao, X., Zha, M. et al. (2005). The Second Global Soil Wetness Project (GSWP- 2): Multi-model Analysis and Implications for our Perception of the Land Surface, COLA Technical Report 185. Calverton, MD: Center for Ocean-Land-Atmosphere Studies.
Donovan, L. A. and Ehleringer, J. R. (1991). Ecophysiological differences among juvenile and reproductive plants of several woody species. Oecologia, 86, 594–97.CrossRefGoogle Scholar
Dessler, A. E., Schoeber, M. R., Wang, T., Davis, S. M. and Rosenlof, K. H. (2013). Stratospheric water vapor feedback. Proceedings of the National Academy of Sciences, 110, 18087–91, doi:10.1073/pnas.1310344110.CrossRefGoogle Scholar
Dunin, F. X., Reyenga, W. and McIlroy, I. C. (1991). Australian lysimeter studies of field evaporation. In Lysimeters for Evapotranspiration and Environmental Measurements, ed. Allen, R. G., Howell, T. A., Pruitt, W. O., Walter, I. A. and Jensen, M. E.. New York: American Society of Civil Engineers, pp. 237–45.
Duursma, R., Barton, C., Lin, Y. et al. (2014). The peaked response of transpiration rate to vapour pressure deficit in field conditions can be explained by the temperature optimum of photosynthesis. Agricultural and Forest Meteorology, 189, 2–10.CrossRefGoogle Scholar
Eagleman, J. R. (1976). The Visualization of Climate. Lexington, MA: Lexington Books.
Eichinger, W. E., Parlange, M. B. and Strickler, H. (1996). On the concept of equilibrium evaporation and the value of the Priestley-Taylor coefficient. Water Resources Research, 32, 161–4.CrossRefGoogle Scholar
Eugster, W., McFadden, J. P and Chapin, F. S. (1997). A comparative approach to regional variation in surface fluxes using mobile eddy correlation towers. Boundary-Layer Meteorology, 85, 293–307.CrossRefGoogle Scholar
Eugster, W., Rouse, W. R., Pielke, R. A. et al. (2000). Land-atmosphere energy exchange in Arctic tundra and boreal forest: available data and feedbacks to climate, Global Change Biology, 6, 84–115.Google Scholar
Ezzahar, J. and Chehbouni, A. (2009). The use of scintillometry for validating aggregation schemes over heterogeneous grids. Agricultural and Forest Meteorology, 149, 2098–109.CrossRefGoogle Scholar
Farley, K. A., Jobbagy, E. B. and Jackson, R. B. (2005). Effects of afforestation on water yield: a global synthesis with implications for policy. Global Change Biology, 11, 1565–76.CrossRefGoogle Scholar
Farquhar, G. (1978). Feed forward responses of stomata to humidity. Functional Plant Biology, 5, 787–800.Google Scholar
Farquhar, G. D. and Sharkey, T. D. (1982). Stomatal conductance and photosynthesis. Annual Reviews in Plant Physiology, 33, 317–45.CrossRefGoogle Scholar
Farquhar, G. D. and vonCaemmerer, S. (1982). Modelling of photosynthetic response to environmental conditions. In Physiological Plant Ecology II: Water Relations and Carbon Assimilation, ed. Nobel, P. S., Osmond, C. B. and Ziegler, H.. New York: Springer, pp. 549–87.CrossRef
Fernandes, R., Korolevych, V. and Wang, S. (2007). Trends in land evapotranspiration over Canada for the period 1960–2000 based on in situ climate observations and a land surface model. Journal of Hydrometeorology, 8, 1016–30.CrossRefGoogle Scholar
Fisher, J. B., Baldocchi, D. D, Misson, L., Dawson, T. E. and Goldstein, A. H. (2007). What the towers don't see at night: nocturnal sap flow in trees and shrubs at two AmeriFlux sites in California. Tree Physiology, 27, 597–610.CrossRefGoogle Scholar
Fisher, J. B., Malhi, Y., Bonal, D. et al. (2009). The land–atmosphere water flux in the tropics. Global Change Biology, 15, 2694–714.CrossRefGoogle Scholar
Fisher, J. B., Tu, K. P. and Baldocchi, D. D. (2008). Global estimates of the land-atmosphere water flux based on monthly AVHRR and ISLSCP-II data, validated at 16 FLUXNET sites. Remote Sensing of Environment, 112, 901–19.CrossRefGoogle Scholar
Fisher, J. B., Whittaker, R. J. and Malhi, Y. (2011). ET come home: potential evapotranspiration in geographical ecology. Global Ecology and Biogeography, 20, 1–18.CrossRefGoogle Scholar
Flint, A. L. and Childs, S. W. (1991). Use of the Priestley–Taylor evaporation equation for soil water limited conditions in a small forest clear cut. Agricultural and Forest Meteorology, 56, 247–60.CrossRefGoogle Scholar
Foken, T. (2008). Micrometeorology. Heidelberg: Springer.
Fritschen, L. J. and Fritschen, C. L. (2005). Bowen ratio energy balance method. In Micrometeorology in Agricultural Systems, ed. Hatfield, J. L. and Baker, J. M.. Madison, WI: American Society of Agronomy, Crop Science Society of America, Soil Science Society of America, pp. 397–405.
Fyfe, J. C., Gillett, N. P. and Zwiers, F. W. (2013). Overestimated global warming over the past 20 years. Nature Climate Change, 3, 767–9.CrossRefGoogle Scholar
Gallardo, M., Jackson, L. E., Schulbach, K. et al. (1996). Production and water use in lettuce under variable water supply. Irrigation Science, 16, 125–37.CrossRefGoogle Scholar
Garratt, J. R. and Francey, R. J. (1978). Bulk characteristics of heat transfer in the unstable, baroclinic atmospheric boundary layer. Boundary Layer Meteorology, 15, 399–421.CrossRefGoogle Scholar
Gentine, P., Entekhabi, D., Chehbouni, A., Boulet, G. and Duchemin, B. (2007). Analysis of evaporative fraction diurnal behaviour. Agricultural and Forest Meteorology, 143, 13–29.CrossRefGoogle Scholar
Golubev, V. S., Lawrimore, J. H., Groisman, P. Y. et al. (2001). Evaporation changes over the contiguous United States and the former USSR: a reassessment. Geophysical Research Letters, 28, 2665–8.CrossRefGoogle Scholar
Goss, M. J. and Ehlers, W. (2009). The role of lysimeters in the development of our understanding of soil water and nutrient dynamics in ecosystems. Soil Use Management, 25, 213–23.CrossRefGoogle Scholar
Granier, A. (1985). A new method of sap flow measurement in tree stems. Annales des Sciences Forestieres, 42, 193–200.CrossRefGoogle Scholar
Grant, R. F., Black, T. A., Gaumont-Guay, D. et al. (2006). Net ecosystem productivity of boreal aspen forests under drought and climate change: Mathematical modelling with Ecosys. Agricultural and Forest Meteorology, 140, 152–70.CrossRefGoogle Scholar
Grant, R. F., Wall, G. W., Kimball, B. A. et al. (1999). Crop water relations under different CO2 and irrigation: testing of Ecosys with the free air CO2 enrichment experiment (FACE). Agricultural and Forest Meteorology, 95, 27–51.CrossRefGoogle Scholar
Hamon, W. R. (1961). Estimating potential evapotranspiration. Journal of the Hydraulics Division, 87, 107–20.Google Scholar
Hargreaves, G. H. (1975). Moisture availability and crop production. Transactions of the American Society of Agricultural Engineers, 18, 980–4.CrossRefGoogle Scholar
Harrold, L. L. and Dreibelbis, F. R. (1967). Evaluation of Agriculture Hydrology by Monolith Lysimeters (1956–1962). U.S. Department of Agriculture Technical Bulletin 1367.
Hartogensis, O. K., Watts, C. J., Rodriguez, J. C. and DeBruin, A. R. (2003). Derivation of an effective height for scintillometers: La Poza experiment in northwest Mexico. Journal of Hydrometeorology, 4, 915–28.2.0.CO;2>CrossRefGoogle Scholar
Hasler, N. and Avissar, R. (2007). What controls evapotranspiration in the Amazon Basin? Journal of Hydrometeorology, 8, 380–95.Google Scholar
Hayhoe, K., Wake, C. P., Huntington, T. G. et al. (2007). Past and future changes in climate and hydrological indicators in the U.S. Northeast. Climate Dynamics, 28, 381–407.CrossRefGoogle Scholar
Healy, R. W. and Scanlon, B. R. (2010). Estimating Groundwater Recharge. Cambridge: Cambridge University Press.CrossRef
Held, I. M. and Soden, B. J. (2006). Robust responses of the hydrological cycle to global warming. Journal of Climate, 19, 5686–99.CrossRefGoogle Scholar
Herbst, M., Rosier, P. T. W., McNeil, D. D., Harding, R. J. and Gowing, D. J. (2008). Seasonal variability of interception evaporation from the canopy of a mixed deciduous forest. Agricultural and Forest Meteorology, 148, 1655–67.CrossRefGoogle Scholar
Hilker, T., Hall, F. G., Coops, N. C. et al. (2013). Remote sensing of transpiration and heat fluxes using multi-angle observations. Remote Sensing of Environment, 137, 31–42.CrossRefGoogle Scholar
Hill, R. J. (1997). Algorithms for obtaining atmospheric surface-layer fluxes from scintillation measurements. Journal of Atmospheric and Oceanic Technology, 14, 456–67.2.0.CO;2>CrossRefGoogle Scholar
Holland, S., Heitman, J. L., Howard, A. et al. (2013). Micro-Bowen ratio system for measuring evapotranspiration in a vineyard inter-row. Agricultural and Forest Meteorology, 177, 93–100.CrossRefGoogle Scholar
Hollinger, D. Y., Goltz, S. M., Davidson, E. A. et al. (1999). Seasonal patterns and environmental control of carbon dioxide and water vapor exchange in an ecotonal boreal forest. Global Change Biology, 5, 891–902.CrossRefGoogle Scholar
Holmes, J. W. (1984). Measuring evapotranspiration by hydrological methods. Agricultural Water Management, 8, 29–40.CrossRefGoogle Scholar
Humphreys, E. R., Black, T. A., Ethier, G. J. et al. (2003). Annual and seasonal variability of sensible and latent heat fluxes above a coastal Douglas-fir forest, British Columbia, Canada. Agricultural and Forest Meteorology, 115, 109–25.CrossRefGoogle Scholar
Huntington, T. G. (2010). Climate warming-induced intensification of the hydrologic cycle: an assessment of the published record and potential impacts on agriculture. Advances in Agronomy, 109, 1–52.CrossRefGoogle Scholar
Iroume, A. and Huber, A. (2002). Comparison of interception losses in a broadleaved native forest and a Pseudotsuga menziesii (Douglas fir) plantation in the Andes Mountains of southern Chile. Hydrological Processes, 16, 2347–61.CrossRefGoogle Scholar
Ito, A. and Oikawa, T. (2000). A model analysis of the relationship between climate perturbations and carbon budget anomalies in global terrestrial ecosystems: 1970 to 1997. Climate Research, 15, 161–83.CrossRefGoogle Scholar
Ittner, E. (1968). Der Tagesgang der Geschwindigkeit des Transpirationsstromes im Stam einer 75-jaehrigen Fichte. Oecologia Plantarum, 3, 177–83.Google Scholar
Jackson, R. D., Kimball, B. A., Reginato, R. J. and Nakayama, F. S. (1973). Diurnal soil water evaporation: time-depth-flux patterns. Soil Science Society of America Journal, 37, 505–9.CrossRefGoogle Scholar
Jarvis, P. G. (1976). The interpretation of the variations in leaf water potential and stomatal conductance found in canopies in the field. Philosophical Transactions of the Royal Society London Series B, 273, 593–610.CrossRefGoogle Scholar
Jasechko, S., Sharp, Z. D., Gibson, J. J., et al. (2013). Terrestrial water fluxes dominated by transpiration. Nature, 496, 347–50.CrossRefGoogle Scholar
Jassal, R. S., Black, T. A., Arevalo, C. et al. (2013). Carbon sequestration and water use of a young hybrid poplar plantation in north-central Alberta. Biomass and Bioenergy, 56, 323–33.CrossRefGoogle Scholar
Jassal, R. S., Black, T. A., Spittlehouse, D., Christian, B. and Nesic, Z. (2009). Evapotranspiration and water use efficiency in Pacific Northwest Douglas-fir stands of different ages following clearcut harvesting. Agricultural and Forest Meteorology, 149, 1168–78.CrossRefGoogle Scholar
Jassal, R. S., Novak, M. D. and Black, T. A. (2003). Effect of surface layer thickness on simultaneous transport of heat and water in a bare soil and its implications for land surface schemes. Atmosphere-Ocean, 41, 259–72.CrossRefGoogle Scholar
Jensen, M. E., Burman, R. D. and Allen, R. G. (1990). Evapotranspiration and Irrigation Water Requirements: ASCE Manuals and Reports on Engineering Practice, 70. New York: The Society.
Jones, H. G. (1992). Plants and Microclimate: A Quantitative Approach to Environmental Plant Physiology. New York: Cambridge University Press.
Jones, J. A., Creed, I. F., Hatcher, K. L. et al. (2012). Ecosystem processes and human influences regulate stream flow response to climate change at long-term ecological research sites. Bioscience, 62, 390–404.CrossRefGoogle Scholar
Jung, M., Reichstein, M., Ciais, P. et al. (2010). Recent decline in the global land evapotranspiration trend due to limited moisture supply. Nature, 467, 951–4.CrossRefGoogle Scholar
Kaminski, K. P., Kørup, K., Nielsen, K. L. et al. (2014). Gas-exchange, water use efficiency and yield responses of elite potato (Solanum tuberosum L.) cultivars to changes in atmospheric carbon dioxide concentration, temperature and relative humidity. Agricultural and Forest Meteorology, 187, 36–45.CrossRefGoogle Scholar
Katul, G. G., Oren, R., Manzoni, S., Higgins, C. and Parlange, M. B. (2012). Evapotranspiration: a process driving mass transport and energy exchange in the soil-plant-atmosphere-climate system. Reviews of Geophysics, 50, doi:10.1029/2011RG000366.CrossRefGoogle Scholar
Keenan, T.F., Hollinger, D.Y., Bohrer, G. et al. (2013). Increase in forest water-use efficiency as atmospheric carbon dioxide concentrations rise. Nature, 499, doi:10.1038/nature1229.CrossRefGoogle Scholar
Kelliher, F. M., Leuning, R., Raupach, M. R. and Schulze, E. D. (1995). Maximum conductances for evaporation from global vegetation types. Agricultural and Forest Meteorology, 73, 1–16.CrossRefGoogle Scholar
Kelliher, F. M., Whitehead, D., McAneney, K. J. and Judd, M. J. (1990). Partitioning evapotranspiration into tree and understorey components in two young Pinus radiata stands. Agricultural and Forest Meteorology, 50, 211–27.CrossRefGoogle Scholar
Kharrufa, N. S. (1985). Simplified equation for evapotranspiration in arid regions. Sonderheft, 5, 39–47.Google Scholar
Kim, D., Oren, R., Oishi, A. C. et al. (2014). Sensitivity of stand transpiration to wind velocity in a mixed broadleaved deciduous forest. Agricultural and Forest Meteorology, 187, 62–71.CrossRefGoogle Scholar
Kiptala, J. K., Mohamed, Y., Mul, M. L. and Van der Zaag, P. (2013). Mapping evapotranspiration trends using MODIS and SEBAL model in a data scarce and heterogeneous landscape in Eastern Africa. Water Resources Research, 49, 8495–510.CrossRefGoogle Scholar
Kilinc, M., Beringer, J., Hutley, L. B., Tapper, N. J. and McGuire, D. A. (2013). Carbon and water exchange of the world's tallest angiosperm forest. Agricultural and Forest Meteorology, 182–183, 215–24.CrossRefGoogle Scholar
Kljun, N., Black, T. A., Griffis, T. J. et al. (2006). Response of net ecosystem productivity of three boreal forest stands to drought. Ecosystems, 9, 1128–44.CrossRefGoogle Scholar
Kool, D., Agam, N., Lazarovitch, N. et al. (2014). A review of approaches for evapotranspiration partitioning. Agricultural and Forest Meteorology, 184, 56–70.CrossRefGoogle Scholar
Kruijt, B., Witte, J. P. M., Jacobs, C. M. J. and Kroon, T. (2008). Effects of rising atmospheric CO2 on evapotranspiration and soil moisture: a practical approach for the Netherlands. Journal of Hydrology, 349, 257–67.CrossRefGoogle Scholar
Kumagai, T., Tateishi, M., Miyazawa, Y. et al. (2014). Estimation of annual forest evapotranspiration from a coniferous plantation watershed in Japan (1): water use components in Japanese cedar stands. Journal of Hydrology, 508, 66–76.CrossRefGoogle Scholar
Kume, T., Takizawa, H., Yoshifuji, N. et al. (2007). Impact of soil drought on sap flow and water status of evergreen trees in a tropical monsoon forest in northern Thailand. Forest Ecology and Management, 238, 220–30.CrossRefGoogle Scholar
Kustas, W. and Anderson, M. (2009). Advances in thermal infrared remote sensing for land surface modeling. Agricultural and Forest Meteorology, 149, 2071–81.CrossRefGoogle Scholar
Labat, D., Godderis, Y., Probst, J. L. and Guyot, J. L. (2004) Evidence for global runoff increase related to climate warming. Advances in Water Resources, 27, 631–42.CrossRefGoogle Scholar
Lagouarde, J. P., Bonnefond, J. M, Kerr, Y. H., McAneney, K. J. and Irvine, M. (2002). Integrated sensible heat flux measurements of a two-surface composite landscape using scintillometry. Boundary Layer Meteorology, 105, 5–35.CrossRefGoogle Scholar
Law, B. E., Falge, E., Gu, L. et al. (2002). Environmental controls over carbon dioxide and water vapor exchange of terrestrial vegetation. Agricultural and Forest Meteorology, 113, 97–120.CrossRefGoogle Scholar
Law, B. E., Williams, M., Anthoni, P., Baldocchi, D. D. and Unsworth, M. H. (2000). Measuring and modeling seasonal variation of carbon dioxide and water vapor exchange of a Pinus ponderosa forest subject to soil water deficit. Global Change Biology, 6, 613–30.CrossRefGoogle Scholar
Lawrence, D. M., Oleson, K. W., Flanner, M. G. et al. (2011). Parameterization improvements and functional and structural advances in Version 4 of the Community Land Model. Journal of Advances in Modeling Earth Systems 3, doi:03010.01029/02011MS000045.Google Scholar
Lawrence, D. M., Thornton, P. E., Oleson, K. W. and Bonan, G. B. (2007). The partitioning of evapotranspiration into transpiration, soil evaporation, and canopy evaporation in a GCM: impacts on land-atmosphere interaction. Journal of Hydrometeorology, 8, 862–80.CrossRefGoogle Scholar
Maitre, D. C. Le, Scott, D. F. and Colvin, C. (1999). A review of information on interactions between vegetation and groundwater. Water SA, 25, 137–52.Google Scholar
Lee, D., Kim, J., Lee, K. S. and Kim, S. (2010). Partitioning of catchment water budget and its implications for ecosystem carbon exchange. Biogeosciences, 7, 1903–14.CrossRefGoogle Scholar
Leuning, R. (1995). A critical appraisal of a combined stomatal photosynthesis model for C-3 plants. Plant, Cell and Environment, 18, 339–55.CrossRefGoogle Scholar
Leuning, R., Zhang, Y. Q., Rajaud, A., Cleugh, H. and Tu, K. (2008). A simple surface conductance model to estimate regional evaporation using MODIS leaf area index and the Penman–Monteith equation. Water Resources Research, 44, doi:10.1029/2007WR006562.CrossRefGoogle Scholar
Liepert, B. G., Feichter, L., Lohmann, U. and Roeckner, E. (2004). Can aerosols spin down the water cycle in a warmer and moister world? Geophysical Research Letters, 31, doi:10.1029/2003GL019060.CrossRefGoogle Scholar
Lhomme, J. P. and Guilioni, L. (2006). Comments on some articles about the complementary relationship. Journal of Hydrology, 323, 1–3.CrossRefGoogle Scholar
Li, Z. L., Tang, R. L., Wan, Z. M. et al. (2009). A review of current methodologies for regional evapotranspiration estimation from remotely sensed data. Sensors, 9, 3801–53.CrossRefGoogle Scholar
Lim, W. H., Roderick, M. L., Hobbins, M. T., Wong, S. C. and Farquhar, G. D. (2013). The energy balance of a US Class A evaporation pan. Agricultural and Forest Meteorology, 182, 314–31.CrossRefGoogle Scholar
Linacre, E. T. (1977). A simple formula for estimating evaporation rates in various climates, using temperature data alone. Agricultural and Forest Meteorology, 18, 409–24.CrossRefGoogle Scholar
Llorens, P., Poch, R., Latron, J. and Gallart, F. (1997). Rainfall interception by a Pinus sylvestris forest patch overgrown in a Mediterranean mountainous abandoned area. Journal of Hydrology, 199, 331–45.CrossRefGoogle Scholar
Long, D., Longuevergne, L. and Scanlon, B. R. (2014). Uncertainty in evapotranspiration from land surface modeling, remote sensing, and GRACE satellites. Water Resources Research, 50, 1131–51.CrossRefGoogle Scholar
Lott, R. B. and Hunt, R. J. (2001). Estimating evapotranspiration in natural and constructed wetlands. Wetlands, 21, 614–28.CrossRefGoogle Scholar
Lu, X. and Zhuang, Q. (2010). Evaluating evapotranspiration and water-use efficiency of terrestrial ecosystems in the conterminous United States using MODIS and AmeriFlux data. Remote Sensing of Environment, 114, 1924–39.CrossRefGoogle Scholar
Lüdeke, M. K. B., Badeck, F.-W., Otto, R. D. et al. (1994). The Frankfurt Biosphere Model: a global process oriented model of seasonal and long-term CO2 exchange between terrestrial ecosystems and the atmosphere, I, Model description and illustrative results for cold deciduous and boreal forests. Climate Research, 4, 143–66.CrossRefGoogle Scholar
Matheny, A., Bohrer, G., Stoy, P. et al. (2014). Characterizing the diurnal patterns of errors in the prediction of evapotranspiration by several land-surface models: an NACP analysis. Journal of Geophysical Research- Biogeosciences, 119, 1458–73.CrossRefGoogle Scholar
Margolis, H. A. and Ryan, M. G. (1997). A physiological basis for biosphere-atmosphere interactions in the boreal forest: an overview. Tree Physiology, 17, 491–9.CrossRefGoogle Scholar
Marshall, D. C. (1958). Measurement of sap flow in conifers by heat transport. Plant Physiology, 33, 385–96.CrossRefGoogle Scholar
McAneney, K. J., Green, A. E. and Astill, M. S. (1995). Large aperture scintillometry: the homogeneous case. Agricultural and Forest Meteorology, 76, 149–62.CrossRefGoogle Scholar
McCumber, M. C. and Pielke, R. A. (1981). Simulation of the effects of surface fluxes of heat and moisture in a Mesoscale numerical model: 1. Soil layer. Journal of Geophysical Research, 86, 9929–38.CrossRefGoogle Scholar
McNaughton, K. G. and Black, T. A. (1973). A study of evapotranspiration from a Douglas-fir forest using the energy balance approach. Water Resources Research, 9, 1579–90.CrossRefGoogle Scholar
McNaughton, K. G. and Jarvis, P. G. (1983). Predicting effects of vegetation changes on transpiration and evaporation. In Water Deficits and Plant Growth, ed. Kozlowski, T. T.. New York: Academic Press, pp. 1–47.CrossRef
McNaughton, K. G. and Spriggs, T. W. (1986). A mixed-layer model for regional evaporation. Boundary-Layer Meteorology, 34, 243–62.CrossRefGoogle Scholar
McNaughton, K. G. and Spriggs, T. W. (1989). An evaluation of the Priestley and Taylor equation and the complementary relationship using results from a mixed-layer model of the convective boundary layer. In Estimation of Areal Evapotranspiration, ed. Black, T. A., Spittlehouse, D. L., Novak, M. D., and Price, D. T.. Vancouver, BC: IAHS Publications, pp. 89–104.
Meijninger, W. M. L., Beyrich, F., Ludi, A., Kohsiek, W. and DeBruin, H. A. R. (2006). Scintillometer-based turbulent fluxes of sensible and latent heat over a heterogeneous land surface – a contribution to LITFASS-2003. Boundary-Layer Meteorology, 121, 89–110.CrossRefGoogle Scholar
Meijninger, W. M. L., Hartogensis, O. K., Kohsiek, W. et al. (2002). Determination of area-averaged sensible heat fluxes with a large aperture scintillometer over a heterogeneous surface: Flevoland field experiment. Boundary Layer Meteorology, 105, 37–62.CrossRefGoogle Scholar
Melillo, M., Kicklighter, D. W., McGuire, A. D., Peterjohn, W. T. and Newkirk, K. (1995). Global change and its effects on soil organic carbon stocks. In The Dahlem Conference, ed. Eddy, J. A. and Oeschger, H.. Chichester, UK: John Wiley and Sons, pp. 176–225.
Menzel, A., Sparks, T. H., Estrella, N., et al. (2006). European phenological response to climate change matches the warming pattern. Global Change Biology, 12, 1969–76.CrossRefGoogle 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, 1014–17.CrossRefGoogle Scholar
Mizutani, K., Yamanoi, K., Ikeda, T. and Watanabe, T. (1997). Applicability of the eddy correlation method to measure sensible heat transfer to forest under rainfall conditions. Agricultural and Forest Meteorology, 86, 193–203.CrossRefGoogle Scholar
Moffat, A. M., Papale, D., Reichstein, M. et al. (2007). Comprehensive comparison of gap filling techniques for eddy covariance net carbon fluxes. Agricultural and Forest Meteorology, 147, 209–32.CrossRefGoogle Scholar
Monin, A. S. and Obukhov, A. N. (1954). Basic laws of turbulent mixing in the ground layer of the atmosphere (in Russian). Tr. Akad. Nauk SSSR Geophiz. Inst., 24, 163–87.Google Scholar
Monteith, J. L. (1981). Principles of Environmental Physics. London: Edward Arnold.
Monteith, J. L. and Unsworth, M. H. (2008). Principles of Environmental Physics,. Amsterdam: Elsevier.
Morillas, L., Leuning, R., Villagarcia, L. et al. (2014). Improving evapotranspiration estimates in Mediterranean drylands: the role of soil evaporation. Water Resources Research, 49, 6572–86.CrossRefGoogle Scholar
Morton, F. I. (1965). Potential evaporation and river basin evaporation. ASCE Journal of Hydraulics Division, 102, 275–91.Google Scholar
Morton, F. I. (1983). Operational estimates of areal evapotranspiration and their significance to the science and practice of hydrology. Journal of Hydrology, 66, 1–76.CrossRefGoogle Scholar
Mu, Q., Heinsch, F. A., Zhao, M. and Running, S. W. (2007). Development of a global evapotranspiration algorithm based on MODIS and global meteorology data. Remote Sensing of Environment, 111, 519–36.CrossRefGoogle Scholar
Myneni, R. B., Keeling, C. D., Tucker, C. J., Asrar, G. and Nemani, R. R. (1997). Increased plant growth in the northern high latitudes from 1981 to 1991. Nature, 386, 698–702.CrossRefGoogle Scholar
Nadezhdina, N., Cermák, J. and Nadezhdin, V. (1998). Heat field deformation method for sap flow measurements. In Proceedings of the 4th International Workshop on Measuring Sap Flow in Intact Plants, ed. Cermák, J. and Nadezhdina, N.. Brno, Czech Republic: Publishing House of Mendel University, pp. 72–92.
Nepstad, D. C., de Carvalho, C. R., Davidson, E. A. et al. (1994). The role of deep roots in the hydrological and carbon cycles of Amazonian forests and pastures. Nature, 372, 666–9.CrossRefGoogle Scholar
Nichols, J., Eichinger, W., Cooper, D. I. et al. (2004). Comparison of Evaporation Estimation Methods for a Riparian Area. IIHR Technical Report No 436. Iowa City, IA: Hydroscience and Engineering, University of Iowa.
Niyogi, D., Alapaty, K., Raman, S. and Chen, F. (2009). Development and evaluation of a coupled photosynthesis-based gas exchange evapotranspiration model (GEM) for mesoscale weather forecasting applications. Journal of Applied Meteorology and Climatology, 48, 349–68.CrossRefGoogle Scholar
O'Grady, A. P., Eamus, D. and Hutley, L. B. (1999). Transpiration increases during the dry season: Patterns of tree water use in eucalypt open-forests of northern Australia. Tree Physiology, 19, 591–7.CrossRefGoogle Scholar
Ozdogan, M., Woodcock, C. E., Salvucci, G. D. and Demir, H. (2006). Changes in summer irrigated crop area and water use in southeastern Turkey from 1993–2002: implications for currentand future water resources. Water Resources Management, 20, 467–88.CrossRefGoogle Scholar
Papale, D. and Valentini, R. (2003). A new assessment of European forests carbon exchanges by eddy fluxes and artificial neural network spatialization. Global Change Biology, 9, 525–35.CrossRefGoogle Scholar
Pauliukonis, N. and Schneider, R. (2001). Temporal patterns in evapotranspiration from lysimeters with three common wetland plant species in the eastern United States. Aquatic Botany, 71, 35–46.CrossRefGoogle Scholar
Pawar, S. J. and Firake, N. N. (2003). Effect of irrigation levels and micro-irrigation methods on yield of cabbage. Journal of Maharashtra Agricultural University, 28, 116–7.Google Scholar
Pearson, R. G., Phillips, S. J., Loranty, M. M. et al. (2013). Shifts in Arctic vegetation and associated feedbacks under climate change. Nature Climate Change, 3, 673–7.CrossRefGoogle Scholar
Penman, H. L. (1948). Natural evaporation from open water, bare soil and grass. Proceedings of the Royal Society London A, 193, 120–45.CrossRefGoogle Scholar
Peschke, G., Scholz, J. and Seidler, C. (1991). Field investigations of moisture and temperature fluxes at atmosphere, soil and vegetation interfaces. In Hydrological Interactions between Atmosphere, Soil and Vegetation, ed. Kienitz, G., Milly, P. C. D., Genuchten, M. Th. Van, Rosbjerg, D. and Shuttleworth, W. J.. Wallingford, UK: IAHS Press, pp. 433–96.
Peterson, A. T., Golubev, V. S. and Groisman, P. Y. (1995). Evaporation losing its strength. Nature, 377, 687–8.CrossRefGoogle Scholar
Piao, S. L., Friedlingstein, P., Ciais, P. et al. (2007). Changes in climate and land use have a larger direct impact than rising CO2 on global river runoff trends. Proceedings of the National Academy of Sciences, 104, 15242–7.CrossRefGoogle Scholar
Potter, N. J. and Zhang, L. (2009). Interannual variability of catchment water balance in Australia. Journal of Hydrology, 369, 120–9.CrossRefGoogle Scholar
Postel, S. L., Daily, G. C. and Ehrlich, P. R. (1996). Human appropriation of renewable fresh water. Science, 271, 785–8.CrossRefGoogle Scholar
Price, D. T. and Black, T. A. (1990). Effects of short-term variation in weather on diurnal canopy CO2 flux and evapotranspiration of a juvenile Douglas-fir stand. Agricultural and Forest Meteorology, 50, 139–58.CrossRefGoogle Scholar
Priestley, C. H. B. and Taylor, R. J. (1972). On the assessment of surface heat flux and evaporation using large-scale parameters. Monthly Weather Review, 100, 81–92.2.3.CO;2>CrossRefGoogle Scholar
Pruitt, W. O. and Angus, D. E. (1960). Large weighing lysimeter for measuring evapotranspiration. Transactions ASAE, 3, 13–18.CrossRefGoogle Scholar
Pypker, T. G., Bond, B. J., Link, T. E., Marks, D. and Unsworth, M. H. (2005). The importance of canopy structure in controlling the interception loss of rainfall: examples from a young and an old-growth Douglas-fir forest. Agricultural and Forest Meteorology, 130, 113–29.CrossRefGoogle Scholar
Qian, T., Dai, A. and Trenberth, K. E. (2004). Hydroclimatic trends in the Mississippi River Basin from 1948 to 2004. Journal of Climate, 20, 4599–614.CrossRefGoogle Scholar
Randel, D. L., Vonder Haar, T. H., Ringerud, M. A. et al. (1996). A new global water vapor dataset. Bulletin of the American Meteorological Society, 77, 1233–46.2.0.CO;2>CrossRefGoogle Scholar
Raupach, M. R. (2001). Combination theory and equilibrium evaporation. Quarterly Journal of the Royal Meteorological Society, 127, 1149–81.CrossRefGoogle Scholar
Raupach, M. R., Haverd, V. and Briggs, P. R. (2014). Sensitivities of the Australian terrestrial water and carbon balances to climate change and variability. Agricultural and Forest Meteorology, 182, 277–91.Google Scholar
Reichardt, K. (1985). Processos de Transferência no Sistema Solo-Planta-Atmosfera, 4ª edição. Campinas: Fundação Cargill.
Reichstein, M., Ciais, P., Papale, P. et al. (2007). Reduction of ecosystem productivity and respiration during the European summer 2003 climate anomaly: a joint flux tower, remote sensing and modelling analysis. Global Change Biology, 13, 634–51.CrossRefGoogle Scholar
Reichstein, M., Falge, E., Baldocchi, D. et al. (2005). On the separation of net ecosystem exchange into assimilation andecosystem respiration: review and improved algorithm. Global Change Biology, 11, 1424–39.CrossRefGoogle Scholar
Richardson, A. D., Keenan, T. F., Migliavacca, M. et al. (2013). Climate change, phenology, and phenological control of vegetation feedbacks to the climate system. Agricultural and Forest Meteorology, 169, 156–73.CrossRefGoogle Scholar
Rinke, A., Melsheimer, C., Dethloff, K. and Heygster, G. (2009). Arctic total water vapor: Comparison of regional climate simulations with observations, and simulated decadal trends. Journal of Hydrometeorology, 10, 113–29.CrossRefGoogle Scholar
Rijtema, P. E. and De Vries, W. (1994) Differences in precipitation excess and nitrogen leaching from agricultural lands and forest plantations. Biomass and Bioenergy, 6, 103–13.CrossRefGoogle Scholar
Robock, A., Konstantin, Y. V., Srinivasan, J. K. et al. (2000). The global soil moisture data bank. Bulletin of the American Meteorological Society, 81, 1281–99.2.3.CO;2>CrossRefGoogle Scholar
Roderick, M. L. and Farquhar, G. D. (2002). The cause of decreased pan evaporation over the past 50 years. Science, 298, 1410–11.Google Scholar
Roderick, M. L., Rotstayn, L. D., Farquhar, G. D. and Hobbins, M. T. (2007). On the attribution of changing pan evaporation. Geophysical Research Letters, 34, doi:10.1029/2007gl031166.CrossRefGoogle Scholar
Romanenko, V. A. (1961). Computation of the Autumn Soil Moisture Using a Universal Relationship for a Large Area. Proceedings of the Ukrainian Hydrometeorological Research Institute, No. 3. Kiev: Ukrainian Hydrometeorological Research Institute.
Running, S. W. and Hunt, E. R. (1993). Generalization of a forest ecosystem process model for other biomes, BIOME-BGC, and an application for global scale models. In Scaling Physiological Processes: Leaf to Globe, ed. Ehleringer, J. and Field, C.. New York: Academic Press, pp. 141–58.CrossRef
Running, S. W., Nemani, R. R., Peterson, D. L. et al. (1989). Mapping regional forest evapotranspiration and photosynthesis by coupling satellite data with ecosystem simulation. Ecology, 70, 1090–101.CrossRefGoogle Scholar
Ryu, Y., Baldocchi, D. D., Ma, S. and Hehn, T. (2008). Interannual variability of evapotranspiration and energy exchange over an annual grassland in California. Journal of Geophysical Research, 113, 1–16.CrossRefGoogle Scholar
Saito, T., Matsuda, H., Momatsu, M. et al. (2014). Forest canopy interception loss is greater than evaporation at the wet canopy in Japanese cypress (Hinoki) and cedar (Sugi) plantations. Journal of Hydrology, 507, 287–99.CrossRefGoogle Scholar
Saleska, S. R., Didan, K., Huete, A. R. and da Rocha, H. R. (2007). Amazon forests green-up during 2005 drought. Science, 318, 612–5.CrossRefGoogle Scholar
Scanlon, B. R., Keese, K. E., Flint, A. L. et al. (2006). Global synthesis of groundwater recharge in semiarid and arid regions. Hydrological Processes, 20, 3335–70.CrossRefGoogle Scholar
Schroter, D., Cramer, W., Leemans, R. et al. (2005). Ecosystem service supply and vulnerability to global change in Europe. Science, 310, 1333–7.CrossRefGoogle Scholar
Scott, R. L., Huxman, T. E., Barron-Gafford, G. A. et al. (2014). When vegetation change alters ecosystem water availability. Global Change Biology, 20, 2198–210.CrossRefGoogle Scholar
Schafer, K. V. R. (2011). Canopy stomatal conductance following drought, disturbance, and death in an upland oak/pine forest of the New Jersey Pine Barrens, USA. Frontiers in Plant Science, 2, doi: 10.3389/fpls.2011.00015.CrossRefGoogle Scholar
Schuepp, P. H., Leclerc, M. Y., MacPherson, J. I. and Desjardins, R. L. (1990). Footprint predictions of scalar fluxes from analytical solutions of the diffusion equation. Boundary-Layer Meteorology, 50, 355–73.CrossRefGoogle Scholar
Schlesinger, W. H. and Jasechko, S. (2014) Transpiration in the global water cycle. Agricultural and Forest Meteorology, 189, 115–17.CrossRefGoogle Scholar
Schmidt, W. (1917). Der Masssenaustausch bei der ungeordneten Stromung in freier Luft und seine Folgen (in Russian). Sitzber. Kais. Akad. Wissen. Wien, 126, 757–804.Google Scholar
Schulze, E. D., Kelliher, F. M., Korner, C., Lloyd, J. and Leuning, R. (1994). Relationships among maximum stomatal conductance, ecosystem surface conductance, carbon assimilation rate, and plant nitrogen nutrition- a global ecological scaling exercise. Annual Review of Ecology and Systematics, 25, 629–60.CrossRefGoogle 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, 883–90.2.0.CO;2>CrossRefGoogle 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, 125–61.CrossRefGoogle Scholar
Sheffield, J. and Wood, E. F. (2008). Global trends and variability in soil moisture and drought characteristics, 1950–2000, from observation-driven simulations of the terrestrial hydrologic cycle. Journal of Climate, 21, 432–58.CrossRefGoogle Scholar
Shukla, J. and Mintz, Y. (1982). Influence of land-surface evapotranspiration on the Earth's climate. Science, 215, 1498–501.CrossRefGoogle Scholar
Shuttleworth, W. J. and Calder, I. R. (1979). Has the Priestley–Taylor equation any relevance to forest evaporation? Journal of Applied Meteorology, 18, 639–46.Google Scholar
Shuttleworth, W. J., Gurney, R. J., Hsu, A. Y. and Ormsby, J. P. (1989). FIFE: the variation in energy partition at surface flux sites. IAHS, 186, 67–74.Google Scholar
Singh, R. and Irmak, A. (2009). Estimation of crop coefficients using satellite remote sensing. Journal of Irrigation and Drainage Engineering, ASCE, 135, 597–608.CrossRefGoogle Scholar
Silberstein, R. P., Vertessy, R. A., Morris, J. and Feikema, P. M. (1999). Modelling the effects of soil moisture and solute conditions on long-term tree growth and water use: a case study from the Shepparton irrigation area, Australia. Agricultural Water Management, 39, 283–315.CrossRefGoogle Scholar
Slatyer, R. O. and McIlroy, I. C. (1961). Practical Micrometeorology. Melbourne: CSIRO.
Smith, N. V., Saatchi, S. S. and Randerson, J. T. (2004). Trends in high northern latitude soil freeze and thaw cycles from 1988 to 2002. Journal of Geophysical Research, 109, doi:10.1029/2003JD004472.CrossRefGoogle Scholar
Smith, C. A., Haigh, J. D. and Toumi, R. (2001). Radiative forcing due to trends in stratospheric water vapour. Geophysical Research Letters, 28, 179–82.CrossRefGoogle Scholar
Soden, B. J., Wetherald, R. T., Stenchikov, G. L. and Robock, A. (2002). Global cooling after the eruption of Mount Pinatubo: a test of climate feedback by water vapor. Science, 296, 727–30.CrossRefGoogle Scholar
Solomon, S., Rosenlof, K. H., Portmann, R. W. et al. (2010). Contributions of stratospheric water vapor to decadal changes in the rate of global warming. Science, 327, 1219–23.CrossRefGoogle Scholar
Sommer, R., Sa, T. D. D., Vielhauer, K. et al. (2002). Transpiration and canopy conductance of secondary vegetation in the eastern Amazon. Agricultural and Forest Meteorology, 112, 103–21.CrossRefGoogle Scholar
Spittlehouse, D. L. (2003). Water availability, climate change and the growth of Douglas-fir in the Georgia Basin. Canadian Water Resources Journal, 28, 673–88.CrossRefGoogle Scholar
Spittlehouse, D. L. (1989). Estimating evapotranspiration from land surfaces in British Columbia. In Estimating Areal Evapotranspiration, ed. Black, T. A., Spittlehouse, D. L., Novak, M. D. and Price, D. T.). International Association of Hydrological Sciences Publication No. 177. Wallingford, UK: International Association of Hydrological Sciences, pp. 245–56.
Stephens, G. L. and Ellis, T. D. (2008). Controls of global-mean precipitation increases in global warming GCM experiments. Journal of Climate, 21, 6141–55.CrossRefGoogle Scholar
Stewart, J. B. (1988). Modelling surface conductance of pine forest. Agricultural and Forest Meteorology, 43, 19–35.CrossRefGoogle Scholar
Sumner, D. M. and Jacobs, J. M. (2005). Utility of Penman–Monteith, Priestley–Taylor, reference evapotranspiration, and pan evaporation methods to estimate pasture evapotranspiration. Journal of Hydrology, 308, 81–104.CrossRefGoogle Scholar
Sun, G., Alstad, K., Chen, J. et al. (2011a). A general predictive model for estimating monthly ecosystem evapotranspiration. Ecohydrology, 4, 244–55.Google Scholar
Sun, G., Caldwell, P., Noormets, A. et al. (2011b). Upscaling key ecosystem functions across the conterminous United States by a water-centric ecosystem model. Journal of Geophysical Research, 116, doi:10.1029/2010JG001573.CrossRefGoogle Scholar
Sun, S., Meng, P., Zhang, J. et al. (2014). Partitioning oak woodland evapotranspiration in the rocky mountainous area of North China was disturbed by foreign vapor, as estimated based on non-steady-state 18O isotopic composition. Agricultural and Forest Meteorology, 184, 36–47.CrossRefGoogle Scholar
Suyker, A. E. and Verma, S. B. (2008). Interannual water vapor and energy exchange in an irrigated maize-based agroecosystem. Agricultural and Forest Meteorology, 148, 417–27.CrossRefGoogle Scholar
Swanson, R. H. (1994). Significant historical developments in thermal methods for measuring sap flow in trees. Agricultural and Forest Meteorology, 72,113–32.CrossRefGoogle Scholar
Tan, C. S. and Black, T. A. (1976). Factors affecting the canopy resistance of a Douglas-fir forest. Boundary Layer Meteorology, 10, 475–88.CrossRefGoogle Scholar
Tan, C. S., Black, T. A. and Nnyamah, J. U. (1978). A simple vapor diffusion model applied to a thinned Douglas-fir stand. Ecology, 59, 1221–9.CrossRefGoogle Scholar
Tanner, C. B. 1960. Energy balance approach to evapotranspiration from crops. Soil Science Society of America Proceedings, 24, 1–9.Google Scholar
Tasumi, M. and Allen, R. G. (2007). Satellite-based ET mapping to assess variation in ET with timing of crop development. Agricultural Water Management, 88, 54–62.CrossRefGoogle Scholar
Teuling, A. J., Hirschi, M., Ohmura, A. et al. (2009). A regional perspective on trends in continental evaporation, Geophysical Research Letters, 36, doi:10.1029/2008GL036584.CrossRefGoogle Scholar
Thornthwaite, C. W. (1948). An approach toward a rational classification of climate. Geographical Review, 38, 55–94.CrossRefGoogle Scholar
Trenberth, K. E. and Dai, A. (2007). Effects of Mount Pinatubo volcanic eruption on the hydrological cycle as an analog of geoengineering. Geophysical Research Letters, 34, doi:10.1029/2007GL030524.CrossRefGoogle Scholar
Trenberth, K. E., Fasullo, J. T. and Kiehl, J. (2009). Earth's global energy budget. Bulletin of the American Meteorological Society, 90, 311–23.CrossRefGoogle Scholar
Trenberth, K. E., Jones, P. D., Ambenje, P. et al. (2007). Observations: surface and atmospheric climate change. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. Solomon, S., Qin, D., Manning, M. et al. Cambridge: Cambridge University Press, pp. 235–336.
Tuzet, A., Perrier, A. and Leuning, R. (2003). A coupled model of stomatal conductance, photosynthesis and transpiration. Plant, Cell and Environment, 26 1097–116.CrossRefGoogle Scholar
van Bavel, C. H. M. (1966). Potential evaporation: the combination concept and its experimental verification. Water Resources Research, 2, 455–67.CrossRefGoogle Scholar
van Groenigen, K. J., Osenberg, C. W. and Hungate, B. A. (2011). Increased soil emissions, of potent greenhouse gases under increased atmospheric CO2 . Nature, 475, 214–6.CrossRefGoogle Scholar
Verma, S. B. (1989). Aerodynamic resistances to transfers of heat, mass and momentum. In Estimation of Areal Evapotranspiration, ed. Black, T. A., Spittlehouse, D. L., Novak, M. D. and Price, D. T.. Vancouver, BC: IAHS Publication, pp. 13–20.
Verseghy, D. L. (1991). CLASS-A Canadian land surface scheme for GCMS. 1. Soil model. International Journal of Climatology, 11, 111–33.Google Scholar
Vertessy, R. A., Hatton, T. J., Benyon, R. G. and Dawes, W. R. (1996). Long-term growth and water balance predictions for a mountain ash (Eucalyptus regnans) forest catchment subject to clear-felling and regeneration. Tree Physiology, 16, 221–32.CrossRefGoogle Scholar
Vesala, T., Ahonen, T., Hari, P., Krissinel, E. and Shokhirev, N. (1996). Analysis of stomatal CO2 uptake by a three-dimensional cylindrically symmetric model. New Phytologist, 132, 235–45.CrossRefGoogle Scholar
Vico, G., Manzoni, S., Palmroth, S., Weih, M. and Katul, G. (2014). A perspective on optimal leaf stomatal conductance under CO2 and light co-limitations. Agricultural and Forest Meteorology, 182, 191–9.Google Scholar
Viewig, G. H. and Ziegler, H. (1960). Thermoelektrische registrierung der geswindigkeit des transpirationsstromes. Berichte der Deutschen Botanischen Gesellschaft, 73, 1–226.Google Scholar
Vourlitis, G. L., Priante, N., Hayashi, M. S. et al. (2002). Seasonal variations in the evapotranspiration of a transitional tropical forest of Mato Grosso, Brazil. Water Resources Research, 38, doi:10.1029/2000WR000122.CrossRefGoogle Scholar
Wang, K. and Dickinson, R. E. (2012). A review of global terrestrial evapotranspiration: observation, modeling, climatology, and climatic variability. Reviews of Geophysics, 50, doi:10.1029/2011RG000373.CrossRefGoogle Scholar
Wang, K. and Liang, S. (2008). An improved method for estimating global evapotranspiration based on satellite estimation of surface net radiation, vegetation index, temperature, and soil moisture. Journal of Hydrometeorology, 9, 712–27.CrossRefGoogle Scholar
Wang, K., Wild, M., Dickinson, D. E. and Liang, S. (2010). Evidence for decadal variation in global terrestrial evapotranspiration between 1982 and 2002. Part II: Results. Journal of Geophysical Research, 50, doi:10.1029/2010JD013847.CrossRef
Wang, L. X., Caylor, K. K., Villegas, J. C. et al. (2010). Partitioning evapotranspiration across gradients of woody plant cover: Assessment of a stable isotope technique. Geophysical Research Letters, 37, doi:10.1029/2010GL043228.CrossRefGoogle Scholar
Ward, H. C., Evans, J. G., Hartogensis, O. K. et al. (2013). A critical revision of the estimation of the latent heat flux from two-wavelength scintillometry. Quarterly Journal of the Royal Meteorological Society, 139, 1912–22.CrossRefGoogle Scholar
Watson, F. G. R., Vertessy, R. A. and Grayson, R. B. (1999). Largescale modelling of forest hydrological processes and their long-term effect on water yield. Hydrological Processes, 13, 689–700.3.0.CO;2-D>CrossRefGoogle Scholar
Wattenbach, M., Zebisch, M. and Hattermann, F. (2007). Hydrological impact assessment of afforestation and change in tree-species composition – a regional case study for the Federal State of Brandenburg (Germany). Journal of Hydrology, 346, 1–17.CrossRefGoogle Scholar
Wentz, F. J., Ricciardulli, L., Hilburn, K. and Mears, C. (2007). How much more rain will global warming bring? Science, 317, 233–5.Google Scholar
Wesely, M. L. and Hart, R. L. (1985). Variability of short term eddy-correlation estimates of mass exchange. In The Forest–Atmosphere Interaction, ed. Hutchinson, B. A. and Hicks, B. B.. Dordrecht: D. Reidel, pp. 591–612.CrossRef
Wild, M. (2009). Global dimming and brightening: a review. Journal of Geophysical Research, 114, doi:10.1029/2008JD011470.CrossRefGoogle Scholar
Wild, M., Gilgen, H., Roesch, A. et al. (2005). From dimming to brightening: decadal changes in solar radiation at earth's surface. Science, 308, 847–50.CrossRefGoogle Scholar
Williams, C. A., Reichstein, R., 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, doi:10.1029/2011WR011586.CrossRefGoogle Scholar
Wilson, K. B. and Baldocchi, D. D. (2000). Seasonal and interannual variability of energy fluxes over a broadleaved temperate deciduous forest in North America. Agricultural and Forest Meteorology, 100, 1–18.CrossRefGoogle Scholar
Wilson, K. B., Goldstein, A., Falge, I., Aubinet, M. and Baldocchi, D. (2002). Energy balance closure at FLUXNET sites. Agricultural and Forest Meteorology, 113, 223–43.CrossRefGoogle Scholar
Wilson, R. G. and Rouse, W. R. (1972). Moisture and temperature limits of the equilibrium evapotranspiration model. Journal of Applied Meteorology 11, 436–42.2.0.CO;2>CrossRefGoogle Scholar
WMO (World Meteorological Organization) (2008). Guide to Meteorological Instruments and Methods of Observation, WMO No. 8,. Geneva: WMO.
Woeikoff, E. (1887). Die Klimate der Erde (in Russian). Jena: H. Costenoble, pp. 396–422.
Wohlfahrt, G., Irschick, C., Thalinger, B. et al. (2010). Insights from independent evapotranspiration estimates for closing the energy balance: a grassland case study. Vadose Zone, 9, doi:10.2136/vzj2009.0158.CrossRefGoogle Scholar
Wollny, E. (1877). Der Einfluss der Pflannzendecke und Beschattung auf die physicalischen Eigenschaften und die Fruchtbarkeit des Bodens (in Russian). Berlin: Verlag.
Woodward, F. I. (1990). Global change: translating plant ecophysiological responses to ecosystems. Trends in Ecology and Evolution, 5, 308–11.CrossRefGoogle Scholar
Wu, A., Black, T. A., Verseghy, D. L., Novak, M. D. and Bailey, W. G. (2000). Testing the α and β methods of estimating evaporation from bare and vegetated surfaces in CLASS. Atmosphere-Ocean, 38, 15–35.CrossRefGoogle Scholar
Wullschleger, S. D., Gunderson, C. A., Hanson, P. J., Wilson, K. B. and Norby, R. J. (2002). Sensitivity of stomatal and canopy conductance to elevated CO2 concentration—interacting variables and perspectives of scale. New Phytologist, 153, 485–96.CrossRefGoogle Scholar
Yang, G., Pu, R., Zhao, C. and Xue, X. (2014). Estimating high spatiotemporal resolution evapotranspiration over a winter wheat field using an IKONOS image based complementary relationship and Lysimeter observations. Agricultural Water Management, 133, 34–43.CrossRefGoogle 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, 701–8.CrossRefGoogle Scholar
Zhang, Y., Leuning, R., Hutley, L. et al. (2010). Using long-term water balances to parameterize surface conductances and calculate evaporation at spatial resolution. Water Resources Research, 46, doi:10.1029/2009WR008716.CrossRefGoogle Scholar
Zveryaev, I. and Allan, R. P. (2010). Summertime precipitation variability over Europe and its links to atmospheric dynamics and evaporation. Journal of Geophysical Research, 115, doi:10/1029/2008JD011213.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×