Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-wpx84 Total loading time: 0 Render date: 2024-08-09T13:08:54.151Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  08 February 2019

Gordon Bonan
Affiliation:
National Center for Atmospheric Research, Boulder, Colorado
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019

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

Abbott, M. B., Bathurst, J. C., Cunge, J. A., O’Connell, P. E., and Rasmussen, J. (1986). An introduction to the European Hydrological System – Systeme Hydrologique Europeen, “SHE”, 2: Structure of a physically-based, distributed modelling system. Journal of Hydrology, 87, 6177.Google Scholar
Aber, J. D., and Melillo, J. M. (1980). Litter decomposition: Measuring relative contributions of organic matter and nitrogen to forest soils. Canadian Journal of Botany, 58, 416421.CrossRefGoogle Scholar
Aber, J. D., and Melillo, J. M. (1982). Nitrogen immobilization in decaying hardwood leaf litter as a function of initial nitrogen and lignin content. Canadian Journal of Botany, 60, 22632269.CrossRefGoogle Scholar
Aber, J. D., Melillo, J. M., and Federer, C. A. (1982). Predicting the effects of rotation length, harvest intensity, and fertilization on fiber yield from northern hardwood forests in New England. Forest Science, 28, 3145.Google Scholar
Aber, J. D., Melillo, J. M., and McClaugherty, C. A. (1990). Predicting long-term patterns of mass loss, nitrogen dynamics, and soil organic matter formation from initial fine litter chemistry in temperate forest ecosystems. Canadian Journal of Botany, 68, 22012208.Google Scholar
Adair, E. C., Parton, W. J., Del Grosso, S. J., et al. (2008). Simple three-pool model accurately describes patterns of long-term litter decomposition in diverse climates. Global Change Biology, 14, 26362660.CrossRefGoogle Scholar
Ågren, G. I., and Bosatta, E. (1987). Theoretical analysis of the long-term dynamics of carbon and nitrogen in soils. Ecology, 68, 11811189.CrossRefGoogle Scholar
Ågren, G. I., and Bosatta, E. (1996). Theoretical Ecosystem Ecology: Understanding Element Cycles. Cambridge: Cambridge University Press.Google Scholar
Ainsworth, E. A., and Rogers, A. (2007). The response of photosynthesis and stomatal conductance to rising [CO2]: Mechanisms and environmental interactions. Plant, Cell and Environment, 30258270.CrossRefGoogle ScholarPubMed
Ainsworth, E. A., Yendrek, C. R., Sitch, S., Collins, W. J., and Emberson, L. D. (2012). The effects of tropospheric ozone on net primary productivity and implications for climate change. Annual Review of Plant Biology, 63, 637661.Google Scholar
Akagi, S. K., Yokelson, R. J., Wiedinmyer, C., et al. (2011). Emission factors for open and domestic biomass burning for use in atmospheric models. Atmospheric Chemistry and Physics, 11, 40394072.Google Scholar
Ali, A. A., Xu, C., Rogers, A., et al. (2016). A global scale mechanistic model of photosynthetic capacity (LUNA V1.0). Geoscientific Model Development, 9, 587606.CrossRefGoogle Scholar
Amenu, G. G., and Kumar, P. (2008). A model for hydraulic redistribution incorporating coupled soil–root moisture transport. Hydrology and Earth System Sciences, 12, 5574.Google Scholar
Amthor, J. S., Goulden, M. L., Munger, J. W., and Wofsy, S. C. (1994). Testing a mechanistic model of forest-canopy mass and energy exchange using eddy correlation: Carbon dioxide and ozone uptake by a mixed oak–maple stand. Australian Journal of Plant Physiology, 21, 623651.Google Scholar
Anderson, E. A. (1976). A Point Energy and Mass Balance Model of a Snow Cover, NOAA Technical Report NWS 19. Silver Spring, MD: Office of Hydrology, National Weather Service.Google Scholar
Anderson, J. L., Balaji, V., Broccoli, A. J., et al. (2004). The new GFDL global atmosphere and land model AM2–LM2: Evaluation with prescribed SST simulations. Journal of Climate, 17, 46414673.Google Scholar
Anderson, M. C., Kustas, W. P., Norman, J. M., et al. (2011). Mapping daily evapotranspiration at field to continental scales using geostationary and polar orbiting satellite imagery. Hydrology and Earth System Sciences, 15, 223239.Google Scholar
Anderson, M. C., Norman, J. M., Mecikalski, J. R., Otkin, J. A., and Kustas, W. P. (2007). A climatological study of evapotranspiration and moisture stress across the continental United States based on thermal remote sensing: 1. Model formulation. Journal of Geophysical Research, 112, D10117, doi:10.1029/2006JD007506.Google Scholar
Andreae, M. O., and Merlet, P. (2001). Emission of trace gases and aerosols from biomass burning. Global Biogeochemical Cycles, 15, 955966.Google Scholar
Andrén, O., and Paustian, K. (1987). Barley straw decomposition in the field: A comparison of models. Ecology, 68, 11901200.Google Scholar
Anten, N. P. R., Miyazawa, K., Hikosaka, K., Nagashima, H., and Hirose, T. (1998a). Leaf nitrogen distribution in relation to leaf age and photon flux density in dominant and subordinate plants in dense stands of a dicotyledonous herb. Oecologia, 113, 314324.Google Scholar
Anten, N. P. R., Schieving, F., and Werger, M. J. A. (1995). Patterns of light and nitrogen distribution in relation to whole canopy carbon gain in C3 and C4 mono- and dicotyledonous species. Oecologia, 101, 504513.CrossRefGoogle ScholarPubMed
Anten, N. P. R., Werger, M. J. A., and Medina, E. (1998b). Nitrogen distribution and leaf area indices in relation to photosynthetic nitrogen use efficiency in savanna grasses. Plant Ecology, 138, 6375.CrossRefGoogle Scholar
Arneth, A., Lloyd, J., Šantrůčková, H., et al. (2002). Response of central Siberian Scots pine to soil water deficit and long-term trends in atmospheric CO2 concentration. Global Biogeochemical Cycles, 16, 1005, doi:10.1029/2000GB001374.CrossRefGoogle Scholar
Arneth, A., Mercado, L., Kattge, J., and Booth, B. B. B. (2012). Future challenges of representing land-processes in studies on land-atmosphere interactions. Biogeosciences, 9, 35873599.Google Scholar
Arneth, A., Monson, R. K., Schurgers, G., Niinemets, Ü., and Palmer, P. I. (2008). Why are estimates of global terrestrial isoprene emissions so similar (and why is this not so for monoterpenes)? Atmospheric Chemistry and Physics, 8, 46054620.Google Scholar
Arneth, A., Niinemets, Ü., Pressley, S., et al. (2007). Process-based estimates of terrestrial ecosystem isoprene emissions: Incorporating the effects of a direct CO2-isoprene interaction. Atmospheric Chemistry and Physics, 7, 3153.CrossRefGoogle Scholar
Arneth, A., Schurgers, G., Lathiere, J., et al. (2011). Global terrestrial isoprene emission models: Sensitivity to variability in climate and vegetation. Atmospheric Chemistry and Physics, 11, 80378052.Google Scholar
Arora, V. K., and Boer, G. J. (2005a). A parameterization of leaf phenology for the terrestrial ecosystem component of climate models. Global Change Biology, 11, 3959.Google Scholar
Arora, V. K., and Boer, G. J. (2005b). Fire as an interactive component of dynamic vegetation models. Journal of Geophysical Research, 110, G02008, doi:10.1029/2005JG000042.Google Scholar
Arora, V. K., and Boer, G. J. (2010). Uncertainties in the 20th century carbon budget associated with land use change. Global Change Biology, 16, 33273348.CrossRefGoogle Scholar
Arora, V. K., Boer, G. J., Christian, J. R., et al. (2009). The effect of terrestrial photosynthesis down regulation on the twentieth-century carbon budget simulated with the CCCma Earth system model. Journal of Climate, 22, 60666088.Google Scholar
Ashworth, K., Chung, S. H., Griffin, R. J., et al. (2015). FORest Canopy Atmosphere Transfer (FORCAsT) 1.0: A 1-D model of biosphere–atmosphere chemical exchange. Geoscientific Model Development, 8, 37653784.Google Scholar
Asner, G. P., Scurlock, J. M. O., and Hicke, J. A. (2003). Global synthesis of leaf area index observations: Implications for ecological and remote sensing studies. Global Ecology and Biogeography, 12, 191205.Google Scholar
Assouline, S. (2013). Infiltration into soils: Conceptual approaches and solutions. Water Resources Research, 49, 17551772, doi:10.1002/wrcr.20155.Google Scholar
Aston, A. R. (1985). Heat storage in a young eucalypt forest. Agricultural and Forest Meteorology, 35, 281297.Google Scholar
Atkin, O. K. (2016). Corrigendum. New Phytologist, 211, 1142.Google Scholar
Atkin, O. K., Atkinson, L. J., Fisher, R. A., et al. (2008). Using temperature-dependent changes in leaf scaling relationships to quantitatively account for thermal acclimation of respiration in a coupled global climate–vegetation model. Global Change Biology, 14, 27092726.CrossRefGoogle Scholar
Atkin, O. K., Bloomfield, K. J., Reich, P. B., et al. (2015). Global variability in leaf respiration in relation to climate, plant functional types and leaf traits. New Phytologist, 206, 614636.Google Scholar
Aubinet, M., Grelle, A., Ibrom, A., et al. (2000). Estimates of the annual net carbon and water exchange of forests: The EUROFLUX methodology. Advances in Ecological Research, 30, 113175.Google Scholar
Aubinet, M., Vesala, T., and Papale, D. (2012). Eddy Covariance: A Practical Guide to Measurement and Data Analysis. Dordrecht: Springer.CrossRefGoogle Scholar
Bailey, R. L., and Dell, T. R. (1973). Quantifying diameter distributions with the Weibull function. Forest Science, 19, 97104.Google Scholar
Baker, I. T., Prihodko, L., Denning, A. S., et al. (2008). Seasonal drought stress in the Amazon: Reconciling models and observations. Journal of Geophysical Research, 113, G00B01, doi:10.1029/2007JG000644.Google Scholar
Baker, I. T., Sellers, P. J., Denning, A. S., et al. (2017). Closing the scale gap between land surface parameterizations and GCMs with a new scheme, SiB3-Bins. Journal of Advances in Modeling Earth Systems, 9, 691711, doi:10.1002/2016MS000764.Google Scholar
Baldocchi, D. (1988). A multi-layer model for estimating sulfur dioxide deposition to a deciduous oak forest canopy. Atmospheric Environment, 22, 869884.Google Scholar
Baldocchi, D. (1989). Turbulent transfer in a deciduous forest. Tree Physiology, 5, 357377.Google Scholar
Baldocchi, D. (1992). A Lagrangian random-walk model for simulating water vapor, CO2 and sensible heat flux densities and scalar profiles over and within a soybean canopy. Boundary-Layer Meteorology, 61, 113144.Google Scholar
Baldocchi, D. (1994). An analytical solution for coupled leaf photosynthesis and stomatal conductance models. Tree Physiology, 14, 10691079.Google Scholar
Baldocchi, D. (2003). Assessing the eddy covariance technique for evaluating carbon dioxide exchange rates of ecosystems: Past, present and future. Global Change Biology, 9, 479492.CrossRefGoogle Scholar
Baldocchi, D. D., and Bowling, D. R. (2003). Modelling the discrimination of 13CO2 above and within a temperate broad-leaved forest canopy on hourly to seasonal time scales. Plant, Cell and Environment, 26, 231244.CrossRefGoogle Scholar
Baldocchi, D. D., and Harley, P. C. (1995). Scaling carbon dioxide and water vapour exchange from leaf to canopy in a deciduous forest. II. Model testing and application. Plant, Cell and Environment, 18, 11571173.Google Scholar
Baldocchi, D. D., and Meyers, T. P. (1988). A spectral and lag-correlation analysis of turbulence in a deciduous forest canopy. Boundary-Layer Meteorology, 45, 3158.CrossRefGoogle Scholar
Baldocchi, D. D., and Meyers, T. P. (1998). On using eco-physiological, micrometeorological and biogeochemical theory to evaluate carbon dioxide, water vapor and trace gas fluxes over vegetation: A perspective. Agricultural and Forest Meteorology, 90, 125.Google Scholar
Baldocchi, D. D., Hicks, B. B., and Camara, P. (1987). A canopy stomatal resistance model for gaseous deposition to vegetated surfaces. Atmospheric Environment, 21, 91101.Google Scholar
Baldocchi, D. D., Hicks, B. B., and Meyers, T. P. (1988). Measuring biosphere–atmosphere exchanges of biologically related gases with micrometeorological methods. Ecology, 69, 13311340.CrossRefGoogle Scholar
Baldocchi, D. D., Hutchison, B. A., Matt, D. R., and McMillen, R. T. (1985). Canopy radiative transfer models for spherical and known leaf inclination angle distributions: A test in an oak–hickory forest. Journal of Applied Ecology, 22, 539555.Google Scholar
Baldocchi, D. D., and Wilson, K. B. (2001). Modeling CO2 and water vapor exchange of a temperate broadleaved forest across hourly to decadal time scales. Ecological Modelling, 142, 155184.Google Scholar
Baldocchi, D. D., Wilson, K. B., and Gu, L. (2002). How the environment, canopy structure and canopy physiological functioning influence carbon, water and energy fluxes of a temperate broad-leaved deciduous forest – an assessment with the biophysical model CANOAK. Tree Physiology, 22, 10651077.Google Scholar
Ball, J. T., Woodrow, I. E., and Berry, J. A. (1987). A model predicting stomatal conductance and its contribution to the control of photosynthesis under different environmental conditions. In Progress in Photosynthesis Research, vol. 4, ed. Biggins, J.. Dordrecht: Martinus Nijhoff, pp. 221224.Google Scholar
Ball, M. C., Cowan, I. R., and Farquhar, G. D. (1988). Maintenance of leaf temperature and the optimisation of carbon gain in relation to water loss in a tropical mangrove forest. Australian Journal of Plant Physiology, 15, 263276.Google Scholar
Balsamo, G., Albergel, C., Beljaars, A., et al. (2015). ERA-Interim/Land: A global land surface reanalysis data set. Hydrology and Earth System Sciences, 19, 389407.CrossRefGoogle Scholar
Balsamo, G., Viterbo, P., Beljaars, A., et al. (2009). A revised hydrology for the ECMWF model: Verification from field site to terrestrial water storage and impact in the integrated forecast system. Journal of Hydrometeorology, 10, 623643.Google Scholar
Band, L. E., Patterson, P., Nemani, R., and Running, S. W. (1993). Forest ecosystem processes at the watershed scale: Incorporating hillslope hydrology. Agricultural and Forest Meteorology, 63, 93126.Google Scholar
Barman, R., and Jain, A. K. (2016). Comparison of effects of cold-region soil/snow processes and the uncertainties from model forcing data on permafrost physical characteristics. Journal of Advances in Modeling Earth Systems, 8, 453466, doi:10.1002/2015MS000504.Google Scholar
Barnard, D. M., and Bauerle, W. L. (2013). The implications of minimum stomatal conductance on modeling water flux in forest canopies. Journal of Geophysical Research Biogeosciences, 118, 13221333, doi:10.1002/jgrg.20112.Google Scholar
Bauerle, W. L., Daniels, A. B., and Barnard, D. M. (2014). Carbon and water flux responses to physiology by environment interactions: A sensitivity analysis of variation in climate on photosynthetic and stomatal parameters. Climate Dynamics, 42, 25392554.Google Scholar
Bauerle, W. L., Oren, R., Way, D. A., et al. (2012). Photoperiodic regulation of the seasonal pattern of photosynthetic capacity and the implications for carbon cycling. Proceedings of the National Academy of Sciences USA, 109, 86128617.Google Scholar
Belcher, S. E., Finnigan, J. J., and Harman, I. N. (2008). Flows through forest canopies in complex terrain. Ecological Applications, 18, 14361453.CrossRefGoogle ScholarPubMed
Belcher, S. E., Harman, I. N., and Finnigan, J. J. (2012). The wind in the willows: Flows in forest canopies in complex terrain. Annual Review of Fluid Mechanics, 44, 479504.Google Scholar
Beljaars, A. C. M., and Holtslag, A. A. M. (1991). Flux parameterization over land surfaces for atmospheric models. Journal of Applied Meteorology, 30, 327341.Google Scholar
Berg, B., Berg, M. P., Bottner, P., et al. (1993). Litter mass loss rates in pine forests of Europe and eastern United States: Some relationships with climate and litter quality. Biogeochemistry, 20, 127159.Google Scholar
Berg, B., Hannus, K., Popoff, T., and Theander, O. (1982). Changes in organic chemical components of needle litter during decomposition. Long-term decomposition in a Scots pine forest. I. Canadian Journal of Botany, 60, 13101319.Google Scholar
Bernacchi, C. J., Bagley, J. E., Serbin, S. P., et al. (2013). Modelling C3 photosynthesis from the chloroplast to the ecosystem. Plant, Cell and Environment, 36, 16411657.Google Scholar
Bernacchi, C. J., Pimentel, C., and Long, S. P. (2003). In vivo temperature response functions of parameters required to model RuBP-limited photosynthesis. Plant, Cell and Environment, 26, 14191430.Google Scholar
Bernacchi, C. J., Portis, A. R., Nakano, H., von Caemmerer, S., and Long, S. P. (2002). Temperature response of mesophyll conductance. Implications for the determination of rubisco enzyme kinetics and for limitations to photosynthesis in vivo. Plant Physiology, 130, 19921998.Google Scholar
Bernacchi, C. J., Singsaas, E. L., Pimentel, C., Portis, A. R. J., and Long, S. P. (2001). Improved temperature response functions for models of Rubisco-limited photosynthesis. Plant, Cell and Environment, 24, 253259.Google Scholar
Best, M. J., Pryor, M. Clark, D. B., et al. (2011). The Joint UK Land Environment Simulator (JULES), model description – Part 1: Energy and water fluxes. Geoscientific Model Development, 4, 677699.CrossRefGoogle Scholar
Beven, K. (1993). Prophecy, reality and uncertainty in distributed hydrological modelling. Advances in Water Resources, 16, 4151.Google Scholar
Beven, K. (2000). Rainfall–Runoff Modelling: The Primer. Chichester: Wiley.Google Scholar
Beven, K. (2002). Towards a coherent philosophy for modelling the environment. Proceedings of the Royal Society London A, 458, 24652484.Google Scholar
Beven, K. (2006). A manifesto for the equifinality thesis. Journal of Hydrology, 320, 1836.CrossRefGoogle Scholar
Beven, K., and Freer, J. (2001). Equifinality, data assimilation, and uncertainty estimation in mechanistic modelling of complex environmental systems using the GLUE methodology. Journal of Hydrology, 249, 1129.Google Scholar
Beven, K. J., and Kirkby, M. J. (1979). A physically based, variable contributing area model of basin hydrology. Hydrological Sciences Bulletin, 24, 4369.CrossRefGoogle Scholar
Beven, K. J., Lamb, R., Quinn, P. F., Romanowicz, R., and Freer, J. (1995). TOPMODEL. In Computer Models of Watershed Hydrology, ed. Singh, V. P.. Highlands Ranch, CO: Water Resources Publications, pp. 627668.Google Scholar
Blackman, F. F. (1905). Optima and limiting factors. Annals of Botany, 19, 281295.Google Scholar
Blanken, P. D., Black, T. A., Yang, P. C., et al. (1997). Energy balance and canopy conductance of a boreal aspen forest: Partitioning overstory and understory components. Journal of Geophysical Research, 102D, 2891528927.Google Scholar
Bohrer, G., Mourad, H., Laursen, T. A., et al. (2005). Finite element tree crown hydrodynamics model (FETCH) using porous media flow within branching elements: A new representation of tree hydrodynamics. Water Resources Research, 41, W11404, doi:10.1029/2005WR004181.Google Scholar
Bolhàr-Nordenkampf, H. R., and Draxler, G. (1993). Functional leaf anatomy. In Photosynthesis and Production in a Changing Environment: A Field and Laboratory Manual, ed. Hall, D. O., Scurlock, J. M. O., Bolhàr-Nordenkampf, H. R., Leegood, R. C., and Long, S. P.. New York, NY: Chapman and Hall, pp. 91112.Google Scholar
Bonacina, C., Comini, G., Fasano, A., and Primicerio, M. (1973). Numerical solution of phase-change problems. International Journal of Heat and Mass Transfer, 16, 18251832.Google Scholar
Bonan, G. B. (1989). Environmental factors and ecological processes controlling vegetation patterns in boreal forests. Landscape Ecology, 3, 111130.Google Scholar
Bonan, G. B. (1990a). Carbon and nitrogen cycling in North American boreal forests. I. Litter quality and soil thermal effects in interior Alaska. Biogeochemistry, 10, 128.Google Scholar
Bonan, G. B. (1990b). Carbon and nitrogen cycling in North American boreal forests. II. Biogeographic patterns. Canadian Journal of Forest Research, 20, 10771088Google Scholar
Bonan, G. B. (1991). A biophysical surface energy budget analysis of soil temperature in the boreal forests of interior Alaska. Water Resources Research, 27, 767781.Google Scholar
Bonan, G. B. (1993). Physiological controls of the carbon balance of boreal forest ecosystems. Canadian Journal of Forest Research, 23, 14531471.Google Scholar
Bonan, G. B. (1995). Land-atmosphere CO2 exchange simulated by a land surface process model coupled to an atmospheric general circulation model. Journal of Geophysical Research, 100D, 28172831.Google Scholar
Bonan, G. B. (1996). A Land Surface Model (LSM Version 1.0) for Ecological, Hydrological, and Atmospheric Studies: Technical Description and User’s Guide, Technical Note NCAR/TN-417+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Bonan, G. B. (2008). Forests and climate change: Forcings, feedbacks, and the climate benefits of forests. Science, 320, 14441449.Google Scholar
Bonan, G. B. (2014). Connecting mathematical ecosystems, real-world ecosystems, and climate science. New Phytologist, 202, 731733.Google Scholar
Bonan, G. B. (2016). Ecological Climatology: Concepts and Applications, 3rd edn. Cambridge: Cambridge University Press.Google Scholar
Bonan, G. B., Davis, K. J., Baldocchi, D., Fitzjarrald, D., and Neumann, H. (1997). Comparison of the NCAR LSM1 land surface model with BOREAS aspen and jack pine tower fluxes. Journal of Geophysical Research, 102D, 2906529075.Google Scholar
Bonan, G. B., and Doney, S. C. (2018). Climate, ecosystems, and planetary futures: The challenge to predict life in Earth system models. Science, 359, eaam8328, doi:10.1126/science.aam8328.Google Scholar
Bonan, G. B., Hartman, M. D., Parton, W. J., and Wieder, W. R. (2013). Evaluating litter decomposition in earth system models with long-term litterbag experiments: An example using the Community Land Model version 4 (CLM4). Global Change Biology, 19, 957974.Google Scholar
Bonan, G. B., and Hayden, B. P. (1990). Using a forest stand simulation model to examine the ecological and climatic significance of the late-Quaternary pine–spruce pollen zone in eastern Virginia, U.S.A. Quaternary Research, 33, 204218.Google Scholar
Bonan, G. B., Lawrence, P. J., Oleson, K. W., et al. (2011). Improving canopy processes in the Community Land Model version 4 (CLM4) using global flux fields empirically inferred from FLUXNET data. Journal of Geophysical Research, 116, G02014, doi:10.1029/2010JG001593.Google Scholar
Bonan, G. B., and Levis, S. (2010). Quantifying carbon–nitrogen feedbacks in the Community Land Model (CLM4). Geophysical Research Letters, 37, L07401, doi:10.1029/2010GL042430.Google Scholar
Bonan, G. B., Levis, S., Kergoat, L., and Oleson, K. W. (2002). Landscapes as patches of plant functional types: An integrating concept for climate and ecosystem models. Global Biogeochemical Cycles, 16, 1021, doi:10.1029/2000GB001360.Google Scholar
Bonan, G. B., Levis, S., Sitch, S., Vertenstein, M., and Oleson, K. W. (2003). A dynamic global vegetation model for use with climate models: Concepts and description of simulated vegetation dynamics. Global Change Biology, 9, 15431566.Google Scholar
Bonan, G. B., Patton, E. G., Harman, I. N., et al. (2018). Modeling canopy-induced turbulence in the Earth system: A unified parameterization of turbulent exchange within plant canopies and the roughness sublayer (CLM-ml v0). Geoscientific Model Development, 11, 14671496.Google Scholar
Bonan, G. B., Williams, M., Fisher, R. A., and Oleson, K. W. (2014). Modeling stomatal conductance in the earth system: Linking leaf water-use efficiency and water transport along the soil–plant–atmosphere continuum. Geoscientific Model Development, 7, 21932222.Google Scholar
Boone, A., Samuelsson, P., Gollvik, S., et al. (2017). The interactions between soil–biosphere–atmosphere land surface model with a multi-energy balance (ISBA-MEB) option in SURFEXv8 – Part 1: Model description. Geoscientific Model Development, 10, 843872.Google Scholar
Bosatta, E., and Ågren, G. I. (1985). Theoretical analysis of decomposition of heterogeneous substrates. Soil Biology and Biochemistry, 17, 601610.Google Scholar
Bosatta, E., and Ågren, G. I. (1991). Dynamics of carbon and nitrogen in the organic matter of the soil: A generic theory. American Naturalist, 138, 227245.Google Scholar
Bosatta, E., and Ågren, G. I. (2003). Exact solutions to the continuous-quality equation for soil organic matter turnover. Journal of Theoretical Biology, 224, 97105.Google Scholar
Botkin, D. B. (1993). Forest Dynamics: An Ecological Model. Oxford: Oxford University Press.Google Scholar
Botkin, D. B., Janak, J. F., and Wallis, J. R. (1972). Some ecological consequences of a computer model of forest growth. Journal of Ecology, 60, 849872.Google Scholar
Bouten, W., Schaap, M. G., Aerts, J., and Vermetten, A. W. M. (1996). Monitoring and modelling canopy water storage amounts in support of atmospheric deposition studies. Journal of Hydrology, 181, 305321.Google Scholar
Boy, M., Sogachev, A., Lauros, J., et al. (2011). SOSA – a new model to simulate the concentrations of organic vapours and sulphuric acid inside the ABL – Part 1: Model description and initial evaluation. Atmospheric Chemistry and Physics, 11, 4351.Google Scholar
Bras, R. L. (1990). Hydrology: An Introduction to Hydrologic Science. Reading, MA: Addison-Wesley.Google Scholar
Bresler, E., and Dagan, G. (1983). Unsaturated flow in spatially variable fields: 2. Application of water flow models to various fields. Water Resources Research, 19, 421428.Google Scholar
Brooks, R. H., and Corey, A. T. (1964). Hydraulic Properties of Porous Media, Hydrology Papers No. 3. Fort Collins, CO: Colorado State University.Google Scholar
Brooks, R. H., and Corey, A. T. (1966). Properties of porous media affecting fluid flow. Journal of the Irrigation and Drainage Division Proceedings of the American Society of Civil Engineers, 92(IR2), 6188.Google Scholar
Brovkin, V., van Bodegom, P. M., Kleinen, T., et al. (2012). Plant-driven variation in decomposition rates improves projections of global litter stock distribution. Biogeosciences, 9, 565576.Google Scholar
Brutsaert, W. (1982). Evaporation into the Atmosphere: Theory, History, and Applications. Dordrecht: Kluwer.Google Scholar
Brutsaert, W. (2005). Hydrology: An Introduction. Cambridge: Cambridge University Press.Google Scholar
Bryan, A. M., Bertman, S. B., Carroll, M. A., et al. (2012). In-canopy gas-phase chemistry during CABINEX 2009: Sensitivity of a 1-D canopy model to vertical mixing and isoprene chemistry. Atmospheric Chemistry and Physics, 12, 88298849.Google Scholar
Brzostek, E. R., Fisher, J. B., and Phillips, R. P. (2014). Modeling the carbon cost of plant nitrogen acquisition: Mycorrhizal trade-offs and multipath resistance uptake improve predictions of retranslocation. Journal of Geophysical Research: Biogeosciences, 119, 16841697, doi:10.1002/2014JG002660.Google Scholar
Buckley, T. N. (2017). Modeling stomatal conductance. Plant Physiology, 174, 572582.Google Scholar
Buckley, T. N., Cescatti, A., and Farquhar, G. D. (2013). What does optimization theory actually predict about crown profiles of photosynthetic capacity when models incorporate greater realism? Plant, Cell and Environment, 36, 15471563.Google Scholar
Buckley, T. N., Miller, J. M., and Farquhar, G. D. (2002). The mathematics of linked optimisation for water and nitrogen use in a canopy. Silva Fennica, 36(3), 639669.Google Scholar
Buckley, T. N., Sack, L., and Farquhar, G. D. (2017). Optimal plant water economy. Plant, Cell and Environment, 40, 881896.Google Scholar
Buckley, T. N., and Schymanski, S. J. (2014). Stomatal optimisation in relation to atmospheric CO2. New Phytologist, 201, 372377.Google Scholar
Bugmann, H. (2001). A review of forest gap models. Climatic Change, 51, 259305.Google Scholar
Burke, E. J., Chadburn, S. E., and Ekici, A. (2017). A vertical representation of soil carbon in the JULES land surface scheme (vn4.3_permafrost) with a focus on permafrost regions. Geoscientific Model Development, 10, 959975.Google Scholar
Calder, I. R. (1977). A model of transpiration and interception loss from a spruce forest in Plynlimon, central Wales. Journal of Hydrology, 33, 247265.Google Scholar
Campbell, G. S. (1974). A simple method for determining unsaturated conductivity from moisture retention data. Soil Science, 117, 311314.Google Scholar
Campbell, G. S. (1985). Soil Physics with Basic: Transport Models for Soil–Plant Systems. Amsterdam: Elsevier.Google Scholar
Campbell, G. S. (1986). Extinction coefficients for radiation in plant canopies calculated using an ellipsoidal inclination angle distribution. Agricultural and Forest Meteorology, 36, 317321.Google Scholar
Campbell, G. S. (1990). Derivation of an angle density function for canopies with ellipsoidal leaf angle distributions. Agricultural and Forest Meteorology, 49, 173176.Google Scholar
Canadell, J., Jackson, R. B., Ehleringer, J. R., et al. (1996). Maximum rooting depth of vegetation types at the global scale. Oecologia, 108, 583595.Google Scholar
Campbell, G. S., and Norman, J. M. (1998). An Introduction to Environmental Biophysics, 2nd edn. New York, NY: Springer-Verlag.Google Scholar
Carlyle-Moses, D. E., and Gash, J. H. C. (2011). Rainfall interception loss by forest canopies. In Forest Hydrology and Biogeochemistry: Synthesis of Past Research and Future Directions, ed. Levia, D. F., Carlyle-Moses, D., and Tanaka, T.. Dordrecht: Springer, pp. 407423.Google Scholar
Carsel, R. F., and Parrish, R. S. (1988). Developing joint probability distributions of soil water retention characteristics. Water Resources Research, 24, 755769.Google Scholar
Carswell, F. E., Meir, P., Wandelli, E. V., et al. (2000). Photosynthetic capacity in a central Amazonian rain forest. Tree Physiology, 20, 179186.Google Scholar
Celia, M. A., Bouloutas, E. T., and Zarba, R. L. (1990). A general mass-conservative numerical solution for the unsaturated flow equation. Water Resources Research, 26, 14831496.Google Scholar
Cellier, P., and Brunet, Y. (1992). Flux–gradient relationships above tall plant canopies. Agricultural and Forest Meteorology, 58, 93117.Google Scholar
Cernusak, L. A., Barbour, M. M., Arndt, S. K., et al. (2016). Stable isotopes in leaf water of terrestrial plants. Plant, Cell and Environment, 39, 10871102.Google Scholar
Cescatti, A. (1997). Modelling the radiative transfer in discontinuous canopies of asymmetric crowns. I. Model structure and algorithms. Ecological Modelling, 101, 263274.CrossRefGoogle Scholar
Cescatti, A., and Niinemets, Ü. (2004). Sunlight capture: Leaf to landscape. In Photosynthetic Adaptation: Chloroplast to Landscape, ed. Smith, W. K., Vogelmann, T. C., and Critchley, C.. New York, NY: Springer, pp. 4285.Google Scholar
Chamberlain, A. C. (1966). Transport of gases to and from grass and grass-like surfaces. Proceedings of the Royal Society London A, 290, 236265.Google Scholar
Chen, B., Liu, J., Chen, J. M., et al. (2016a). Assessment of foliage clumping effects on evapotranspiration estimates in forested ecosystems. Agricultural and Forest Meteorology, 216, 8292.Google Scholar
Chen, C. P., Zhu, X.-G., and Long, S. P. (2008). The effect of leaf-level spatial variability in photosynthetic capacity on biochemical parameter estimates using the Farquhar model: A theoretical analysis. Plant Physiology, 148, 11391147.Google Scholar
Chen, F., and Dudhia, J. (2001). Coupling an advanced land surface–hydrology model with the Penn State–NCAR MM5 modeling system. Part I: Model implementation and sensitivity. Monthly Weather Review, 129, 569585.Google Scholar
Chen, J., and Kumar, P. (2001). Topographic influence on the seasonal and interannual variation of water and energy balance of basins in North America. Journal of Climate, 14, 19892014.Google Scholar
Chen, J.-L., Reynolds, J. F., Harley, P. C., and Tenhunen, J. D. (1993). Coordination theory of leaf nitrogen distribution in a canopy. Oecologia, 93, 6369.Google Scholar
Chen, J. M., and Black, T. A. (1992). Defining leaf area index for non-flat leaves. Plant, Cell and Environment, 15, 421429.Google Scholar
Chen, J. M., Menges, C. H., and Leblanc, S. G. (2005). Global mapping of foliage clumping index using multi-angular satellite data. Remote Sensing of Environment, 97, 447457.Google Scholar
Chen, J. M., Mo, G., Pisek, J., et al. (2012). Effects of foliage clumping on the estimation of global terrestrial gross primary productivity. Global Biogeochemical Cycles, 26, GB1019, doi:10.1029/2010GB003996.Google Scholar
Chen, Y., Ryder, J., Bastrikov, V., et al. (2016b). Evaluating the performance of land surface model ORCHIDEE-CAN v1.0 on water and energy flux estimation with a single- and multi- layer energy budget scheme. Geoscientific Model Development, 9, 29512972.Google Scholar
Chen, Z. Q., Govindaraju, R. S., and Kavvas, M. L. (1994). Spatial averaging of unsaturated flow equations under infiltration conditions over areally heterogeneous fields: 1. Development of models. Water Resources Research, 30, 523533.Google Scholar
Cherkauer, K. A., and Lettenmaier, D. P. (1999). Hydrologic effects of frozen soils in the upper Mississippi River basin. Journal of Geophysical Research, 104D, 1959919610.CrossRefGoogle Scholar
Choat, B., Jansen, S., Brodribb, T. J., et al. (2012). Global convergence in the vulnerability of forests to drought. Nature, 491, 752755.Google Scholar
Choudhury, B. J., and Monteith, J. L. (1988). A four-layer model for the heat budget of homogeneous land surfaces. Quarterly Journal of the Royal Meteorological Society, 114, 373398.Google Scholar
Christoffersen, B. O., Gloor, M., Fauset, S., et al. (2016). Linking hydraulic traits to tropical forest function in a size-structured and trait-driven model (TFS v.1-Hydro). Geoscientific Model Development, 9, 42274255.Google Scholar
Cionco, R. M. (1965). A mathematical model for air flow in a vegetative canopy. Journal of Applied Meteorology, 4, 517522.Google Scholar
Cionco, R. M. (1978). Analysis of canopy index values for various canopy densities. Boundary-Layer Meteorology, 15, 8193.Google Scholar
Clapp, R. B., and Hornberger, G. M. (1978). Empirical equations for some soil hydraulic properties. Water Resources Research, 14, 601604.Google Scholar
Clapp, R. B., Hornberger, G. M., and Cosby, B. J. (1983). Estimating spatial variability in soil moisture with a simplified dynamic model. Water Resources Research, 19, 739745.Google Scholar
Clark, D. B., Mercado, L. M., Sitch, S., et al. (2011a). The Joint UK Land Environment Simulator (JULES), model description – Part 2: Carbon fluxes and vegetation dynamics. Geoscientific Model Development, 4, 701722.Google Scholar
Clark, M. P., Hendrikx, J., Slater, A. G., et al. (2011b). Representing spatial variability of snow water equivalent in hydrologic and land-surface models: A review. Water Resources Research, 47, W07539, doi:10.1029/2011WR010745.Google Scholar
Clark, M. P., Fan, Y., Lawrence, D. M., et al. (2015a). Improving the representation of hydrologic processes in Earth System Models. Water Resources Research, 51, 59295956, doi:10.1002/2015WR017096.Google Scholar
Clark, M. P., Nijssen, B., Lundquist, J. D., et al. (2015b). A unified approach for process-based hydrologic modeling: 1. Modeling concept. Water Resources Research, 51, 24982514, doi:10.1002/2015WR017198.Google Scholar
Clark, M. P., Nijssen, B., Lundquist, J. D., (2015c). A unified approach for process-based hydrologic modeling: 2. Model implementation and case studies. Water Resources Research, 51, 25152542, doi:10.1002/2015WR017200.Google Scholar
Clarke, R. H., Dyer, A. J., Brook, R. R., Reid, D. G., and Troup, A. J. (1971). The Wangara Experiment: Boundary Layer Data, Division of Meteorological Physics Technical Paper Number 19. Melbourne: Commonwealth Scientific and Industrial Research Organization.Google Scholar
Collatz, G. J., Ball, J. T., Grivet, C., and Berry, J. A. (1991). Physiological and environmental regulation of stomatal conductance, photosynthesis and transpiration: A model that includes a laminar boundary layer. Agricultural and Forest Meteorology, 54, 107136.CrossRefGoogle Scholar
Collatz, G. J., Berry, J. A., Farquhar, G. D., and Pierce, J. (1990). The relationship between the Rubisco reaction mechanism and models of photosynthesis. Plant, Cell and Environment, 13, 219225.Google Scholar
Collatz, G. J., Ribas-Carbo, M., and Berry, J. A. (1992). Coupled photosynthesis–stomatal conductance model for leaves of C4 plants. Australian Journal of Plant Physiology, 19, 519538.Google Scholar
Connor, D. J., Sadras, V. O., and Hall, A. J. (1995). Canopy nitrogen distribution and the photosynthetic performance of sunflower crops during grain filling – a quantitative analysis. Oecologia, 101, 274281.Google Scholar
Cornwell, W. K, Cornelissen, J. H. C., Amatangelo, K., et al. (2008). Plant species traits are the predominant control on litter decomposition rates within biomes worldwide. Ecology Letters, 11, 10651071.Google Scholar
Cosby, B. J., Hornberger, G. M., Clapp, R. B., and Ginn, T. R. (1984). A statistical exploration of the relationships of soil moisture characteristics to the physical properties of soils. Water Resources Research, 20, 682690.Google Scholar
Cowan, I. R. (1968). Mass, heat and momentum exchange between stands of plants and their atmospheric environment. Quarterly Journal of the Royal Meteorological Society, 94, 523544.Google Scholar
Cowan, I. R. (1977). Stomatal behaviour and environment. Advances in Botanical Research, 4, 117228.Google Scholar
Cowan, I. R. (1982). Regulation of water use in relation to carbon gain in higher plants. In Encyclopedia of Plant Physiology, New Series, vol. 12B. Physiological Plant Ecology. II. Water Relations and Carbon Assimilation, ed. Lange, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H.. Berlin: Springer-Verlag, pp. 589613.Google Scholar
Cowan, I. R., and Farquhar, G. D. (1977). Stomatal function in relation to leaf metabolism and environment. In Integration of Activity in the Higher Plant, ed. Jennings, D. H.. Cambridge: Cambridge University Press, pp. 471505.Google Scholar
Cox, P. M. (2001). Description of the “TRIFFID” Dynamic Global Vegetation Model, Technical Note 24. Bracknell: Met Office Hadley Centre.Google Scholar
Cox, P. M., Betts, R. A., Bunton, C. B., et al. (1999). The impact of new land surface physics on the GCM simulation of climate and climate sensitivity. Climate Dynamics, 15, 183203.Google Scholar
Cox, P. M., Betts, R. A., Jones, C. D., Spall, S. A., and Totterdell, I. J. (2000). Acceleration of global warming due to carbon-cycle feedbacks in a coupled climate model. Nature, 408, 184187.Google Scholar
Cox, P. M., Huntingford, C., and Harding, R. J. (1998). A canopy conductance and photosynthesis model for use in a GCM land surface scheme. Journal of Hydrology, 212/213, 7994.Google Scholar
Craig, H., and Gordon, L. I. (1965). Deuterium and oxygen-18 variations in the ocean and the marine atmosphere. In Stable Isotopes in Oceanographic Studies and Paleotemperatures, ed. Tongiorgi, E.. Pisa: Consiglio Nazionale delle Ricerche - Laboratorio di Geologia Nucleare, pp. 9130.Google Scholar
Crank, J., and Nicolson, P. (1947). A practical method for numerical evaluation of solutions of partial differential equations of the heat-conduction type. Mathematical Proceedings of the Cambridge Philosophical Society, 43, 5067.Google Scholar
Crawford, N. H., and Burges, S. J. (2004). History of the Stanford Watershed Model. Water Resources IMPACT, 6(2), 35.Google Scholar
Crawford, N. H., and Linsley, R. K. (1966). Digital Simulation in Hydrology: Stanford Watershed Model IV, Technical Report No. 39, Department of Civil Engineering. Stanford, CA: Stanford University.Google Scholar
Currie, W. S., Harmon, M. E., Burke, I. C., et al. (2010). Cross-biome transplants of plant litter show decomposition models extend to a broader climatic range but lose predictability at the decadal time scale. Global Change Biology, 16, 17441761.Google Scholar
Dagan, G., and Bresler, E. (1983). Unsaturated flow in spatially variable fields: 1. Derivation of models of infiltration and redistribution. Water Resources Research, 19, 413420.Google Scholar
Dai, Y., Dickinson, R. E., and Wang, Y.-P. (2004). A two-big-leaf model for canopy temperature, photosynthesis, and stomatal conductance. Journal of Climate, 17, 22812299.Google Scholar
Damour, G., Simonneau, T., Cochard, H., and Urban, L. (2010). An overview of models of stomatal conductance at the leaf level. Plant, Cell and Environment, 33, 14191438.Google Scholar
Dang, Q. L., Margolis, H. A., and Collatz, G. J. (1998). Parameterization and testing of a coupled photosynthesis–stomatal conductance model for boreal trees. Tree Physiology, 18, 141153.Google Scholar
Dang, Q. L., Margolis, H. A., Coyea, M. R., Sy, M., and Collatz, G. J. (1997a). Regulation of branch-level gas exchange of boreal trees: Roles of shoot water potential and vapor pressure difference. Tree Physiology, 17, 521535.Google Scholar
Dang, Q. L., Margolis, H. A., Sy, M., et al. (1997b). Profiles of photosynthetically active radiation, nitrogen and photosynthetic capacity in the boreal forest: Implications for scaling from leaf to canopy. Journal of Geophysical Research, 102D, 2884528859.Google Scholar
Davidson, E. A. (1991). Fluxes of nitrous oxide and nitric oxide from terrestrial ecosystems. In Microbial Production and Consumption of Greenhouse Gases: Methane, Nitrogen Oxides and Halomethanes, ed. Rogers, J. E. and Whitman, W. B.. Washington, DC: American Society for Microbiology, pp. 219235.Google Scholar
Davidson, E. A., Keller, M., Erickson, H. E., Verchot, L. V., and Veldkamp, E. (2000). Testing a conceptual model of soil emissions of nitrous and nitric oxides. BioScience, 50, 667680.Google Scholar
Deardorff, J. W. (1978). Efficient prediction of ground surface temperature and moisture, with inclusion of a layer of vegetation. Journal of Geophysical Research, 83C, 18891903.Google Scholar
De Kauwe, M. G., Medlyn, B. E., Zaehle, S., et al. (2013). Forest water use and water use efficiency at elevated CO2: A model-data intercomparison at two contrasting temperate forest FACE sites. Global Change Biology, 19, 17591779.Google Scholar
De Kauwe, M. G., Medlyn, B. E., Zaehle, S., (2014). Where does the carbon go? A model–data intercomparison of vegetation carbon allocation and turnover processes at two temperate forest free-air CO2 enrichment sites. New Phytologist, 203, 883899.Google Scholar
De Kauwe, M. G., Kala, J., Lin, Y.-S., et al. (2015). A test of an optimal stomatal conductance scheme within the CABLE land surface model. Geoscientific Model Development, 8, 431452.Google Scholar
Del Grosso, S. J., Parton, W. J., Mosier, A. R., et al. (2000). General model for N2O and N2 gas emissions from soils due to denitrification. Global Biogeochemical Cycles, 14, 10451060.Google Scholar
Del Grosso, S. J., Parton, W. J., Mosier, A. R., (2005a). Modeling soil CO2 emissions from ecosystems. Biogeochemistry, 73, 7191.Google Scholar
Del Grosso, S. J., Mosier, A. R., Parton, W. J., and Ojima, D. S. (2005b). DAYCENT model analysis of past and contemporary soil N2O and net greenhouse gas flux for major crops in the USA. Soil & Tillage Research, 83, 924.Google Scholar
Del Grosso, S. J., Ojima, D. S., Parton, W. J., et al. (2009). Global scale DAYCENT model analysis of greenhouse gas emissions and mitigation strategies for cropped soils. Global and Planetary Change, 67, 4450.Google Scholar
Denmead, O. T., and Bradley, E. F. (1985). Flux–gradient relationships in a forest canopy. In The Forest–Atmosphere Interaction, ed. Hutchinson, B. A. and Hicks, B. B.. Dordrecht: Reidel, pp. 421442.Google Scholar
de Pury, D. G. G., and Farquhar, G. D. (1997). Simple scaling of photosynthesis from leaves to canopies without the errors of big-leaf models. Plant, Cell and Environment, 20, 537557.Google Scholar
De Ridder, K. (2010). Bulk transfer relations for the roughness sublayer. Boundary-Layer Meteorology, 134, 257267.Google Scholar
de Vries, D. A. (1963). Thermal properties of soils. In Physics of Plant Environment, ed. van Wijk, W. R.. Amsterdam: North-Holland, pp. 210235.Google Scholar
Dewar, R., Mauranen, A., Mäkelä, A., et al. (2018). New insights into the covariation of stomatal, mesophyll and hydraulic conductances from optimization models incorporating nonstomatal limitations to photosynthesis. New Phytologist, 217, 571585.Google Scholar
de Wit, C. T. (1965). Photosynthesis of Leaf Canopies, Agricultural Research Reports Number 663. Wageningen: Center for Agricultural Publications and Documentation.Google Scholar
Diaz-Espejo, A., Bernacchi, C. J., Collatz, G. J., and Sharkey, T. D. (2012). Models of photosynthesis. In Terrestrial Photosynthesis in a Changing Environment: A Molecular, Physiological and Ecological Approach, ed. Flexas, J., Loreto, F., and Medrano, H.. Cambridge: Cambridge University Press, pp. 98112.Google Scholar
Dickinson, R. E. (1983). Land surface processes and climate-surface albedos and energy balance. Advances in Geophysics, 25, 305353.Google Scholar
Dickinson, R. E. (1988). The force-restore model for surface temperatures and its generalizations. Journal of Climate, 1, 10861097.Google Scholar
Dickinson, R. E., Jäger, J., Washington, W. M., and Wolski, R. (1981). Boundary Subroutine for the NCAR Global Climate Model, Technical Note NCAR/TN-173+IA. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Dickinson, R. E., Henderson-Sellers, A., and Kennedy, P. J. (1993). Biosphere–Atmosphere Transfer Scheme (BATS) Version 1e as Coupled to the NCAR Community Climate Model, Technical Note NCAR/TN-387+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Dickinson, R. E., Henderson-Sellers, A., Kennedy, P. J., and Wilson, M. F. (1986). Biosphere–Atmosphere Transfer Scheme (BATS) for the NCAR Community Climate Model, Technical Note NCAR/TN-275+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Dickinson, R. E., Berry, J. A., Bonan, G. B., et al. (2002). Nitrogen controls on climate model evapotranspiration. Journal of Climate, 15, 278295.Google Scholar
Dickinson, R. E., Shaikh, M., Bryant, R., and Graumlich, L. (1998). Interactive canopies for a climate model. Journal of Climate, 11, 28232836.Google Scholar
Dingman, S. L. (1994). Physical Hydrology. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Dolman, A. J. (1993). A multiple-source land surface energy balance model for use in general circulation models. Agricultural and Forest Meteorology, 65, 2145.Google Scholar
Dolman, A. J., and Gregory, D. (1992). The parametrization of rainfall interception in GCMs. Quarterly Journal of the Royal Meteorological Society, 118, 455467.Google Scholar
Dolman, A. J., and Wallace, J. S. (1991). Lagrangian and K-theory approaches in modelling evaporation from sparse canopies. Quarterly Journal of the Royal Meteorological Society, 117, 13251340.Google Scholar
Domingues, T. F., Berry, J. A., Martinelli, L. A., Ometto, J. P. H. B., and Ehleringer, J. R. (2005). Parameterization of canopy structure and leaf-level gas exchange for an eastern Amazonian tropical rain forest (Tapajós National Forest, Pará, Brazil). Earth Interactions, 9(17), 123.Google Scholar
Donigian, A. S., Jr., and Imhoff, J. (2010). History and evolution of watershed modeling derived from the Stanford Watershed Model. In Watershed Models, ed. Singh, V. P. and Frevert, D. K.. Boca Raton, FL: CRC Press, pp. 2145.Google Scholar
Dreccer, M. F., Schapendonk, A. H. C. M., van Oijen, M., Pot, C. S., and Rabbinge, R. (2000a). Radiation and nitrogen use at the leaf and canopy level by wheat and oilseed rape during the critical period for grain number definition. Australian Journal of Plant Physiology, 27, 899910.Google Scholar
Dreccer, M. F., van Oijen, M., Schapendonk, A. H. C. M., Pot, C. S., and Rabbinge, R. (2000b). Dynamics of vertical leaf nitrogen distribution in a vegetative wheat canopy: Impact on canopy photosynthesis. Annals of Botany, 86, 821831.Google Scholar
Drewry, D. T., and Albertson, J. D. (2006). Diagnosing model error in canopy-atmosphere exchange using empirical orthogonal function analysis. Water Resources Research, 42, W06421, doi:10.1029/2005WR004496.Google Scholar
Drewry, D. T., Kumar, P., Long, S., et al. (2010). Ecohydrological responses of dense canopies to environmental variability: 1. Interplay between vertical structure and photosynthetic pathway. Journal of Geophysical Research, 115, G04022, doi:10.1029/2010JG001340.Google Scholar
Drouet, J.-L., and Bonhomme, R. (1999). Do variations in local leaf irradiance explain changes to leaf nitrogen within row maize canopies? Annals of Botany, 84, 6169.Google Scholar
Drouet, J.-L., and Bonhomme, R. (2004). Effect of 3D nitrogen, dry mass per area and local irradiance on canopy photosynthesis within leaves of contrasted heterogeneous maize crops. Annals of Botany, 93, 699710.Google Scholar
Du, Z., Weng, E., Xia, J., et al. (2018). Carbon–nitrogen coupling under three schemes of model representation: Traceability analysis. Geoscientific Model Development Discussions, https://doi.org/10.5194/gmd-2018-41.Google Scholar
Duarte, H. F., Raczka, B. M., Ricciuto, D. M., et al. (2017). Evaluating the Community Land Model (CLM 4.5) at a coniferous forest site in northwestern United States using flux and carbon-isotope measurements. Biogeosciences, 14, 43154340.Google Scholar
Ducharne, A., Koster, R. D., Suarez, M. J., Stieglitz, M., and Kumar, P. (2000). A catchment-based approach to modeling land surface processes in a general circulation model: 2. Parameter estimation and model demonstration. Journal of Geophysical Research, 105D, 2482324838.Google Scholar
Ducharne, A., Laval, K., and Polcher, J. (1998). Sensitivity of the hydrological cycle to the parameterization of soil hydrology in a GCM. Climate Dynamics, 14, 307327.Google Scholar
Ducoudré, N. I., Laval, K., and Perrier, A. (1993). SECHIBA, a new set of parameterizations of the hydrologic exchanges at the land–atmosphere interface within the LMD atmospheric general circulation model. Journal of Climate, 6, 248273.Google Scholar
Dümenil, L., and Todini, E. (1992). A rainfall-runoff scheme for use in the Hamburg climate model. In Advances in Theoretical Hydrology: A Tribute to James Dooge, ed. O’Kane, J. P.. Amsterdam: Elsevier, pp. 129157.Google Scholar
Duursma, R. A. (2015). Plantecophys – An R package for analysing and modelling leaf gas exchange data. PLoS ONE, 10(11), e0143346, doi:10.1371/journal.pone.0143346.Google Scholar
Duursma, R. A., Barton, C. V. M., Lin, Y.-S., 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, 189/190, 210.Google Scholar
Duursma, R. A., and Medlyn, B. E. (2012). MAESPA: A model to study interactions between water limitation, environmental drivers and vegetation function at tree and stand levels, with an example application to [CO2] × drought interactions. Geoscientific Model Development, 5, 919940.Google Scholar
Dyer, A. J. (1974). A review of flux–profile relationships. Boundary-Layer Meteorology, 7, 363372.Google Scholar
Dyer, A. J., and Hicks, B. B. (1970). Flux–gradient relationships in the constant flux layer. Quarterly Journal of the Royal Meteorological Society, 96, 715721.Google Scholar
Eagleson, P. S. (1978). Climate, soil, and vegetation 3. A simplified model of soil moisture movement in the liquid phase. Water Resources Research, 14, 722730.Google Scholar
Egea, G., Verhoef, A., and Vidale, P. L. (2011). Towards an improved and more flexible representation of water stress in coupled photosynthesis–stomatal conductance models. Agricultural and Forest Meteorology, 151, 13701384.Google Scholar
Ek, M. B., Mitchell, K. E., Lin, Y., et al. (2003). Implementation of Noah land surface model advances in the National Centers for Environmental Prediction operational mesoscale Eta model. Journal of Geophysical Research, 108, 8851, doi:10.1029/2002JD003296.Google Scholar
Ellsworth, D. S., and Reich, P. B. (1993). Canopy structure and vertical patterns of photosynthesis and related leaf traits in a deciduous forest. Oecologia, 96, 169178.Google Scholar
Eltahir, E. A. B., and Bras, R. L. (1993). A description of rainfall interception over large areas. Journal of Climate, 6, 10021008.Google Scholar
Entekhabi, D., and Eagleson, P. S. (1989). Land surface hydrology parameterization for atmospheric general circulation models including subgrid scale spatial variability. Journal of Climate, 2, 816831.Google Scholar
Essery, R., Best, M., and Cox, P. (2001). MOSES 2.2 Technical Documentation, Technical Note 30. Bracknell: Met Office Hadley Centre.Google Scholar
Essery, R. L. H., Best, M. J., Betts, R. A., Cox, P. M., and Taylor, C. M. (2003). Explicit representation of subgrid heterogeneity in a GCM land surface scheme. Journal of Hydrometeorology, 4, 530543.Google Scholar
Ethier, G. J., and Livingston, N. J. (2004). On the need to incorporate sensitivity to CO2 transfer conductance into the Farquhar–von Caemmerer–Berry leaf photosynthesis model. Plant, Cell and Environment, 27, 137153.Google Scholar
Evans, J. R. (1989). Photosynthesis and nitrogen relationships in leaves of C3 plants. Oecologia, 78, 919.Google Scholar
Evans, J. R. (1993a). Photosynthetic acclimation and nitrogen partitioning within a lucerne canopy. I. Canopy characteristics. Australian Journal of Plant Physiology, 20, 5567.Google Scholar
Evans, J. R. (1993b). Photosynthetic acclimation and nitrogen partitioning within a lucerne canopy. II. Stability through time and comparison with a theoretical optimum. Australian Journal of Plant Physiology, 20, 6982.Google Scholar
Ewers, B. E., Gower, S. T., Bond-Lamberty, B., and Wang, C. K. (2005). Effects of stand age and tree species on canopy transpiration and average stomatal conductance of boreal forests. Plant, Cell and Environment, 28, 660678.Google Scholar
Ewers, B. E., Oren, R., and Sperry, J. S. (2000). Influence of nutrient versus water supply on hydraulic architecture and water balance in Pinus taeda. Plant, Cell and Environment, 23, 10551066.Google Scholar
Famiglietti, J. S., and Wood, E. F. (1994). Multiscale modeling of spatially variable water and energy balance processes. Water Resources Research, 30, 30613078.Google Scholar
Farouki, O. T. (1981). Thermal Properties of Soils, CRREL Monograph 81–1. Hanover, NH: U.S Army Corps of Engineers, Cold Regions Research and Engineering Laboratory.Google Scholar
Farquhar, G. D., Buckley, T. N., and Miller, J. M. (2002). Optimal stomatal control in relation to leaf area and nitrogen content. Silva Fennica, 36, 625637.Google Scholar
Farquhar, G. D., and Cernusak, L. A. (2005). On the isotopic composition of leaf water in the non-steady state. Functional Plant Biology, 32, 293303.Google Scholar
Farquhar, G. D., Cernusak, L. A., and Barnes, B. (2007). Heavy water fractionation during transpiration. Plant Physiology, 143, 1118.Google Scholar
Farquhar, G. D., Ehleringer, J. R., and Hubick, K. T. (1989a). Carbon isotope discrimination and photosynthesis. Annual Review of Plant Physiology and Plant Molecular Biology, 40, 503537.Google Scholar
Farquhar, G. D., Hubick, K. T., Condon, A. G., and Richards, R. A. (1989b). Carbon isotope fractionation and plant water-use efficiency. In Stable Isotopes in Ecological Research, ed. Rundel, P. W., Ehleringer, J. R., and Nagy, K. A.. New York, NY: Springer-Verlag, pp. 2140.Google Scholar
Farquhar, G. D., and Lloyd, J. (1993). Carbon and oxygen isotope effects in the exchange of carbon dioxide between terrestrial plants and the atmosphere. In Stable Isotopes and Plant Carbon–Water Relations, ed. Ehleringer, J. R., Hall, A. E., and Farquhar, G. D.. San Diego, CA: Academic Press, pp. 4770.Google Scholar
Farquhar, G. D., O’Leary, M. H., and Berry, J. A. (1982). On the relationship between carbon isotope discrimination and the intercellular carbon dioxide concentration in leaves. Australian Journal of Plant Physiology, 9, 121137.Google Scholar
Farquhar, G. D., and Richards, R. A. (1984). Isotopic composition of plant carbon correlates with water-use efficiency of wheat genotypes. Australian Journal of Plant Physiology, 11, 539552.Google Scholar
Farquhar, G. D., and von Caemmerer, S. (1982). Modelling of photosynthetic response to environmental conditions. In Encyclopedia of Plant Physiology, New Series, vol. 12B. Physiological Plant Ecology. II. Water Relations and Carbon Assimilation, ed. Lange, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H.. Berlin: Springer-Verlag, pp. 549587.Google Scholar
Farquhar, G. D., von Caemmerer, S., and Berry, J. A. (1980). A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta, 149, 7890.Google Scholar
Farquhar, G. D., and Wong, S. C. (1984). An empirical model of stomatal conductance. Australian Journal of Plant Physiology, 11, 191210.Google Scholar
Feddes, R. A., and Raats, P. A. C. (2004). Parameterizing the soil – water – plant root system. In Unsaturated-zone Modeling: Progress, Challenges and Applications, ed. Feddes, R. A., de Rooij, G. H., and van Dam, J. C.. Dordrecht: Kluwer, pp. 95141.Google Scholar
Federer, C. A. (1979). A soil–plant–atmosphere model for transpiration and availability of soil water. Water Resources Research, 15, 555562.Google Scholar
Field, C. (1983). Allocating leaf nitrogen for the maximization of carbon gain: Leaf age as a control on the allocation program. Oecologia, 56, 341347.Google Scholar
Field, C., and Mooney, H. A. (1986). The photosynthesis–nitrogen relationship in wild plants. In On the Economy of Plant Form and Function, ed. Givnish, T. J.. Cambridge: Cambridge University Press, pp. 2555.Google Scholar
Finnigan, J. (2004). Advection and modeling. In Handbook of Micrometeorology: A Guide for Surface Flux Measurement and Analysis, ed. Lee, X., Massman, W., and Law, B.. Dordrecht: Kluwer, pp. 209244.Google Scholar
Finnigan, J., Harman, I., Ross, A., and Belcher, S. (2015). First-order turbulence closure for modelling complex canopy flows. Quarterly Journal of the Royal Meteorological Society, 141, 29072916.Google Scholar
Finnigan, J. J., and Raupach, M. R. (1987). Transfer processes in plant canopies in relation to stomatal characteristics. In Stomatal Function, ed. Zeiger, E., Farquhar, G. D., and Cowan, I. R.. Stanford, CA: Stanford University Press, pp. 385429.Google Scholar
Finnigan, J. J., Shaw, R. H., and Patton, E. G. (2009). Turbulence structure above a vegetation canopy. Journal of Fluid Mechanics, 637, 387424.Google Scholar
Firestone, M. K., and Davidson, E. A. (1989). Microbiological basis of NO and N2O production and consumption in soil. In Exchange of Trace Gases between Terrestrial Ecosystems and the Atmosphere, ed. Andreae, M. O. and Schimel, D. S.. New York, NY: Wiley, pp. 721.Google Scholar
Fisher, J. B., Huntzinger, D. N., Schwalm, C. R., and Sitch, S. (2014). Modeling the terrestrial biosphere. Annual Review of Environment and Resources, 39, 91123.Google Scholar
Fisher, J. B., Sitch, S., Malhi, Y., et al. (2010a). Carbon cost of plant nitrogen acquisition: A mechanistic, globally applicable model of plant nitrogen uptake, retranslocation, and fixation. Global Biogeochemical Cycles, 24, GB1014, doi:10.1029/2009GB003621.Google Scholar
Fisher, R., McDowell, N., Purves, D., et al. (2010b). Assessing uncertainties in a second-generation dynamic vegetation model caused by ecological scale limitations. New Phytologist, 187, 666681.Google Scholar
Fisher, R. A., Koven, C. D., Anderegg, W. R. L., et al. (2018). Vegetation demographics in Earth System Models: A review of progress and priorities. Global Change Biology, 24, 3554.Google Scholar
Fisher, R. A., Muszala, S., Vertenstein, M., et al. (2015). Taking off the training wheels: The properties of a dynamic vegetation model without climate envelopes, CLM4.5(ED). Geoscientific Model Development, 8, 35933619.Google Scholar
Fisher, R. A., Williams, M., Lola da Costa, A., et al. (2007). The response of an Eastern Amazonian rain forest to drought stress: Results and modelling analyses from a throughfall exclusion experiment. Global Change Biology, 13, 23612378.Google Scholar
Flanagan, L. B. (1993). Environmental and biological influences on the stable oxygen and hydrogen isotopic composition of leaf water. In Stable Isotopes and Plant Carbon–Water Relations, ed. Ehleringer, J. R., Hall, A. E., and Farquhar, G. D.. San Diego, CA: Academic Press, pp. 7190.Google Scholar
Flanagan, L. B., Cai, T., Black, T. A., et al. (2012). Measuring and modeling ecosystem photosynthesis and the carbon isotope composition of ecosystem-respired CO2 in three boreal coniferous forests. Agricultural and Forest Meteorology, 153, 165176.Google Scholar
Flanagan, L. B., Comstock, J. P., and Ehleringer, J. R. (1991). Comparison of modeled and observed environmental influences on the stable oxygen and hydrogen isotope composition of leaf water in Phaseolus vulgaris L. Plant Physiology, 96, 588596.Google Scholar
Flanner, M. G., Zender, C. S., Randerson, J. T., and Rasch, P. J. (2007). Present-day climate forcing and response from black carbon in snow. Journal of Geophysical Research, 112, D11202, doi:10.1029/2006JD008003.Google Scholar
Flatau, P. J., Walko, R. L., and Cotton, W. R. (1992). Polynomial fits to saturation vapor pressure. Journal of Applied Meteorology, 31, 15071513.Google Scholar
Flechard, C. R., Massad, R.-S., Loubet, B., et al. (2013). Advances in understanding, models and parameterizations of biosphere–atmosphere ammonia exchange. Biogeosciences, 10, 51835225.Google Scholar
Flexas, J., Brugnoli, E., and Warren, C. R. (2012). Mesophyll conductance to CO2. In Terrestrial Photosynthesis in a Changing Environment: A Molecular, Physiological and Ecological Approach, ed. Flexas, J., Loreto, F., and Medrano, H.. Cambridge: Cambridge University Press, pp. 169185.Google Scholar
Flexas, J., Ribas-Carbó, M., Diaz-Espejo, A., Galmés, J., and Medrano, H. (2008). Mesophyll conductance to CO2: Current knowledge and future prospects. Plant, Cell and Environment, 31, 602621.Google Scholar
Foken, T. (2006). 50 years of the Monin–Obukhov similarity theory. Boundary-Layer Meteorology, 119, 431447.Google Scholar
Foken, T. (2008). The energy balance closure problem: An overview. Ecological Applications, 18, 13511367.Google Scholar
Foley, J. A., Prentice, I. C., Ramankutty, N., et al. (1996). An integrated biosphere model of land surface processes, terrestrial carbon balance, and vegetation dynamics. Global Biogeochemical Cycles, 10, 603628.Google Scholar
Ford, E. D., and Newbould, P. J. (1971). The leaf canopy of a coppiced deciduous woodland: I. Development and structure. Journal of Ecology, 59, 843862.Google Scholar
Forkel, R., Klemm, O., Graus, M., et al. (2006). Trace gas exchange and gas phase chemistry in a Norway spruce forest: A study with a coupled 1-dimensional canopy atmospheric chemistry emission model. Atmospheric Environment, 40, S28S42.Google Scholar
Franks, P. J., Berry, J. A., Lombardozzi, D. L., and Bonan, G. B. (2017). Stomatal function across temporal and spatial scales: Deep-time trends, land–atmosphere coupling and global models. Plant Physiology, 174, 583602.Google Scholar
Freeze, R. A., and Harlan, R. L. (1969). Blueprint for a physically-based, digitally-simulated hydrologic response model. Journal of Hydrology, 9, 237258.Google Scholar
Friedlingstein, P., Cox, P., Betts, R., et al. (2006). Climate–carbon cycle feedback analysis: Results from the C4MIP model intercomparison. Journal of Climate, 19, 33373353.Google Scholar
Friedlingstein, P., Joel, G., Field, C. B., and Fung, I. Y. (1999). Toward an allocation scheme for global terrestrial carbon models. Global Change Biology, 5, 755770.Google Scholar
Friedlingstein, P., Meinshausen, M., Arora, V. K., et al. (2014). Uncertainties in CMIP5 climate projections due to carbon cycle feedbacks. Journal of Climate, 27, 511526.Google Scholar
Friend, A. D. (1995). PGEN: An integrated model of leaf photosynthesis, transpiration, and conductance. Ecological Modelling, 77, 233255.Google Scholar
Friend, A. D. (2001). Modelling canopy CO2 fluxes: Are ‘big-leaf’ simplifications justified? Global Ecology and Biogeography, 10, 603619.Google Scholar
Friend, A. D., and Kiang, N. Y. (2005). Land surface model development for the GISS GCM: Effects of improved canopy physiology on simulated climate. Journal of Climate, 18, 28832902.Google Scholar
Friend, A. D., Shugart, H. H., and Running, S. W. (1993). A physiology-based gap model of forest dynamics. Ecology, 74, 792797.Google Scholar
Friend, A. D., Stevens, A. K., Knox, R. G., and Cannell, M. G. R. (1997). A process-based, terrestrial biosphere model of ecosystem dynamics (Hybrid v3.0). Ecological Modelling, 95, 249287.Google Scholar
Fung, I. Y., Doney, S. C., Lindsay, K., and John, J. (2005). Evolution of carbon sinks in a changing climate. Proceedings of the National Academy of Sciences USA, 102, 1120111206.Google Scholar
Gao, W., and Wesely, M. L. (1994). Numerical modeling of the turbulent fluxes of chemically reactive trace gases in the atmospheric boundary layer. Journal of Applied Meteorology, 33, 835847.Google Scholar
Gao, W., Wesely, M. L., and Doskey, P. V. (1993). Numerical modeling of the turbulent diffusion and chemistry of NOx, O3, isoprene, and other reactive trace gases in and above a forest canopy. Journal of Geophysical Research, 98D, 1833918353.Google Scholar
Gardner, W. R. (1960). Dynamic aspects of water availability to plants. Soil Science, 89, 6373.Google Scholar
Garratt, J. R. (1978). Flux profile relations above tall vegetation. Quarterly Journal of the Royal Meteorological Society, 104, 199211.Google Scholar
Garratt, J. R. (1980). Surface influence upon vertical profiles in the atmospheric near-surface layer. Quarterly Journal of the Royal Meteorological Society, 106, 803819.Google Scholar
Garratt, J. R. (1983). Surface influence upon vertical profiles in the nocturnal boundary layer. Boundary-Layer Meteorology, 26, 6980.Google Scholar
Garratt, J. R. (1992). The Atmospheric Boundary Layer. Cambridge: Cambridge University Press.Google Scholar
Garratt, J. R., and Hicks, B. B. (1973). Momentum, heat and water vapour transfer to and from natural and artificial surfaces. Quarterly Journal of the Royal Meteorological Society, 99, 680687.Google Scholar
Gastal, F., and Lemaire, G. (2002). N uptake and distribution in crops: An agronomical and ecophysiological perspective. Journal of Experimental Botany, 53, 789799.Google Scholar
Gastellu-Etchegorry, J. P., Demarez, V., Pinel, V., and Zagolski, F. (1996). Modeling radiative transfer in heterogeneous 3-D vegetation canopies. Remote Sensing of Environment, 58, 131156.Google Scholar
Gat, J. R. (1996). Oxygen and hydrogen isotopes in the hydrologic cycle. Annual Review of Earth and Planetary Sciences, 24, 225262.Google Scholar
Gates, D. M. (1962). Energy Exchange in the Biosphere. New York, NY: Harper and Row.Google Scholar
Gates, D. M. (1963). Leaf temperature and energy exchange. Archiv für Meteorologie, Geophysik und Bioklimatologie, 12B, 321336.Google Scholar
Gates, D. M. (1965). Energy, plants, and ecology. Ecology, 46, 113.Google Scholar
Gates, D. M. (1966). Transpiration and energy exchange. Quarterly Review of Biology, 41, 353364.Google Scholar
Gates, D. M. (1980). Biophysical Ecology. New York, NY: Springer-Verlag.Google Scholar
Gates, D. M., Alderfer, R., and Taylor, E. (1968). Leaf temperatures of desert plants. Science, 159, 994995.Google Scholar
Gates, D. M., and Papian, L. E. (1971). Atlas of Energy Budgets of Plant Leaves. New York, NY: Academic Press.Google Scholar
Geiger, R. (1927). Das Klima der bodennahen Luftschicht. Braunschweig, Germany: Friedr. Vieweg & Sohn.Google Scholar
Gelhar, L. W. (1986). Stochastic subsurface hydrology from theory to applications. Water Resources Research, 22, 135S145S.Google Scholar
Ghimire, B., Riley, W. J., Koven, C. D., Mu, M., and Randerson, J. T. (2016). Representing leaf and root physiological traits in CLM improves global carbon and nitrogen cycling predictions. Journal of Advances in Modeling Earth Systems, 8, 598613, doi:10.1002/2015MS000538.Google Scholar
Gholz, H. L., Wedin, D. A., Smitherman, S. M., Harmon, M. E., and Parton, W. J. (2000). Long-term dynamics of pine and hardwood litter in contrasting environments: Toward a global model of decomposition. Global Change Biology, 6, 751765.Google Scholar
Givnish, T. J., and Vermeij, G. J. (1976). Sizes and shapes of liane leaves. American Naturalist, 110, 743778.Google Scholar
Goel, N. S., and Strebel, D. E. (1984). Simple beta distribution representation of leaf orientation in vegetation canopies. Agronomy Journal, 76, 800802.Google Scholar
Goll, D. S., Vuichard, N., Maignan, F., et al. (2017). A representation of the phosphorus cycle for ORCHIDEE (revision 4520). Geoscientific Model Development, 10, 37453770.Google Scholar
Golley, F. B. (1993). A History of the Ecosystem Concept in Ecology: More than the Sum of the Parts. New Haven, CT: Yale University Press.Google Scholar
Goudriaan, J. (1977). Crop Micrometeorology: A Simulation Study. Wageningen: Center for Agricultural Publishing and Documentation.Google Scholar
Goudriaan, J. (1982). Potential production processes. In Simulation of Plant Growth and Crop Production, ed. de Vries, F. W. T. Penning, and van Laar, H. H.. Wageningen: Center for Agricultural Publishing and Documentation, pp. 98113.Google Scholar
Goudriaan, J. (1988). The bare bones of leaf-angle distribution in radiation models for canopy photosynthesis and energy exchange. Agricultural and Forest Meteorology, 43, 155169.Google Scholar
Goudriaan, J., and van Laar, H. H. (1994). Modelling Potential Crop Growth Processes: Textbook with Exercises. Dordrecht: Kluwer.Google Scholar
Goulden, M. L., McMillan, A. M. S., Winston, G. C., et al. (2011). Patterns of NPP, GPP, respiration, and NEP during boreal forest succession. Global Change Biology, 17, 855871.Google Scholar
Grace, J. (1981). Some effects of wind on plants. In Plants and their Atmospheric Environment, ed. Grace, J., Ford, E. D., and Jarvis, P. G.. Oxford: Blackwell, pp. 3156.Google Scholar
Green, W. H., and Ampt, G. A. (1911). Studies on soil physics: Part I – The flow of air and water through soils. Journal of Agricultural Science, 4, 124.Google Scholar
Grindlay, D. J. C. (1997). Towards an explanation of crop nitrogen demand based on the optimization of leaf nitrogen per unit leaf area. Journal of Agricultural Science, 128, 377396.Google Scholar
Grote, R., Morfopoulos, C., Niinemets, Ü., et al. (2014). A fully integrated isoprenoid emissions model coupling emissions to photosynthetic characteristics. Plant, Cell and Environment, 37, 19651980.Google Scholar
Gu, L., Shugart, H. H., Fuentes, J. D., Black, T. A., and Shewchuk, S. R. (1999). Micrometeorology, biophysical exchanges and NEE decomposition in a two-story boreal forest – development and test of an integrated model. Agricultural and Forest Meteorology, 94, 123148.Google Scholar
Gu, L., Pallardy, S. G., Tu, K., Law, B. E., and Wullschleger, S. D. (2010). Reliable estimation of biochemical parameters from C3 leaf photosynthesis–intercellular carbon dioxide response curves. Plant, Cell and Environment, 33, 18521874.Google Scholar
Guenther, A., Hewitt, C. N., Erickson, D., et al. (1995). A global model of natural volatile organic compound emissions. Journal of Geophysical Research, 100D, 88738892.Google Scholar
Guenther, A., Karl, T., Harley, P., et al. (2006). Estimates of global terrestrial isoprene emissions using MEGAN (Model of Emissions of Gases and Aerosols from Nature). Atmospheric Chemistry and Physics, 6, 31813210.Google Scholar
Guenther, A. B., Jiang, X., Heald, C. L., et al. (2012). The Model of Emissions of Gases and Aerosols from Nature version 2.1 (MEGAN2.1): An extended and updated framework for modeling biogenic emissions. Geoscientific Model Development, 5, 14711492.Google Scholar
Gutschick, V. P. (1991). Joining leaf photosynthesis models and canopy photon-transport models. In Photon–Vegetation Interactions: Applications in Optical Remote Sensing and Plant Ecology, ed. Myneni, R. B. and Ross, J.. Berlin: Springer-Verlag, pp 501535.Google Scholar
Gutschick, V. P., and Wiegel, F. W. (1988). Optimizing the canopy photosynthetic rate by patterns of investment in specific leaf mass. American Naturalist, 132, 6786.Google Scholar
Haese, B., Werner, M., and Lohmann, G. (2013). Stable water isotopes in the coupled atmosphere–land surface model ECHAM5-JSBACH. Geoscientific Model Development, 6, 14631480.Google Scholar
Hagemann, S., and Gates, L. D. (2003). Improving a subgrid runoff parameterization scheme for climate models by the use of high resolution data derived from satellite observations. Climate Dynamics, 21, 349359.Google Scholar
Halldin, S. (1985). Leaf and bark area distribution in a pine forest. In The Forest–Atmosphere Interaction, ed. Hutchinson, B. A. and Hicks, B. B.. Dordrecht: Reidel, pp. 3958.Google Scholar
Hantson, S., Arneth, A., Harrison, S. P., et al. (2016). The status and challenge of global fire modelling. Biogeosciences, 13, 33593375.Google Scholar
Hari, P., Mäkelä, A., Korpilahti, E., and Holmberg, M. (1986). Optimal control of gas exchange. Tree Physiology, 2, 169175.Google Scholar
Harley, P. C., and Sharkey, T. D. (1991). An improved model of C3 photosynthesis at high CO2: Reversed O2 sensitivity explained by lack of glycerate reentry into the chloroplast. Photosynthesis Research, 27, 169178.Google Scholar
Harley, P. C., Thomas, R. B., Reynolds, J. F., and Strain, B. R. (1992). Modelling photosynthesis of cotton grown in elevated CO2. Plant, Cell and Environment, 15, 271282.Google Scholar
Harman, I. N. (2012). The role of roughness sublayer dynamics within surface exchange schemes. Boundary-Layer Meteorology, 142, 120.Google Scholar
Harman, I. N., and Finnigan, J. J. (2007). A simple unified theory for flow in the canopy and roughness sublayer. Boundary-Layer Meteorology, 123, 339363.Google Scholar
Harman, I. N., and Finnigan, J. J. (2008). Scalar concentration profiles in the canopy and roughness sublayer. Boundary-Layer Meteorology, 129, 323351.Google Scholar
Harmon, M. E., Silver, W. L., Fasth, B., et al. (2009). Long-term patterns of mass loss during the decomposition of leaf and fine root litter: An intersite comparison. Global Change Biology, 15, 13201338.Google Scholar
Hartman, M. D., Merchant, E. R., Parton, W. J., et al. (2011). Impact of historical land-use changes on greenhouse gas exchange in the U.S. Great Plains, 1883–2003. Ecological Applications, 21, 11051119.Google Scholar
Hartmann, D. L. (1994). Global Physical Climatology. San Diego, CA: Academic Press.Google Scholar
Haverd, V., Cuntz, M., Leuning, R., and Keith, H. (2007). Air and biomass heat storage fluxes in a forest canopy: Calculation within a soil vegetation atmosphere transfer model. Agricultural and Forest Meteorology, 147, 125139.Google Scholar
Haverd, V., Cuntz, M., Nieradzik, L. P., and Harman, I. N. (2016). Improved representations of coupled soil–canopy processes in the CABLE land surface model (Subversion revision 3432). Geoscientific Model Development, 9, 31113122.Google Scholar
Haverd, V., Leuning, R., Griffith, D., van Gorsel, E., and Cuntz, M. (2009). The turbulent Lagrangian time scale in forest canopies constrained by fluxes, concentrations and source distributions. Boundary-Layer Meteorology, 130, 209228.Google Scholar
Haverkamp, R., and Vauclin, M. (1979). A note on estimating finite difference interblock hydraulic conductivity values for transient unsaturated flow problems. Water Resources Research, 15, 181187.Google Scholar
Haverkamp, R., Vauclin, M., Touma, J., Wierenga, P. J., and Vachaud, G. (1977). A comparison of numerical simulation models for one-dimensional infiltration. Soil Science Society of America Journal, 41, 285294.Google Scholar
Heald, C. L., Henze, D. K., Horowitz, L. W., et al. (2008). Predicted change in global secondary organic aerosol concentrations in response to future climate, emissions, and land use change. Journal of Geophysical Research, 113, D05211, doi:10.1029/2007JD009092.Google Scholar
Hendricks Franssen, H. J., Stöckli, R., Lehner, I., Rotenberg, E., and Seneviratne, S. I. (2010). Energy balance closure of eddy-covariance data: A multisite analysis for European FLUXNET stations. Agricultural and Forest Meteorology, 150, 15531567.Google Scholar
Hetherington, A. M., and Woodward, F. I. (2003). The role of stomata in sensing and driving environmental change. Nature, 424, 901908.Google Scholar
Hicks, B. B., Baldocchi, D. D., Meyers, T. P., Hosker, R. P., Jr., and Matt, D. R. (1987). A preliminary multiple resistance routine for deriving dry deposition velocities from measured quantities. Water, Air, and Soil Pollution, 36, 311330.Google Scholar
Hill, T. C., Williams, M., Woodward, F. I., and Moncrieff, J. B. (2011). Constraining ecosystem processes from tower fluxes and atmospheric profiles. Ecological Applications, 21, 14741489.Google Scholar
Hills, R. G., Hudson, D. B., Porro, I., and Wierenga, P. J. (1989). Modeling one-dimensional infiltration into very dry soils: 2. Estimation of the soil water parameters and model predictions. Water Resources Research, 25, 12711282.Google Scholar
Hirose, T., and Werger, M. J. A. (1987). Maximizing daily canopy photosynthesis with respect to the leaf nitrogen allocation pattern in the canopy. Oecologia, 72, 520526.Google Scholar
Hirose, T., and Werger, M. J. A. (1994). Photosynthetic capacity and nitrogen partitioning among species in the canopy of a herbaceous plant community. Oecologia, 100, 203212.Google Scholar
Hirose, T., Werger, M. J. A., Pons, T. L., and van Rheenen, J. W. A. (1988). Canopy structure and leaf nitrogen distribution in a stand of Lysimachia vulgaris L. as influenced by stand density. Oecologia, 77, 145150.Google Scholar
Hirose, T., Werger, M. J. A., and van Rheenen, J. W. A. (1989). Canopy development and leaf nitrogen distribution in a stand of Carex acutiformis. Ecology, 70, 16101618.Google Scholar
Hoelzemann, J. J., Schultz, M. G., Brasseur, G. P., Granier, C., and Simon, M. (2004). Global Wildland Fire Emission Model (GWEM): Evaluating the use of global area burnt satellite data. Journal of Geophysical Research, 109, D14S04, doi:10.1029/2003JD003666.Google Scholar
Hollinger, D. Y. (1989). Canopy organization and foliage photosynthetic capacity in a broad-leaved evergreen montane forest. Functional Ecology, 3, 562.Google Scholar
Hollinger, D. Y. (1996). Optimality and nitrogen allocation in a tree canopy. Tree Physiology, 16, 627634.Google Scholar
Holtslag, A. A. M., and Beljaars, A. C. M. (1989). Surface flux parameterization schemes: Developments and experiences at KNMI. In Parameterization of Fluxes over Land Surface: Proceedings of a Workshop Held at ECMWF 24–26 October 1988. European Centre for Medium-Range Weather Forecasts, Reading, UK, pp. 121147.Google Scholar
Holtslag, A. A. M., Svensson, G., Baas, P., et al. (2013). Stable atmospheric boundary layers and diurnal cycles: Challenges for weather and climate models. Bulletin of the American Meteorological Society, 94, 16911706.Google Scholar
Horn, H. S. (1971). The Adaptive Geometry of Trees. Princeton, NJ: Princeton University Press.Google Scholar
Horn, H. S. (1975). Forest succession. Scientific American, 232(5), 9098.Google Scholar
Hornberger, G., and Wiberg, P. (2005). Numerical Methods in the Hydrological Sciences. Washington, DC: American Geophysical Union.Google Scholar
Hornberger, G. M., Raffensperger, J. P., Wiberg, P. L., and Eshleman, K. N. (1998). Elements of Physical Hydrology. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Hrachowitz, M., and Clark, M. P. (2017). HESS Opinions: The complementary merits of competing modelling philosophies in hydrology. Hydrology and Earth System Sciences, 21, 39533973.Google Scholar
Huang, Y., Lu, X., Shi, Z., et al. (2018). Matrix approach to land carbon cycle modeling: A case study with the Community Land Model. Global Change Biology, 24, 13941404.Google Scholar
Hurtt, G. C., Chini, L. P., Frolking, S., et al. (2011). Harmonization of land–use scenarios for the period 1500–2100: 600 years of global gridded annual land–use transitions, wood harvest, and resulting secondary lands. Climatic Change, 109, 117161.Google Scholar
Hurtt, G. C., Frolking, S., Fearon, M. G., et al. (2006). The underpinnings of land-use history: Three centuries of global gridded land-use transitions, wood-harvest activity, and resulting secondary lands. Global Change Biology, 12, 12081229.Google Scholar
Hurtt, G. C., Moorcroft, P. R., Pacala, S. W., and Levin, S. A. (1998). Terrestrial models and global change: Challenges for the future. Global Change Biology, 4, 581590.Google Scholar
Hutchinson, B. A., Matt, D. R., McMillen, R. T., et al. (1986). The architecture of a deciduous forest canopy in eastern Tennessee, U.S.A. Journal of Ecology, 74, 635646.Google Scholar
Innis, G. S. (1975). Role of total systems models in the grassland biome study. In Systems Analysis and Simulation in Ecology, vol. III. ed. Patten, B. C.. New York, NY: Academic Press, pp. 1347.Google Scholar
Innis, G. S. (1978). Grassland Simulation Model. New York, NY: Springer-Verlag.Google Scholar
Inoue, E. (1963). On the turbulent structure of airflow within crop canopies. Journal of the Meteorological Society of Japan Ser. II, 41, 317326.Google Scholar
Jackson, R. B., Canadell, J., Ehleringer, J. R., et al. (1996). A global analysis of root distributions for terrestrial biomes. Oecologia, 108, 389411.Google Scholar
Jackson, R. B., Mooney, H. A., and Schulze, E.-D. (1997). A global budget for fine root biomass, surface area, and nutrient contents. Proceedings of the National Academy of Sciences USA, 94, 73627366.Google Scholar
Jacobson, M. Z. (2005). Fundamentals of Atmospheric Modeling, 2nd edn. Cambridge: Cambridge University Press.Google Scholar
Jain, A., Yang, X., Kheshgi, H., et al. (2009). Nitrogen attenuation of terrestrial carbon cycle response to global environmental factors. Global Biogeochemical Cycles, 23, GB4028, doi:10.1029/2009GB003519.Google 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 B, 273, 593610.Google Scholar
Jarvis, P. G., and McNaughton, K. G. (1986). Stomatal control of transpiration: Scaling up from leaf to region. Advances in Ecological Research, 15, 149.Google Scholar
Jenkinson, D. S. (1990). The turnover of organic carbon and nitrogen in soil. Philosophical Transactions of the Royal Society London B, 329, 361368.Google Scholar
Jenkinson, D. S., Adams, D. E., and Wild, A. (1991). Model estimates of CO2 emissions from soil in response to global warming. Nature, 351, 304306.Google Scholar
Jetten, V. G. (1996). Interception of tropical rain forest: Performance of a canopy water balance model. Hydrological Processes, 10, 671685.Google Scholar
Ji, J. (1995). A climate–vegetation interaction model: Simulating physical and biological processes at the surface. Journal of Biogeography, 22, 445451.Google Scholar
Jobbágy, E. G., and Jackson, R. B. (2000). The vertical distribution of soil organic carbon and its relation to climate and vegetation. Ecological Applications, 10, 423436.Google Scholar
Johnson, I. R., and Thornley, J. H. M. (1984). A model of instantaneous and daily canopy photosynthesis. Journal of Theoretical Biology, 107, 531545.Google Scholar
Johnson, K. D., Entekhabi, D., and Eagleson, P. S. (1993). The implementation and validation of improved land-surface hydrology in an atmospheric general circulation model. Journal of Climate, 6, 10091026.Google Scholar
Jones, H. G. (2014). Plants and Microclimate: A Quantitative Approach to Environmental Plant Physiology, 3rd edn. Cambridge: Cambridge University Press.Google Scholar
Jordan, R. (1991). A One-Dimensional Temperature Model for a Snow Cover: Technical Documentation for SNTHERM.89, Special Report 91–16. Hanover, NH: U.S. Army Corps of Engineers Cold Regions Research and Engineering Laboratory.Google Scholar
Juang, J.-Y., Katul, G. G., Siqueira, M. B., Stoy, P. C., and McCarthy, H. R. (2008). Investigating a hierarchy of Eulerian closure models for scalar transfer inside forested canopies. Boundary-Layer Meteorology, 128, 132.Google Scholar
Juang, J.-Y., Katul, G., Siqueira, M., Stoy, P., and Novick, K. (2007). Separating the effects of albedo from eco-physiological changes on surface temperature along a successional chronosequence in the southeastern United States. Geophysical Research Letters, 34, L21408, doi:10.1029/2007GL031296.Google Scholar
Jumikis, A. R. (1966). Thermal Soil Mechanics. New Brunswick, NJ: Rutgers University Press.Google Scholar
Kala, J., De Kauwe, M. G., Pitman, A. J., et al. (2015). Implementation of an optimal stomatal conductance scheme in the Australian Community Climate Earth Systems Simulator (ACCESS1.3b). Geoscientific Model Development, 8, 38773889.Google Scholar
Kattge, J., and Knorr, W. (2007). Temperature acclimation in a biochemical model of photosynthesis: A reanalysis of data from 36 species. Plant, Cell and Environment, 30, 11761190.Google Scholar
Kattge, J., Knorr, W., Raddatz, T., and Wirth, C. (2009). Quantifying photosynthetic capacity and its relationship to leaf nitrogen content for global-scale terrestrial biosphere models. Global Change Biology, 15, 976991.Google Scholar
Katul, G. G., Finnigan, J. J., Poggi, D., Leuning, R., and Belcher, S. E. (2006). The influence of hilly terrain on canopy-atmosphere carbon dioxide exchange. Boundary-Layer Meteorology, 118, 189216.Google Scholar
Katul, G. G., Mahrt, L., Poggi, D., and Sanz, C. (2004). One- and two-equation models for canopy turbulence. Boundary-Layer Meteorology, 113, 81109.Google Scholar
Katul, G., Manzoni, S., Palmroth, S., and Oren, R. (2010). A stomatal optimization theory to describe the effects of atmospheric CO2 on leaf photosynthesis and transpiration. Annals of Botany, 105, 431442.Google 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, RG3002, doi:10.1029/2011RG000366.Google Scholar
Katul, G. G., Palmroth, S., and Oren, R. (2009). Leaf stomatal responses to vapour pressure deficit under current and CO2-enriched atmosphere explained by the economics of gas exchange. Plant, Cell and Environment, 32, 968979.Google Scholar
Keenan, T., Sabate, S., and Gracia, C. (2010). Soil water stress and coupled photosynthesis–conductance models: Bridging the gap between conflicting reports on the relative roles of stomatal, mesophyll conductance and biochemical limitations to photosynthesis. Agricultural and Forest Meteorology, 150, 443453.Google Scholar
Keenan, T. F., and Niinemets, Ü. (2016). Global leaf trait estimates biased due to plasticity in the shade. Nature Plants, 3, 16201, doi:10.1038/nplants.2016.201.Google Scholar
Keller, K. M., Lienert, S., Bozbiyik, A., et al. (2017). 20th century changes in carbon isotopes and water-use efficiency: Tree-ring-based evaluation of the CLM4.5 and LPX-Bern models. Biogeosciences, 14, 26412673.Google 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, 116.Google Scholar
Kelly, R. H., Parton, W. J., Hartman, M. D., et al. (2000). Intra-annual and interannual variability of ecosystem processes in shortgrass steppe. Journal of Geophysical Research, 105D, 2009320100.Google Scholar
Kloster, S., Mahowald, N. M., Randerson, J. T., et al. (2010). Fire dynamics during the 20th century simulated by the Community Land Model. Biogeosciences, 7, 18771902.Google Scholar
Knauer, J., Werner, C., and Zaehle, S. (2015). Evaluating stomatal models and their atmospheric drought response in a land surface scheme: A multibiome analysis. Journal of Geophysical Research: Biogeosciences, 120, 18941911, doi:10.1002/2015JG003114.Google Scholar
Knauer, J., Zaehle, S., Medlyn, B. E., et al. (2018). Towards physiologically meaningful water-use efficiency estimates from eddy covariance data. Global Change Biology, 24, 694710.Google Scholar
Knorr, W., and Heimann, M. (2001). Uncertainties in global terrestrial biosphere modeling 1. A comprehensive sensitivity analysis with a new photosynthesis and energy balance scheme. Global Biogeochemical Cycles, 15, 207225.Google Scholar
Knox, R. G., Longo, M., Swann, A. L. S., et al. (2015). Hydrometeorological effects of historical land-conversion in an ecosystem-atmosphere model of Northern South America. Hydrology and Earth System Sciences, 19, 241273.Google Scholar
Kobayashi, H., Baldocchi, D. D., Ryu, Y., et al. (2012). Modeling energy and carbon fluxes in a heterogeneous oak woodland: A three-dimensional approach. Agricultural and Forest Meteorology, 152, 83100.Google Scholar
Kondo, J., and Kawanaka, A. (1986). Numerical study on the bulk heat transfer coefficient for a variety of vegetation types and densities. Boundary-Layer Meteorology, 37, 285296.Google Scholar
Konrad, W., Roth-Nebelsick, A., and Grein, M. (2008). Modelling of stomatal density response to atmospheric CO2. Journal of Theoretical Biology, 253, 638658.Google Scholar
Koster, R. D., and Milly, P. C. D. (1997). The interplay between transpiration and runoff formulations in land surface schemes used with atmospheric models. Journal of Climate, 10, 15781591.Google Scholar
Koster, R. D., and Suarez, M. J. (1992). A comparative analysis of two land surface heterogeneity representations. Journal of Climate, 5, 13791390.Google Scholar
Koven, C. D., Riley, W. J., Subin, Z. M., et al. (2013). The effect of vertically resolved soil biogeochemistry and alternate soil C and N models on C dynamics of CLM4. Biogeosciences, 10, 71097131.Google Scholar
Kowalczyk, E. A., Stevens, L., Law, R. M., et al. (2013). The land surface model component of ACCESS: Description and impact on the simulated surface climatology. Australian Meteorological and Oceanographic Journal, 63, 6582.Google Scholar
Kowalczyk, E. A., Wang, Y. P., Law, R. M., et al. (2006). The CSIRO Atmosphere Biosphere Land Exchange (CABLE) Model for Use in Climate Models and as an Offline Model, Research Paper 13. Aspendale, Australia: CSIRO Marine and Atmospheric Research.Google Scholar
Krinner, G., Viovy, N., de Noblet-Ducoudré, N., et al. (2005). A dynamic global vegetation model for studies of the coupled atmosphere–biosphere system. Global Biogeochemical Cycles, 19, GB1015, doi:10.1029/2003GB002199.Google Scholar
Kucharik, C. J., Foley, J. A., Delire, C., et al. (2000). Testing the performance of a Dynamic Global Ecosystem Model: Water balance, carbon balance, and vegetation structure. Global Biogeochemical Cycles, 14, 795825.Google Scholar
Kull, O., Broadmeadow, M., Kruijt, B., and Meir, P. (1999). Light distribution and foliage structure in an oak canopy. Trees, 14, 5564.Google Scholar
Kull, O., and Kruijt, B. (1998). Leaf photosynthetic light response: A mechanistic model for scaling photosynthesis to leaves and canopies. Functional Ecology, 12, 767777.Google Scholar
Kumagai, T. (2011). Transpiration in forest ecosystems. In Forest Hydrology and Biogeochemistry: Synthesis of Past Research and Future Directions, ed. Levia, D. F., Carlyle-Moses, D., and Tanaka, T.. Dordrecht: Springer, pp. 389406.Google Scholar
Kustas, W. P., Anderson, M. C., Norman, J. M, and Li, F. (2007). Utility of radiometric–aerodynamic temperature relations for heat flux estimation. Boundary-Layer Meteorology, 122, 167187.Google Scholar
Kustas, W. P., and Norman, J. M. (1999). Evaluation of soil and vegetation heat flux predictions using a simple two-source model with radiometric temperatures for partial canopy cover. Agricultural and Forest Meteorology, 94, 1329.Google Scholar
Kwa, C. (1993). Modeling the grasslands. Historical Studies in the Physical and Biological Sciences, 24, 125155.Google Scholar
Kwa, C. (2005). Local ecologies and global science: Discourses and strategies of the International Geosphere–Biosphere Programme. Social Studies of Science, 35, 923950.Google Scholar
Lasslop, G., Thonicke, K., and Kloster, S. (2014). SPITFIRE within the MPI Earth system model: Model development and evaluation. Journal of Advances in Modeling Earth Systems, 6, 740755, doi:10.1002/2013MS000284.Google Scholar
Lathière, J., Hauglustaine, D. A., Friend, A. D., et al. (2006). Impact of climate variability and land use changes on global biogenic volatile organic compound emissions. Atmospheric Chemistry and Physics, 6, 21292146.Google Scholar
Launiainen, S., Katul, G. G., Grönholm, T., and Vesala, T. (2013). Partitioning ozone fluxes between canopy and forest floor by measurements and a multi-layer model. Agricultural and Forest Meteorology, 173, 8599.Google Scholar
Launiainen, S., Katul, G. G., Kolari, P., Vesala, T., and Hari, P. (2011). Empirical and optimal stomatal controls on leaf and ecosystem level CO2 and H2O exchange rates. Agricultural and Forest Meteorology, 151, 16721689.Google Scholar
Launiainen, S., Katul, G. G., Lauren, A., and Kolari, P. (2015). Coupling boreal forest CO2, H2O and energy flows by a vertically structured forest canopy – Soil model with separate bryophyte layer. Ecological Modelling, 312, 385405.Google Scholar
Lawrence, D., Fisher, R., Koven, C., et al. (2018). Technical Description of Version 5.0 of the Community Land Model (CLM). https://escomp.github.io/ctsm-docs/.Google Scholar
Lawrence, D. M., Hurtt, G. C., Arneth, A., et al. (2016). The Land Use Model Intercomparison Project (LUMIP) contribution to CMIP6: Rationale and experimental design. Geoscientific Model Development, 9, 29732998.Google Scholar
Lawrence, P. J., Feddema, J. J., Bonan, G. B., et al. (2012). Simulating the biogeochemical and biogeophysical impacts of transient land cover change and wood harvest in the Community Climate System Model (CCSM4) from 1850 to 2100. Journal of Climate, 25, 30713095.Google Scholar
Lee, J.-E., Oliveira, R. S., Dawson, T. E., and Fung, I. (2005). Root functioning modifies seasonal climate. Proceedings of the National Academy of Sciences USA, 102, 1757617581.Google Scholar
Lee, T. J., and Pielke, R. A. (1992). Estimating the soil surface specific humidity. Journal of Applied Meteorology, 31, 480484.Google Scholar
Lehmann, F., and Ackerer, Ph. (1998). Comparison of iterative methods for improved solutions of the fluid flow equation in partially saturated porous media. Transport in Porous Media, 31, 275292.Google Scholar
Lehnebach, R., Beyer, R., Letort, V., and Heuret, P. (2018). The pipe model theory half a century on: A review. Annals of Botany, 121, 773795.Google Scholar
Leigh, A., Sevanto, S., Ball, M. C., et al. (2012). Do thick leaves avoid thermal damage in critically low wind speeds? New Phytologist, 194, 477487.Google Scholar
Leij, F. J., Alves, W. J., van Genuchten, M. T., and Williams, J. R. (1996). UNSODA – The UNSODA Unsaturated Soil Hydraulic Database User’s Manual Version 1.0, EPA/600/R-96/095. Cincinnati, OH: U.S. Environmental Protection Agency.Google Scholar
Lemaire, G., Onillon, B., Gosse, G., Chartier, M., and Allirand, J. M. (1991). Nitrogen distribution within a lucerne canopy during regrowth: Relation with light distribution. Annals of Botany, 68, 483488.Google Scholar
Lemeur, R. (1973). A method for simulating the direct solar radiation regime in sunflower, Jerusalem artichoke, corn and soybean canopies using actual stand structure data. Agricultural Meteorology, 12, 229247.Google Scholar
Lemeur, R. (1990). Modelling stomatal behaviour and photosynthesis of Eucalyptus grandis. Australian Journal of Plant Physiology, 17, 159175.Google Scholar
Lemeur, R. (1995). A critical appraisal of a combined stomatal–photosynthesis model for C3 plants. Plant, Cell and Environment, 18, 339355.Google Scholar
Lemeur, R. (2000). Estimation of scalar source/sink distributions in plant canopies using Lagrangian dispersion analysis: Corrections for atmospheric stability and comparison with a multilayer canopy model. Boundary-Layer Meteorology, 96, 293314.Google Scholar
Lemeur, R. (2002). Temperature dependence of two parameters in a photosynthesis model. Plant, Cell and Environment, 25, 12051210.Google Scholar
Leuning, R., Kelliher, F. M., de Pury, D. G. G., and Schulze, E.-D. (1995). Leaf nitrogen, photosynthesis, conductance and transpiration: Scaling from leaves to canopies. Plant, Cell and Environment, 18, 11831200.Google Scholar
Leuning, R., van Gorsel, E., Massman, W. J., and Issac, P. R. (2012). Reflections on the surface energy imbalance problem. Agricultural and Forest Meteorology, 156, 6574.Google Scholar
Levis, S., Foley, J. A., and Pollard, D. (1999). CO2, climate, and vegetation feedbacks at the Last Glacial Maximum. Journal of Geophysical Research, 104D, 3119131198.Google Scholar
Levis, S., Foley, J. A., and Pollard, D. (2000). Large-scale vegetation feedbacks on a doubled CO2 climate. Journal of Climate, 13, 13131325.Google Scholar
Levis, S., Bonan, G. B., Vertenstein, M., and Oleson, K. W. (2004). The Community Land Model’s Dynamic Global Vegetation Model (CLM-DGVM): Technical Description and User’s Guide, Technical Note NCAR/TN-459+IA. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Levis, S., Hartman, M. D., and Bonan, G. B. (2014). The Community Land Model underestimates land-use CO2 emissions by neglecting soil disturbance from cultivation. Geoscientific Model Development, 7, 613620.Google Scholar
Levis, S., Wiedinmyer, C., Bonan, G. B., and Guenther, A. (2003). Simulating biogenic volatile organic compound emissions in the Community Climate System Model. Journal of Geophysical Research, 108, 4659, doi:10.1029/2002JD003203.Google Scholar
Li, C., Frolking, S., and Frolking, T. A. (1992). A model of nitrous oxide evolution from soil driven by rainfall events: 1. Model structure and sensitivity. Journal of Geophysical Research, 97D, 97599776.Google Scholar
Li, C., Aber, J., Stange, F., Butterbach-Bahl, K., and Papen, H. (2000). A process-oriented model of N2O and NO emissions from forest soils: 1. Model development. Journal of Geophysical Research, 105D, 43694384.Google Scholar
Li, F., Levis, S., and Ward, D. S. (2013). Quantifying the role of fire in the Earth system – Part 1: Improved global fire modeling in the Community Earth System Model (CESM1). Biogeosciences, 10, 22932314.Google Scholar
Li, F., Zeng, X. D., and Levis, S. (2012a). A process-based fire parameterization of intermediate complexity in a Dynamic Global Vegetation Model. Biogeosciences, 9, 27612780.Google Scholar
Li, H., Huang, M., Wigmosta, M. S., et al. (2011). Evaluating runoff simulations from the Community Land Model 4.0 using observations from flux towers and a mountainous watershed. Journal of Geophysical Research, 116, D24120, doi:10.1029/2011JD016276.Google Scholar
Li, L., Wang, Y.-P., Yu, Q., et al. (2012b). Improving the responses of the Australian community land surface model (CABLE) to seasonal drought. Journal of Geophysical Research, 117, G04002, doi:10.1029/2012JG002038.Google Scholar
Liang, X., Lettenmaier, D. P., and Wood, E. F. (1996). One-dimensional statistical dynamic representation of subgrid spatial variability of precipitation in the two-layer variable infiltration capacity model. Journal of Geophysical Research, 101D, 2140321422.Google Scholar
Liang, X., Lettenmaier, D. P., Wood, E. F., and Burges, S. J. (1994). A simple hydrologically based model of land surface water and energy fluxes for general circulation models. Journal of Geophysical Research, 99D, 1441514428.Google Scholar
Lin, Y.-S., Medlyn, B. E., Duursma, R. A., et al. (2015). Optimal stomatal behaviour around the world. Nature Climate Change, 5, 459464.Google Scholar
Lindroth, A., Mölder, M., and Lagergren, F. (2010). Heat storage in forest biomass improves energy balance closure. Biogeosciences, 7, 301313.Google Scholar
Liou, K. N. (2002). An Introduction to Atmospheric Radiation, 2nd edn. San Diego, CA: Academic Press.Google Scholar
Liston, G. E. (2004). Representing subgrid snow cover heterogeneities in regional and global models. Journal of Climate, 17, 13811397.Google Scholar
Liu, Y., and Gupta, H. V. (2007). Uncertainty in hydrologic modeling: Toward an integrated data assimilation framework. Water Resources Research, 43, W07401, doi:10.1029/2006WR005756.Google Scholar
Lloyd, C. R., Gash, J. H. C., and Shuttleworth, W. J. (1988). The measurement and modelling of rainfall interception by Amazonian rain forest. Agricultural and Forest Meteorology, 43, 277294.Google Scholar
Lloyd, J. (1991). Modelling stomatal responses to environment in Macadamia integrifolia. Australian Journal of Plant Physiology, 18, 649660.Google Scholar
Lloyd, J., and Farquhar, G. D. (2008). Effects of rising temperatures and [CO2] on the physiology of tropical forest trees. Philosophical Transactions of the Royal Society B, 363, 18111817.Google Scholar
Lloyd, J., Patiño, S., Paiva, R. Q., et al. (2010). Optimisation of photosynthetic carbon gain and within-canopy gradients of associated foliar traits for Amazon forest trees. Biogeosciences, 7, 18331859.Google Scholar
Lloyd, J., and Taylor, J. A. (1994). On the temperature dependence of soil respiration. Functional Ecology, 8, 315323.Google Scholar
Lombardozzi, D., Levis, S., Bonan, G., Hess, P. G., and Sparks, J. P. (2015b). The influence of chronic ozone exposure on global carbon and water cycles. Journal of Climate, 28, 292305.Google Scholar
Lombardozzi, D., Sparks, J. P., and Bonan, G. (2013). Integrating O3 influences on terrestrial processes: Photosynthetic and stomatal response data available for regional and global modeling. Biogeosciences, 10, 68156831.Google Scholar
Lombardozzi, D. L., Bonan, G. B., Smith, N. G., Dukes, J. S., and Fisher, R. A. (2015a). Temperature acclimation of photosynthesis and respiration: A key uncertainty in the carbon cycle-climate feedback. Geophysical Research Letters, 42, 86248631, doi:10.1002/2015GL065934.Google Scholar
Lombardozzi, D. L., Smith, N. G., Cheng, S. J., et al. (2018). Triose phosphate limitation in photosynthesis models reduces leaf photosynthesis and global terrestrial carbon storage. Environmental Research Letters, 13, 074025, doi:10.1088/1748-9326/aacf68.Google Scholar
Lombardozzi, D. L., Zeppel, M. J. B., Fisher, R. A., and Tawfik, A. (2017). Representing nighttime and minimum conductance in CLM4.5: Global hydrology and carbon sensitivity analysis using observational constraints. Geoscientific Model Development, 10, 321331.Google Scholar
Long, S. P., Postl, W. F., and Bolhàr-Nordenkampf, H. R. (1993). Quantum yields for uptake of carbon dioxide in C3 vascular plants of contrasting habitats and taxonomic groups. Planta, 189, 226234.Google Scholar
Long, T. A. (2005). The Forest and the Mainframe: The Dynamics of Modeling and Field Study in the Coniferous Forest Biome, 1969–1980, M.S. Thesis. Corvallis, OR: Oregon State University.Google Scholar
Louis, J.-F. (1979). A parametric model of vertical eddy fluxes in the atmosphere. Boundary Layer Meteorology, 17, 187202.Google Scholar
Louis, J. F., Tiedtke, M., and Geleyn, J. F. (1982). A short history of the operational PBL-parameterization at ECMWF. In Workshop on Planetary Boundary Layer Parameterization 25–27 November 1981. Reading, UK: European Centre for Medium Range Weather Forecasts, pp. 5979.Google Scholar
Lu, S., Ren, T., Gong, Y., and Horton, R. (2007). An improved model for predicting soil thermal conductivity from water content at room temperature. Soil Science Society of America Journal, 71, 814.Google Scholar
Lunardini, V. J. (1981). Heat Transfer in Cold Climates. New York, NY: Van Nostrand Reinhold.Google Scholar
Luo, Y., Shi, Z., Lu, X., et al. (2017). Transient dynamics of terrestrial carbon storage: Mathematical foundation and its applications. Biogeosciences, 14, 145161.Google Scholar
Luo, Y., White, L. W., Canadell, J. G., et al. (2003). Sustainability of terrestrial carbon sequestration: A case study in Duke Forest with inversion approach. Global Biogeochemical Cycles, 17, 1021, doi:10.1029/2002GB001923.Google Scholar
Luo, Y. Q., Randerson, J. T., Abramowitz, G., et al. (2012). A framework for benchmarking land models. Biogeosciences, 9, 38573874.Google Scholar
MacBean, N., Maignan, F., Peylin, P., et al. (2015). Using satellite data to improve the leaf phenology of a global terrestrial biosphere model. Biogeosciences, 12, 71857208.Google Scholar
Mahat, V., Tarboton, D. G., and Molotch, N. P. (2013). Testing above- and below-canopy representations of turbulent fluxes in an energy balance snowmelt model. Water Resources Research, 49, doi:10.1002/wrcr.20073.Google Scholar
Mahowald, N. M., Muhs, D. R., Levis, S., et al. (2006). Change in atmospheric mineral aerosols in response to climate: Last glacial period, preindustrial, modern, and doubled carbon dioxide climates. Journal of Geophysical Research, 111, D10202, doi:10.1029/2005JD006653.Google Scholar
Mahrt, L. (1987). Grid-averaged surface fluxes. Monthly Weather Review, 115, 15501560.Google Scholar
Majoube, M. (1971). Fractionnement en oxygène-18 et en deutérium entre l’eau et sa vapeur. Journal de Chimie Physique, 68, 14231436.Google Scholar
Makar, P. A., Fuentes, J. D., Wang, D., Staebler, R. M., and Wiebe, H. A. (1999). Chemical processing of biogenic hydrocarbons within and above a temperate deciduous forest. Journal of Geophysical Research, 104D, 35813603.Google Scholar
Makar, P. A., Staebler, R. M., Akingunola, A., et al. (2017). The effects of forest canopy shading and turbulence on boundary layer ozone. Nature Communications, 8, 15243, doi:10.1038/ncomms15243.Google Scholar
Manabe, S. (1969). Climate and the ocean circulation. I. The atmospheric circulation and the hydrology of the Earth’s surface. Monthly Weather Review, 97, 739774.Google Scholar
Manabe, S., Smagorinsky, J., and Strickler, R. F. (1965). Simulated climatology of a general circulation model with a hydrologic cycle. Monthly Weather Review, 93, 769798.Google Scholar
Mantoglou, A., and Gelhar, L. W. (1987a). Stochastic modeling of large-scale transient unsaturated flow systems. Water Resources Research, 23, 3746.Google Scholar
Mantoglou, A., and Gelhar, L. W. (1987b). Capillary tension head variance, mean soil moisture content, and effective specific soil moisture capacity of transient unsaturated flow in stratified soils. Water Resources Research, 23, 4756.Google Scholar
Mantoglou, A., and Gelhar, L. W. (1987c). Effective hydraulic conductivities of transient unsaturated flow in stratified soils. Water Resources Research, 23, 5767.Google Scholar
Manzoni, S., and Porporato, A. (2009). Soil carbon and nitrogen mineralization: Theory and models across scales. Soil Biology and Biochemistry, 41, 13551379.Google Scholar
Manzoni, S., Vico, G., Katul, G., et al. (2011). Optimizing stomatal conductance for maximum carbon gain under water stress: A meta-analysis across plant functional types and climates. Functional Ecology, 25, 456467.Google Scholar
Manzoni, S., Vico, G., Katul, G., (2013). Hydraulic limits on maximum plant transpiration and the emergence of the safety–efficiency trade-off. New Phytologist, 198, 169178.Google Scholar
Markesteijn, L., Poorter, L., Bongers, F., Paz, H., and Sack, L. (2011). Hydraulics and life history of tropical dry forest tree species: Coordination of species’ drought and shade tolerance. New Phytologist, 191, 480495.Google Scholar
Markkanen, T., Rannik, Ü., Marcolla, B., Cescatti, A., and Vesala, T. (2003). Footprints and fetches for fluxes over forest canopies with varying structure and density. Boundary-Layer Meteorology, 106, 437459.Google Scholar
Massman, W. J. (1982). Foliage distribution in old-growth coniferous tree canopies. Canadian Journal of Forest Research, 12, 1017.Google Scholar
Massman, W. J. (1997). An analytical one-dimensional model of momentum transfer by vegetation of arbitrary structure. Boundary-Layer Meteorology, 83, 407421.Google Scholar
Massman, W. J. (1998). A review of the molecular diffusivities of H2O, CO2, CH4, CO, O3, SO2, NH3, N2O, NO, and NO2 in air, O2 and N2 near STP. Atmospheric Environment, 32, 11111127.Google Scholar
Massman, W. J. (1999). A model study of kBH–1 for vegetated surfaces using ‘localized near-field’ Lagrangian theory. Journal of Hydrology, 223, 2743.Google Scholar
Massman, W. J., and Weil, J. C. (1999). An analytical one-dimensional second-order closure model of turbulence statistics and the Lagrangian time scale within and above plant canopies of arbitrary structure. Boundary-Layer Meteorology, 91, 81107.Google Scholar
Matheny, A. M., Fiorella, R. P., Bohrer, G., et al. (2017). Contrasting strategies of hydraulic control in two co-dominant temperate tree species. Ecohydrology, 10, e1815, doi:10.1002/eco.1815.Google Scholar
Maurer, E. P., Wood, A. W., Adam, J. C., Lettenmaier, D. P., and Nijssen, B. (2002). A long-term hydrologically based dataset of land surface fluxes and states for the conterminous United States. Journal of Climate, 15, 32373251.Google Scholar
McCaughey, J. H. (1985). Energy balance storage terms in a mature mixed forest at Petawawa, Ontario – a case study. Boundary-Layer Meteorology, 31, 89101.Google Scholar
McCaughey, J. H., and Saxton, W. L. (1988). Energy balance storage terms in a mixed forest. Agricultural and Forest Meteorology, 44, 118.Google Scholar
McGuire, A. D., Melillo, J. M., Joyce, L. A., et al. (1992). Interactions between carbon and nitrogen dynamics in estimating net primary productivity for potential vegetation in North America. Global Biogeochemical Cycles, 6, 101124.Google Scholar
McNaughton, K. G., and van den Hurk, B. J. J. M. (1995). A ‘Lagrangian’ revision of the resistors in the two-layer model for calculating the energy budget of a plant canopy. Boundary-Layer Meteorology, 74, 261288.Google Scholar
Meador, W. E., and Weaver, W. R. (1980). Two-stream approximations to radiative transfer in planetary atmospheres: A unified description of existing methods and a new improvement. Journal of the Atmospheric Sciences, 37, 630643.Google Scholar
Medlyn, B. E. (2004). A MAESTRO retrospective. In Forests at the Land–Atmosphere Interface, ed. Mencuccini, M., Grace, J., Moncrieff, J., and McNaughton, K. G.. Wallingford: CAB International, pp. 105121.Google Scholar
Medlyn, B. E., Barton, C. V. M., Broadmeadow, M. S. J., et al. (2001). Stomatal conductance of forest species after long-term exposure to elevated CO2 concentration: A synthesis. New Phytologist, 149, 247264.Google Scholar
Medlyn, B. E., Dreyer, E., Ellsworth, D., et al. (2002). Temperature response of parameters of a biochemically based model of photosynthesis. II. A review of experimental data. Plant, Cell and Environment, 25, 11671179.Google Scholar
Medlyn, B. E., Duursma, R. A., De Kauwe, M. G., and Prentice, I. C. (2013). The optimal stomatal response to atmospheric CO2 concentration: Alternative solutions, alternative interpretations. Agricultural and Forest Meteorology, 182/183, 200203.Google Scholar
Medlyn, B. E., Duursma, R. A., Eamus, D., et al. (2011). Reconciling the optimal and empirical approaches to modelling stomatal conductance. Global Change Biology, 17, 21342144.Google Scholar
Medlyn, B. E., McMurtrie, R. E., Dewar, R. C., and Jeffreys, M. P. (2000). Soil processes dominate the long-term response of forest net primary productivity to increased temperature and atmospheric CO2 concentration. Canadian Journal of Forest Research, 30, 873888.Google Scholar
Medvigy, D., Wofsy, S. C., Munger, J. W., Hollinger, D. Y., and Moorcroft, P. R. (2009). Mechanistic scaling of ecosystem function and dynamics in space and time: Ecosystem Demography model version 2. Journal of Geophysical Research, 114, G01002, doi:10.1029/2008JG000812.Google Scholar
Meentemeyer, V. (1978). Macroclimate and lignin control of litter decomposition rates. Ecology, 59, 465472.Google Scholar
Meinzer, F. C., Goldstein, G., Jackson, P., et al. (1995). Environmental and physiological regulation of transpiration in tropical forest gap species: The influence of boundary layer and hydraulic properties. Oecologia, 101, 514522.Google Scholar
Meir, P., Kruijt, B., Broadmeadow, M., et al. (2002). Acclimation of photosynthetic capacity to irradiance in tree canopies in relation to leaf nitrogen concentration and leaf mass per unit area. Plant, Cell and Environment, 25, 343357.Google Scholar
Melillo, J. M., McGuire, A. D., Kicklighter, D. W., et al. (1993). Global climate change and terrestrial net primary production. Nature, 363, 234240.Google Scholar
Melton, J. R., and Arora, V. K. (2014). Sub-grid scale representation of vegetation in global land surface schemes: Implications for estimation of the terrestrial carbon sink. Biogeosciences, 11, 10211036.Google Scholar
Melton, J. R., Wania, R., Hodson, E. L., et al. (2013). Present state of global wetland extent and wetland methane modelling: Conclusions from a model inter-comparison project (WETCHIMP). Biogeosciences, 10, 753788.Google Scholar
Merlivat, L. (1978). Molecular diffusivities of H216O, HD16O, and H218O in gases. Journal of Chemical Physics, 69, 28642871.Google Scholar
Meyerholt, J., and Zaehle, S. (2015). The role of stoichiometric flexibility in modelling forest ecosystem responses to nitrogen fertilization. New Phytologist, 208, 10421055.Google Scholar
Meyerholt, J., Zaehle, S., and Smith, M. J. (2016). Variability of projected terrestrial biosphere responses to elevated levels of atmospheric CO2 due to uncertainty in biological nitrogen fixation. Biogeosciences, 13, 14911518.Google Scholar
Meyers, T., and Paw U, K. T. (1986). Testing of a higher-order closure model for modeling airflow within and above plant canopies. Boundary-Layer Meteorology, 37, 297311.Google Scholar
Meyers, T. P., Finkelstein, P., Clarke, J., Ellestad, T. G., and Sims, P. F. (1998). A multilayer model for inferring dry deposition using standard meteorological measurements. Journal of Geophysical Research, 103D, 2264522661.Google Scholar
Meyers, T. P., and Hollinger, S. E. (2004). An assessment of storage terms in the surface energy balance of maize and soybean. Agricultural and Forest Meteorology, 125, 105115.Google Scholar
Michaletz, S. T., Weiser, M. D., McDowell, N. G., et al. (2016). The energetic and carbon economic origins of leaf thermoregulation. Nature Plants, 2, 16129, doi:10.1038/nplants.2016.129.Google Scholar
Michaletz, S. T., Weiser, M. D., Zhou, J., et al. (2015). Plant thermoregulation: Energetics, trait–environment interactions, and carbon economics. Trends in Ecology and Evolution, 30, 714724.Google Scholar
Michiles, A. A. S., and Gielow, R. (2008). Above-ground thermal energy storage rates, trunk heat fluxes and surface energy balance in a central Amazonian rainforest. Agricultural and Forest Meteorology, 148, 917930.Google Scholar
Miller, S. D., Goulden, M. L., Hutyra, L. R., et al. (2011). Reduced impact logging minimally alters tropical rainforest carbon and energy exchange. Proceedings of the National Academy of Sciences USA, 108, 1943119435.Google Scholar
Millington, R. J. (1959). Gas diffusion in porous media. Science, 130, 100102.Google Scholar
Milly, P. C. D. (1985). A mass-conservative procedure for time-stepping in models of unsaturated flow. Advances in Water Resources, 8, 3236.Google Scholar
Milly, P. C. D. (1988). Advances in modeling of water in the unsaturated zone. Transport in Porous Media, 3, 491514.Google Scholar
Milly, P. C. D., Malyshev, S. L., Shevliakova, E., et al. (2014). An enhanced model of land water and energy for global hydrologic and Earth-system studies. Journal of Hydrometeorology, 15, 17391761.Google Scholar
Milly, P. C. D., and Shmakin, A. B. (2002). Global modeling of land water and energy balances. Part I: The land dynamics (LaD) model. Journal of Hydrometeorology, 3, 283299.Google Scholar
Miner, G. L., Bauerle, W. L., and Baldocchi, D. D. (2017). Estimating the sensitivity of stomatal conductance to photosynthesis: A review. Plant, Cell and Environment, 40, 12141238.Google Scholar
Mintz, Y., and Walker, G. K. (1993). Global fields of soil moisture and land surface evapotranspiration derived from observed precipitation and surface air temperature. Journal of Applied Meteorology, 32, 13051334.Google Scholar
Miralles, D. G., Gash, J. H., Holmes, T. R. H., de Jeu, R. A. M., and Dolman, A. J. (2010). Global canopy interception from satellite observations. Journal of Geophysical Research, 115, D16122, doi:10.1029/2009JD013530.Google Scholar
Mirfenderesgi, G., Bohrer, G., Matheny, A. M., et al. (2016). Tree level hydrodynamic approach for resolving aboveground water storage and stomatal conductance and modeling the effects of tree hydraulic strategy. Journal of Geophysical Research: Biogeosciences, 121, 17921813, doi:10.1002/2016JG003467.Google Scholar
Mölder, M., Grelle, A., Lindroth, A., and Halldin, S. (1999). Flux-profile relationships over a boreal forest – roughness sublayer corrections. Agricultural and Forest Meteorology, 98/99, 645658.Google Scholar
Monsi, M., and Saeki, T. (1953). Über den Lichtfaktor in den Pflanzengesellschaften und seine Bedeutung für die Stoffproduktion. Japanese Journal of Botany, 14, 2252.Google Scholar
Monsi, M., and Saeki, T. (2005). On the factor light in plant communities and its importance for matter production. Annals of Botany, 95, 549567.Google Scholar
Monteith, J. L. (1965). Evaporation and environment. In The State and Movement of Water in Living Organisms (19th Symposia of the Society for Experimental Biology), ed. Fogg, G. E.. New York: Academic Press, pp. 205234.Google Scholar
Monteith, J. L. (1981a). Evaporation and surface temperature. Quarterly Journal of the Royal Meteorological Society, 107, 127.CrossRefGoogle Scholar
Monteith, J. L. (1981b). Coupling of plants to the atmosphere. In Plants and their Atmospheric Environment, ed. Grace, J., Ford, E. D., and Jarvis, P. G.. Oxford: Blackwell, pp. 129.Google Scholar
Monteith, J. L., and Unsworth, M. H. (2013). Principles of Environmental Physics, 4th edn. Amsterdam: Elsevier.Google Scholar
Moorcroft, P. R., Hurtt, G. C., and Pacala, S. W. (2001). A method for scaling vegetation dynamics: The ecosystem demography model (ED). Ecological Monographs, 71, 557586.Google Scholar
Moore, C. J., and Fisch, G. (1986). Estimating heat storage in Amazonian tropical forest. Agricultural and Forest Meteorology, 38, 147168.Google Scholar
Moore, R. J. (2007). The PDM rainfall-runoff model. Hydrology and Earth System Sciences, 11, 483499.Google Scholar
Morfopoulos, C., Prentice, I. C., Keenan, T. F., et al. (2013). A unifying conceptual model for the environmental responses of isoprene emissions from plants. Annals of Botany, 112, 12231238.Google Scholar
Morfopoulos, C., Sperlich, D., Peñuelas, J., et al. (2014). A model of plant isoprene emission based on available reducing power captures responses to atmospheric CO2. New Phytologist, 203, 125139.Google Scholar
Morison, J. I. L. (1987). Intercellular CO2 concentration and stomatal response to CO2. In Stomatal Function, ed. Zeiger, E., Farquhar, G. D., and Cowan, I. R.. Stanford, CA: Stanford University Press, pp. 229251.Google Scholar
Morison, J. I. L., and Jarvis, P. G. (1983). Direct and indirect effects of light on stomata. II. In Commelina communis L. Plant, Cell and Environment, 6, 103109.Google Scholar
Muzylo, A., Llorens, P., Valente, F., et al. (2009). A review of rainfall interception modelling. Journal of Hydrology, 370, 191206.Google Scholar
Nemitz, E., Milford, C., and Sutton, M. A. (2001). A two-layer canopy compensation point model for describing bi-directional biosphere–atmosphere exchange of ammonia. Quarterly Journal of the Royal Meteorological Society, 127, 815833.Google Scholar
Niinemets, Ü. (1998). Adjustment of foliage structure and function to a canopy light gradient in two co-existing deciduous trees. Variability in leaf inclination angles in relation to petiole morphology. Trees, 12, 446451.Google Scholar
Niinemets, Ü. (1999). Components of leaf dry mass per area – thickness and density – alter leaf photosynthetic capacity in reverse directions in woody plants. New Phytologist, 144, 3547.Google Scholar
Niinemets, Ü. (2007). Photosynthesis and resource distribution through plant canopies. Plant, Cell and Environment, 30, 10521071.Google Scholar
Niinemets, Ü. (2012). Optimization of foliage photosynthetic capacity in tree canopies: Towards identifying missing constraints. Tree Physiology, 32, 505509.Google Scholar
Niinemets, Ü., and Anten, N. P. R. (2009). Packing the photosynthetic machinery: From leaf to canopy. In Photosynthesis in silico: Understanding Complexity from Molecules to Ecosystems, ed. Laisk, A., Nedbal, L., and Govindjee, . Dordrecht: Springer, pp. 363399.Google Scholar
Niinemets, Ü., Bilger, W., Kull, O., and Tenhunen, J. D. (1999a). Responses of foliar photosynthetic electron transport, pigment stoichiometry, and stomatal conductance to interacting environmental factors in a mixed species forest canopy. Tree Physiology, 19, 839852.Google Scholar
Niinemets, Ü., Diaz-Espejo, A., Flexas, J., Galmés, J., and Warren, C. R. (2009). Importance of mesophyll diffusion conductance in estimation of plant photosynthesis in the field. Journal of Experimental Botany, 60, 22712282.Google Scholar
Niinemets, Ü., Keenan, T. F., and Hallik, L. (2015). A worldwide analysis of within-canopy variations in leaf structural, chemical and physiological traits across plant functional types. New Phytologist, 205, 973993.Google Scholar
Niinemets, Ü., and Kull, O. (1995). Effects of light availability and tree size on the architecture of assimilative surface in the canopy of Picea abies: Variation in needle morphology. Tree Physiology, 15, 307315.Google Scholar
Niinemets, Ü., Kull, O., and Tenhunen, J. D. (1998). An analysis of light effects on foliar morphology, physiology, and light interception in temperate deciduous woody species of contrasting shade tolerance. Tree Physiology, 18, 681696.Google Scholar
Niinemets, Ü., Kull, O., and Tenhunen, J. D. (2004). Within-canopy variation in the rate of development of photosynthetic capacity is proportional to integrated quantum flux density in temperate deciduous trees. Plant, Cell and Environment, 27, 293313.Google Scholar
Niinemets, Ü., and Tenhunen, J. D. (1997). A model separating leaf structural and physiological effects on carbon gain along light gradients for the shade-tolerant species Acer saccharum. Plant, Cell and Environment, 20, 845866.Google Scholar
Niinemets, Ü., Tenhunen, J. D., Harley, P. C., and Steinbrecher, R. (1999b). A model of isoprene emission based on energetic requirements for isoprene synthesis and leaf photosynthetic properties for Liquidambar and Quercus. Plant, Cell and Environment, 22, 13191335.Google Scholar
Nijssen, B., O’Donnell, G. M., Lettenmaier, D. P., Lohmann, D., and Wood, E. F. (2001). Predicting the discharge of global rivers. Journal of Climate, 14, 33073323.Google Scholar
Nilson, T. (1971). A theoretical analysis of the frequency of gaps in plant stands. Agricultural Meteorology, 8, 2538.Google Scholar
Niu, G.-Y., and Yang, Z.-L. (2004). Effects of vegetation canopy processes on snow surface energy and mass balances. Journal of Geophysical Research, 109, D23111, doi:10.1029/2004JD004884.Google Scholar
Niu, G.-Y., and Yang, Z.-L. (2006). Effects of frozen soil on snowmelt runoff and soil water storage at a continental scale. Journal of Hydrometeorology, 7, 937952.CrossRefGoogle Scholar
Niu, G.-Y., and Yang, Z.-L. (2007). An observation-based formulation of snow cover fraction and its evaluation over large North American river basins. Journal of Geophysical Research, 112, D21101, doi:10.1029/2007JD008674.Google Scholar
Niu, G.-Y., Yang, Z.-L., Dickinson, R. E., and Gulden, L. E. (2005). A simple TOPMODEL-based runoff parameterization (SIMTOP) for use in global climate models. Journal of Geophysical Research, 10, D21106, doi:10.1029/2005JD006111.Google Scholar
Niu, G.-Y., Yang, Z.-L., Mitchell, K. E., et al. (2011). The community Noah land surface model with multiparameterization options (Noah-MP): 1. Model description and evaluation with local-scale measurements. Journal of Geophysical Research, 116, D12109, doi:10.1029/2010JD015139.Google Scholar
Noilhan, J., and Planton, S. (1989). A simple parameterization of land surface processes for meteorological models. Monthly Weather Review, 117, 536549.Google Scholar
Norman, J. M. (1979). Modeling the complete crop canopy. In Modification of the Aerial Environment of Plants, ed. Barfield, B. J. and Gerber, J. F.. St. Joseph, MI: American Society of Agricultural Engineers, pp. 249277.Google Scholar
Norman, J. M. (1982). Simulation of microclimates. In Biometeorology in Integrated Pest Management, ed. Hatfield, J. L. and Thomason, I. J.. New York, NY: Academic Press, pp. 6599.Google Scholar
Norman, J. M. (1989). Synthesis of canopy processes. In Plant Canopies: Their Growth, Form and Function, ed. Russell, G., Marshall, B., and Jarvis, P. G.. Cambridge: Cambridge University Press, pp. 161175.Google Scholar
Norman, J. M. (1993). Scaling processes between leaf and canopy levels. In Scaling Physiological Processes: Leaf to Globe, ed. Ehleringer, J. R. and Field, C. B.. New York, NY: Academic Press, pp. 4176.Google Scholar
Norman, J. M., and Campbell, G. (1983). Application of a plant–environment model to problems in irrigation. In Advances in Irrigation, vol. 2, ed. Hillel, D.. New York, NY: Academic Press, pp. 155188.Google Scholar
Norman, J. M., Kustas, W. P., and Humes, K. S. (1995). Source approach for estimating soil and vegetation energy fluxes in observations of directional radiometric surface temperature. Agricultural and Forest Meteorology, 77, 263293.Google Scholar
O’Connell, P. E. (1991). A historical perspective. In Recent Advances in the Modeling of Hydrologic Systems, ed. Bowles, D. S. and O’Connell, P. E.. Dordrecht: Kluwer, pp. 330.Google Scholar
Odum, H. T. (1957). Trophic structure and productivity of Silver Springs, Florida. Ecological Monographs, 27, 55112.CrossRefGoogle Scholar
Odum, H. T. (1960). Ecological potential and analogue circuits for the ecosystem. American Scientist, 48, 18.Google Scholar
Ogée, J., Brunet, Y., Loustau, D., Berbigier, P., and Delzon, S. (2003). MuSICA, a CO2, water and energy multilayer, multileaf pine forest model: Evaluation from hourly to yearly time scales and sensitivity analysis. Global Change Biology, 9, 697717.Google Scholar
Oleson, K. W., Bonan, G. B., Feddema, J. J., Vertenstein, M., and Kluzek, E. (2010a). Technical Description of an Urban Parameterization for the Community Land Model (CLMU), Technical Note NCAR/TN-480+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Oleson, K. W., Dai, Y., Bonan, G., et al. (2004). Technical Description of the Community Land Model (CLM), Technical Note NCAR/TN-461+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Oleson, K. W., Lawrence, D. M., Bonan, G. B., et al. (2010b). Technical Description of Version 4.0 of the Community Land Model (CLM), Technical Note NCAR/TN-478+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Oleson, K. W., Lawrence, D. M., Bonan, G. B., et al. (2013). Technical Description of Version 4.5 of the Community Land Model (CLM), Technical Note NCAR/TN-503+STR. Boulder, CO: National Center for Atmospheric Research.Google Scholar
Oliphant, A. J., Grimmond, C. S. B., Zutter, H. N., et al. (2004). Heat storage and energy balance fluxes for a temperate deciduous forest. Agricultural and Forest Meteorology, 126, 185201.Google Scholar
Olson, J. S. (1963). Energy storage and the balance of producers and decomposers in ecological systems. Ecology, 44, 322331.Google Scholar
Olson, J. S. (1965). Equations for cesium transfer in a Liriodendron forest. Health Physics, 11, 13851392.Google Scholar
Oren, R., Sperry, J. S., Katul, G. G., et al. (1999). Survey and synthesis of intra- and interspecific variation in stomatal sensitivity to vapour pressure deficit. Plant, Cell and Environment, 22, 15151526.Google Scholar
Pacala, S. W., Canham, C. D., Saponara, J., et al. (1996). Forest models defined by field measurements: Estimation, error analysis and dynamics. Ecological Monographs, 66, 143.Google Scholar
Pacifico, F., Harrison, S. P., Jones, C. D., et al. (2011). Evaluation of a photosynthesis-based biogenic isoprene emission scheme in JULES and simulation of isoprene emissions under present-day climate conditions. Atmospheric Chemistry and Physics, 11, 43714389.Google Scholar
Paniconi, C., Aldama, A. A., and Wood, E. F. (1991). Numerical evaluation of iterative and noniterative methods for the solution of the nonlinear Richards equation. Water Resources Research, 27, 11471163.Google Scholar
Paniconi, C., and Putti, M. (1994). A comparison of Picard and Newton iteration in the numerical solution of multidimensional variably saturated flow problems. Water Resources Research, 30, 33573374.Google Scholar
Panofsky, H. A. (1963). Determination of stress from wind and temperature measurements. Quarterly Journal of the Royal Meteorological Society, 89, 8594.Google Scholar
Parkhurst, D. F., and Loucks, O. L. (1972). Optimal leaf size in relation to environment. Journal of Ecology, 60, 505537.Google Scholar
Parton, W. J., Hartman, M., Ojima, D., and Schimel, D. (1998). DAYCENT and its land surface submodel: Description and testing. Global and Planetary Change, 19, 3548.Google Scholar
Parton, W. J., Holland, E. A., Del Grosso, S. J., et al. (2001). Generalized model for NOx and N2O emissions from soils. Journal of Geophysical Research, 106D, 1740317419.CrossRefGoogle Scholar
Parton, W. J., Mosier, A. R., Ojima, D. S., et al. (1996). Generalized model for N2 and N2O production from nitrification and denitrification. Global Biogeochemical Cycles, 10, 401412.Google Scholar
Parton, W. J., Ojima, D. S., Cole, C. V., and Schimel, D. S. (1994). A general model for soil organic matter dynamics: Sensitivity to litter chemistry, texture and management. In Quantitative Modeling of Soil Forming Processes, ed. Bryant, R. B. and Arnold, R. W.. Madison, WI: Soil Science Society of America, pp. 147167.Google Scholar
Parton, W. J., Schimel, D. S., Cole, C. V., and Ojima, D. S. (1987). Analysis of factors controlling soil organic matter levels in Great Plains grasslands. Soil Science Society of America Journal, 51, 11731179.CrossRefGoogle Scholar
Parton, W. J., Scurlock, J. M. O., Ojima, D. S., et al. (1993). Observations and modeling of biomass and soil organic matter dynamics for the grassland biome worldwide. Global Biogeochemical Cycles, 7, 785809.Google Scholar
Parton, W., Silver, W. L., Burke, I. C., et al. (2007). Global-scale similarities in nitrogen release patterns during long-term decomposition. Science, 315, 361364.Google Scholar
Parton, W. J., Stewart, J. W. B., and Cole, C. V. (1988). Dynamics of C, N, P and S in grassland soils: A model. Biogeochemistry, 5, 109131.Google Scholar
Pastor, J., and Post, W. M. (1985). Development of a Linked Forest Productivity–Soil Process Model, ORNL/TM-9519. Oak Ridge, TN: Oak Ridge National Laboratory.Google Scholar
Pastor, J., and Post, W. M. (1986). Influence of climate, soil moisture, and succession on forest carbon and nitrogen cycles. Biogeochemistry, 2, 327.Google Scholar
Pastor, J., and Post, W. M. (1988). Response of northern forests to CO2-induced climate change. Nature, 334, 5558.Google Scholar
Pastor, J., and Post, W. M. (1996). LINKAGES – An individual-based forest ecosystem model. Climatic Change, 34, 253261.Google Scholar
Patten, B. C. (1975). Systems Analysis and Simulation in Ecology, vol. III. New York, NY: Academic Press.Google Scholar
Patton, E. G., Horst, T. W., Sullivan, P. P., et al. (2011). The Canopy Horizontal Array Turbulence Study. Bulletin of the American Meteorological Society, 92, 593611.Google Scholar
Patton, E. G., Sullivan, P. P., Shaw, R. H., Finnigan, J. J., and Weil, J. C. (2016). Atmospheric stability influences on coupled boundary layer and canopy turbulence. Journal of the Atmospheric Sciences, 73, 16211647.Google Scholar
Paulson, C. A. (1970). The mathematical representation of wind speed and temperature profiles in the unstable atmospheric surface layer. Journal of Applied Meteorology, 9, 857861.Google Scholar
Paustian, K., Parton, W. J., and Persson, J. (1992). Modeling soil organic matter in organic-amended and nitrogen-fertilized long-term plots. Soil Science Society of America Journal, 56, 476488.Google Scholar
Paw U, K. T. (1987). Mathematical analysis of the operative temperature and energy budget. Journal of Thermal Biology, 12, 227233.Google Scholar
Peltoniemi, M. S., Duursma, R. A., and Medlyn, B. E. (2012). Co-optimal distribution of leaf nitrogen and hydraulic conductance in plant canopies. Tree Physiology, 32, 510519.Google Scholar
Penman, H. L. (1940). Gas and vapour movements in the soil: I. The diffusion of vapours through porous solids. Journal of Agricultural Science, 30, 437462.Google Scholar
Penman, H. L. (1948). Natural evaporation from open water, bare soil and grass. Proceedings of the Royal Society London A, 193, 120145.Google Scholar
Pereira, F. L., Gash, J. H. C., David, J. S., and Valente, F. (2009). Evaporation of intercepted rainfall from isolated evergreen oak trees: Do the crowns behave as wet bulbs? Agricultural and Forest Meteorology, 149, 667679.Google Scholar
Peters-Lidard, C. D., Blackburn, E., Liang, X., and Wood, E. F. (1998). The effect of soil thermal conductivity parameterization on surface energy fluxes and temperatures. Journal of the Atmospheric Sciences, 55, 12091224.Google Scholar
Philip, J. R. (1957a). Evaporation, and moisture and heat fields in the soil. Journal of Meteorology, 14, 354366.Google Scholar
Philip, J. R. (1957b). The theory of infiltration: 4. Sorptivity and algebraic infiltration equations. Soil Science, 84, 257264.Google Scholar
Physick, W. L., and Garratt, J. R. (1995). Incorporation of a high-roughness lower boundary into a mesoscale model for studies of dry deposition over complex terrain. Boundary-Layer Meteorology, 74, 5571.Google Scholar
Pisek, J., Sonnentag, O., Richardson, A. D., and Mõttus, M. (2013). Is the spherical leaf inclination angle distribution a valid assumption for temperate and boreal broadleaf tree species? Agricultural and Forest Meteorology, 169, 186194.Google Scholar
Pitman, A. J., Henderson-Sellers, A., and Yang, Z.-L. (1990). Sensitivity of regional climates to localized precipitation in global models. Nature, 346, 734737.Google Scholar
Pleim, J., and Ran, L. (2011). Surface flux modeling for air quality applications. Atmosphere, 2, 271302.Google Scholar
Pleim, J. E., Xiu, A., Finkelstein, P. L., and Otte, T. L. (2001). A coupled land-surface and dry deposition model and comparison to field measurements of surface heat, moisture, and ozone fluxes. Water, Air, and Soil Pollution: Focus, 1, 243252.Google Scholar
Polcher, J., McAvaney, B., Viterbo, P., et al. (1998). A proposal for a general interface between land surface schemes and general circulation models. Global and Planetary Change, 19, 261276.Google Scholar
Pollard, D., and Thompson, S. L. (1995). Use of a land-surface-transfer scheme (LSX) in a global climate model: The response to doubling stomatal resistance. Global and Planetary Change, 10, 129161.Google Scholar
Poorter, H., Niinemets, Ü, Poorter, L., Wright, I. J., and Villar, R. (2009). Causes and consequences of variation in leaf mass per area (LMA): A meta-analysis. New Phytologist, 182, 565588.Google Scholar
Potter, C. S., Randerson, J. T., Field, C. B., et al. (1993). Terrestrial ecosystem production: A process model based on global satellite and surface data. Global Biogeochemical Cycles, 7, 811841.Google Scholar
Prentice, I. C., Bondeau, A., Cramer, W., et al. (2007). Dynamic Global Vegetation Modeling: Quantifying terrestrial ecosystem responses to large-scale environmental change. In Terrestrial Ecosystems in a Changing World, ed. Canadell, J. G., Pataki, D. E., and Pitelka, L. F.. Berlin: Springer, pp. 175192.Google Scholar
Prentice, I. C., Dong, N., Gleason, S. M., Maire, V., and Wright, I. J. (2014). Balancing the costs of carbon gain and water transport: Testing a new theoretical framework for plant functional ecology. Ecology Letters, 17, 8291.Google Scholar
Prentice, I. C., Liang, X., Medlyn, B. E., and Wang, Y.-P. (2015). Reliable, robust and realistic: The three R’s of next-generation land-surface modelling. Atmospheric Chemistry and Physics, 15, 59876005.Google Scholar
Prince, S. D., and Goward, S. N. (1995). Global primary production: A remote sensing approach. Journal of Biogeography, 22, 815835.Google Scholar
Pugh, T. A. M., Arneth, A., Olin, S., et al. (2015). Simulated carbon emissions from land-use change are substantially enhanced by accounting for agricultural management. Environmental Research Letters, 10, 124008, doi:10.1088/1748-9326/10/12/124008.Google Scholar
Purves, D. W., Lichstein, J. W., Strigul, N., and Pacala, S. W. (2008). Predicting and understanding forest dynamics using a simple tractable model. Proceedings of the National Academy of Sciences USA, 105, 1701817022.Google Scholar
Pyles, R. D., Weare, B. C., and Paw U, K. T. (2000). The UCD Advanced Canopy–Atmosphere–Soil Algorithm: Comparisons with observations from different climate and vegetation regimes. Quarterly Journal of the Royal Meteorological Society, 126, 29512980.Google Scholar
Qualls, R. J., and Brutsaert, W. (1996). Effect of vegetation density on the parameterization of scalar roughness to estimate spatially distributed sensible heat fluxes. Water Resources Research, 32, 645652.Google Scholar
Raats, P. A. C. (2007). Uptake of water from soils by plant roots. Transport in Porous Media, 68, 528.Google Scholar
Raczka, B., Duarte, H. F., Koven, C. D., et al. (2016). An observational constraint on stomatal function in forests: Evaluating coupled carbon and water vapor exchange with carbon isotopes in the Community Land Model (CLM4.5). Biogeosciences, 13, 51835204.Google Scholar
Raddatz, T. J., Reick, C. H., Knorr, W., et al. (2007). Will the tropical land biosphere dominate the climate–carbon cycle feedback during the twenty-first century? Climate Dynamics, 29, 565574.CrossRefGoogle Scholar
Rafique, R., Xia, J., Hararuk, O., et al. (2016). Divergent predictions of carbon storage between two global land models: Attribution of the causes through traceability analysis. Earth System Dynamics, 7, 649658.Google Scholar
Rafique, R., Xia, J., Hararuk, O., et al. (2017). Comparing the performance of three land models in global C cycle simulations: A detailed structural analysis. Land Degradation and Development, 28, 524533.Google Scholar
Raich, J. W., Rastetter, E. B., Melillo, J. M., et al. (1991). Potential net primary productivity in South America: Application of a global model. Ecological Applications, 1, 399429.Google Scholar
Rajaram, H., Bahr, J. M., Blöschl, G., et al. (2015). A reflection on the first 50 years of Water Resources Research. Water Resources Research, 51, 78297837, doi:10.1002/2015WR018089.Google Scholar
Ran, L., Pleim, J., Song, C., et al. (2017). A photosynthesis-based two-leaf canopy stomatal conductance model for meteorology and air quality modeling with WRF/CMAQ PX LSM. Journal of Geophysical Research: Atmospheres, 122, 19301952, doi:10.1002/2016JD025583.Google Scholar
Randerson, J. T., Hoffman, F. M., Thornton, P. E., et al. (2009). Systematic assessment of terrestrial biogeochemistry in coupled climate–carbon models. Global Change Biology, 15, 24622484.Google Scholar
Randerson, J. T., Thompson, M. V., Malmstrom, C. M., Field, C. B., and Fung, I. Y. (1996). Substrate limitations for heterotrophs: Implications for models that estimate the seasonal cycle of atmospheric CO2. Global Biogeochemical Cycles, 10, 585602.Google Scholar
Raschke, K. (1960). Heat transfer between the plant and the environment. Annual Review of Plant Physiology, 11, 111126.Google Scholar
Raupach, M. R. (1979). Anomalies in flux-gradient relationships over forest. Boundary Layer Meteorology, 16, 467486.Google Scholar
Raupach, M. R. (1988). Canopy transport processes. In Flow and Transport in the Natural Environment: Advances and Applications, ed. Steffen, W. L. and Denmead, O. T.. Berlin: Springer-Verlag, pp. 95127.Google Scholar
Raupach, M. R. (1989a). Applying Lagrangian fluid mechanics to infer scalar source distributions from concentration profiles in plant canopies. Agricultural and Forest Meteorology, 47, 85108.Google Scholar
Raupach, M. R. (1989b). A practical Lagrangian method for relating scalar concentrations to source distributions in vegetation canopies. Quarterly Journal of the Royal Meteorological Society, 115, 609632.Google Scholar
Raupach, M. R. (1989c). Stand overstorey processes. Philosophical Transactions of the Royal Society London B, 324, 175190.Google Scholar
Raupach, M. R. (1994). Simplified expressions for vegetation roughness length and zero-plane displacement as functions of canopy height and area index. Boundary-Layer Meteorology, 71, 211216.Google Scholar
Raupach, M. R. (1995). Corrigenda. Boundary-Layer Meteorology, 76, 303304.Google Scholar
Raupach, M. R., Finkele, K., and Zhang, L. (1997). SCAM (Soil-Canopy-Atmosphere Model): Description and Comparisons with Field Data, Technical Report No. 132. Canberra: CSIRO Centre for Environmental Mechanics.Google Scholar
Raupach, M. R., and Finnigan, J. J. (1988). ‘Single-layer models of evaporation from plant canopies are incorrect but useful, whereas multilayer models are correct but useless’: Discuss. Australian Journal of Plant Physiology, 15, 705716.Google Scholar
Raupach, M. R., Finnigan, J. J., and Brunet, Y. (1996). Coherent eddies and turbulence in vegetation canopies: The mixing-layer analogy. Boundary-Layer Meteorology, 78, 351382.Google Scholar
Rawls, W. J., Ahuja, L. R., Brakensiek, D. L., and Shirmohammadi, A. (1993). Infiltration and soil water movement. In Handbook of Hydrology, ed. Maidment, D. R.. New York, NY: McGraw-Hill, pp. 5.15.51.Google Scholar
Rawls, W. J., and Brakensiek, D. L. (1985). Prediction of soil water properties for hydrologic modeling. In Watershed Management in the Eighties, ed. Jones, E. B. and Ward, T. J.. New York, NY: American Society of Civil Engineers, pp. 293299.Google Scholar
Rawls, W. J., Brakensiek, D. L., and Saxton, K. E. (1982). Estimation of soil water properties. Transactions of the ASAE, 25, 13161320.Google Scholar
Reich, P. B., Ellsworth, D. S., and Walters, M. B. (1998). Leaf structure (specific leaf area) modulates photosynthesis–nitrogen relations: Evidence from within and across species and functional groups. Functional Ecology, 12, 948958.Google Scholar
Reick, C. H., Raddatz, T., Brovkin, V., and Gayler, V. (2013). Representation of natural and anthropogenic land cover change in MPI-ESM. Journal of Advances in Modeling Earth Systems, 5, 459482, doi:10.1002/jame.20022.Google Scholar
Richards, L. A. (1931). Capillary conduction of liquids through porous mediums. Physics, 1, 318333.Google Scholar
Richardson, A. D., Anderson, R. S., Arain, M. A., et al. (2012). Terrestrial biosphere models need better representation of vegetation phenology: Results from the North American Carbon Program Site Synthesis. Global Change Biology, 18, 566584.Google Scholar
Richtmyer, R. D., and Morton, K. W. (1967). Difference Methods for Initial-Value Problems, 2nd edn. New York, NY: Wiley.Google Scholar
Riddick, S., Ward, D., Hess, P., et al. (2016). Estimate of changes in agricultural terrestrial nitrogen pathways and ammonia emissions from 1850 to present in the Community Earth System Model. Biogeosciences, 13, 33973426.Google Scholar
Riley, W. J., Still, C. J., Torn, M. S., and Berry, J. A. (2002). A mechanistic model of H218O and C18OO fluxes between ecosystems and the atmosphere: Model description and sensitivity analyses. Global Biogeochemical Cycles, 16, 1095, doi:10.1029/2002GB001878.Google Scholar
Riley, W. J., Subin, Z. M., Lawrence, D. M., et al. (2011). Barriers to predicting changes in global terrestrial methane fluxes: Analyses using CLM4Me, a methane biogeochemistry model integrated in CESM. Biogeosciences, 8, 19251953.Google Scholar
Rogers, A. (2014). The use and misuse of Vc,max in Earth System Models. Photosynthesis Research, 119, 1529.Google Scholar
Rogers, A., Medlyn, B. E., Dukes, J. S., et al. (2017). A roadmap for improving the representation of photosynthesis in Earth system models. New Phytologist, 213, 2242.Google Scholar
Rosenzweig, C., and Abramopoulos, F. (1997). Land-surface model development for the GISS GCM. Journal of Climate, 10, 20402054.Google Scholar
Ross, J. (1975). Radiative transfer in plant communities. In Vegetation and the Atmosphere: vol. 1. Principles, ed. Monteith, J. L.. New York, NY: Academic Press, pp. 1355.Google Scholar
Ross, J. (1981). The Radiation Regime and Architecture of Plant Stands. The Hague: Dr. W. Junk.Google Scholar
Running, S. W., and Coughlan, J. C. (1988). A general model of forest ecosystem processes for regional applications. I. Hydrological balance, canopy gas exchange and primary production processes. Ecological Modelling, 42, 125154.Google Scholar
Running, S. W., and Gower, S. T. (1991). FOREST-BGC, a general model of forest ecosystem processes for regional applications. II. Dynamic carbon allocation and nitrogen budgets. Tree Physiology, 9, 147160.Google Scholar
Running, S. W., and Hunt, E. R., Jr. (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. R. and Field, C. B.. New York, NY: Academic Press, pp. 141158.Google Scholar
Running, S. W., Nemani, R. R., Heinsch, F. A., et al. (2004). A continuous satellite-derived measure of global terrestrial primary production. BioScience, 54, 547560.Google Scholar
Running, S. W., Thornton, P. E., Nemani, R., and Glassy, J. M. (2000). Global terrestrial gross and net primary productivity from the Earth Observing System. In Methods in Ecosystem Science, ed. Sala, O. E.. New York, NY: Springer-Verlag, pp. 4457.Google Scholar
Running, S. W., Waring, R. H., and Rydell, R. A. (1975). Physiological control of water flux in conifers: A computer simulation model. Oecologia, 18, 116.Google Scholar
Rutter, A. J. (1967). An analysis of evaporation from a stand of Scots pine. In Forest Hydrology: Proceedings of a National Science Foundation Advanced Science Seminar held at The Pennsylvania State University, University Park, Pennsylvania, Aug 29–Sept 10, 1965, ed. Sopper, W. E. and Lull, H. W.. Oxford: Pergamon Press, pp. 403417.Google Scholar
Rutter, A. J., Kershaw, K. A., Robins, P. C., and Morton, A. J. (1971). A predictive model of rainfall interception in forests, 1. Derivation of the model from observations in a plantation of Corsican pine. Agricultural Meteorology, 9, 367384.Google Scholar
Rutter, A. J., Morton, A. J., and Robins, P. C. (1975). A predictive model of rainfall interception in forests. II. Generalization of the model and comparison with observations in some coniferous and hardwood stands. Journal of Applied Ecology, 12, 367380.Google Scholar
Ryan, M. G. (1991). Effects of climate change on plant respiration. Ecological Applications, 1, 157167.Google Scholar
Ryder, J., Polcher, J., Peylin, P., et al. (2016). A multi-layer land surface energy budget model for implicit coupling with global atmospheric simulations. Geoscientific Model Development, 9, 223245.Google Scholar
Ryel, R. J., Caldwell, M. M., Yoder, C. K., Or, D., Leffler, A. J. (2002). Hydraulic redistribution in a stand of Artemisia tridentata: Evaluation of benefits to transpiration assessed with a simulation model. Oecologia, 130, 173184.Google Scholar
Ryu, Y., Baldocchi, D. D., Kobayashi, H., et al. (2011). Integration of MODIS land and atmosphere products with a coupled-process model to estimate gross primary productivity and evapotranspiration from 1 km to global scales. Global Biogeochemical Cycles, 25, GB4017, doi:10.1029/2011GB004053.Google Scholar
Saikawa, E., Schlosser, C. A., and Prinn, R. G. (2013). Global modeling of soil nitrous oxide emissions from natural processes. Global Biogeochemical Cycles, 27, 972989, doi:10.1002/gbc.20087.Google Scholar
Sakaguchi, K., and Zeng, X. (2009). Effects of soil wetness, plant litter, and under-canopy atmospheric stability on ground evaporation in the Community Land Model (CLM3.5). Journal of Geophysical Research, 114, D01107, doi:10.1029/2008JD010834.Google Scholar
Sands, P. J. (1995). Modelling canopy production. II. From single-leaf photosynthetic parameters to daily canopy photosynthesis. Australian Journal of Plant Physiology, 22, 603614.Google Scholar
Sato, H., Itoh, A., and Kohyama, T. (2007). SEIB–DGVM: A new Dynamic Global Vegetation Model using a spatially explicit individual-based approach. Ecological Modelling, 200, 279307.Google Scholar
Saylor, R. D. (2013). The Atmospheric Chemistry and Canopy Exchange Simulation System (ACCESS): Model description and application to a temperate deciduous forest canopy. Atmospheric Chemistry and Physics, 13, 693715.Google Scholar
Saylor, R. D., Wolfe, G. M., Meyers, T. P., and Hicks, B. B. (2014). A corrected formulation of the Multilayer Model (MLM) for inferring gaseous dry deposition to vegetated surfaces. Atmospheric Environment, 92, 141145.Google Scholar
Schaap, M. G., Leij, F. J., and van Genuchten, M. T. (1998). Neural network analysis for hierarchical prediction of soil hydraulic properties. Soil Science Society of America Journal, 62, 847855.Google Scholar
Schaap, M. G., Leij, F. J., and van Genuchten, M. T. (2001). ROSETTA: A computer program for estimating soil hydraulic parameters with hierarchical pedotransfer functions. Journal of Hydrology, 251, 163176.Google Scholar
Schaphoff, S., von Bloh, W., Rammig, A., et al. (2018). LPJmL4 – a dynamic global vegetation model with managed land – Part 1: Model description. Geoscientific Model Development, 11, 13431375.Google Scholar
Schieving, F., Werger, M. J. A., and Hirose, T. (1992). Canopy structure, nitrogen distribution and whole canopy photosynthetic carbon gain in growing and flowering stands of tall herbs. Vegetatio, 102, 173181.Google Scholar
Schimel, D. S., Kittel, T. G. F., Knapp, A. K., et al. (1991). Physiological interactions along resource gradients in a tallgrass prairie. Ecology, 72, 672684.Google Scholar
Schmidt, G. A., Kelley, M., Nazarenko, L., et al. (2014). Configuration and assessment of the GISS ModelE2 contributions to the CMIP5 archive. Journal of Advances in Modeling Earth Systems, 6, 141184, doi:10.1002/2013MS000265.CrossRefGoogle Scholar
Schuepp, P. H. (1993). Leaf boundary layers. New Phytologist, 125, 477507.Google Scholar
Schultz, M. G., Heil, A., Hoelzemann, J. J., et al. (2008). Global wildland fire emissions from 1960 to 2000. Global Biogeochemical Cycles, 22, GB2002, doi:10.1029/2007GB003031.CrossRefGoogle Scholar
Schultz, N. M., Lee, X., Lawrence, P. J., Lawrence, D. M., and Zhao, L. (2016). Assessing the use of subgrid land model output to study impacts of land cover change. Journal of Geophysical Research: Atmospheres, 121, 61336147, doi:10.1002/2016JD025094.Google Scholar
Schwarz, P. A., Law, B. E., Williams, M., et al. (2004). Climatic versus biotic constraints on carbon and water fluxes in seasonally drought-affected ponderosa pine ecosystems. Global Biogeochemical Cycles, 18, GB4007, doi:10.1029/2004GB002234.Google Scholar
Schymanski, S. J., Or, D., and Zwieniecki, M. (2013). Stomatal control and leaf thermal and hydraulic capacitances under rapid environmental fluctuations. PLoS ONE, 8(1), e54231, doi:10.1371/journal.pone.0054231.Google Scholar
Séférian, R., Delire, C., Decharme, B., et al. (2016). Development and evaluation of CNRM Earth system model – CNRM-ESM1. Geoscientific Model Development, 9, 14231453.Google Scholar
Sellers, P. J. (1985). Canopy reflectance, photosynthesis and transpiration. International Journal of Remote Sensing, 6, 13351372.Google Scholar
Sellers, P. J., Berry, J. A., Collatz, G. J., Field, C. B., and Hall, F. G. (1992). Canopy reflectance, photosynthesis, and transpiration. III. A reanalysis using improved leaf models and a new canopy integration scheme. Remote Sensing of Environment, 42, 187216.Google Scholar
Sellers, P. J., Fennessy, M. J., and Dickinson, R. E. (2007). A numerical approach to calculating soil wetness and evapotranspiration over large grid areas. Journal of Geophysical Research, 112, D18106, doi:10.1029/2007JD008781.Google Scholar
Sellers, P. J., and Lockwood, J. G. (1981). A computer simulation of the effects of differing crop types on the water balance of small catchments over long time periods. Quarterly Journal of the Royal Meteorological Society, 107, 395414.Google Scholar
Sellers, P. J., Los, S. O., Tucker, C. J., et al. (1996b). A revised land surface parameterization (SiB2) for atmospheric GCMs. Part II: The generation of global fields of terrestrial biophysical parameters from satellite data. Journal of Climate, 9, 706737.Google Scholar
Sellers, P. J., Mintz, Y., Sud, Y. C., and Dalcher, A. (1986). A simple biosphere model (SiB) for use within general circulation models. Journal of the Atmospheric Sciences, 43, 505531.Google Scholar
Sellers, P. J., Randall, D. A., Collatz, G. J., et al. (1996a). A revised land surface parameterization (SiB2) for atmospheric GCMs. Part I: Model formulation. Journal of Climate, 9, 676705.Google Scholar
Seth, A., Giorgi, F., and Dickinson, R. E. (1994). Simulating fluxes from heterogeneous land surfaces: Explicit subgrid method employing the biosphere–atmosphere transfer scheme (BATS). Journal of Geophysical Research, 99D, 1865118667.Google Scholar
Sharkey, T. D., Bernacchi, C. J., Farquhar, G. D., and Singsaas, E. L. (2007). Fitting photosynthetic carbon dioxide response curves for C3 leaves. Plant, Cell and Environment, 30, 10351040.Google Scholar
Shaw, R. H., and Pereira, A. R. (1982). Aerodynamic roughness of a plant canopy: A numerical experiment. Agricultural Meteorology, 26, 5165.Google Scholar
Shevliakova, E., Pacala, S. W., Malyshev, S., et al. (2009). Carbon cycling under 300 years of land use change: Importance of the secondary vegetation sink. Global Biogeochemical Cycles, 23, GB2022, doi:10.1029/2007GB003176.Google Scholar
Shi, M., Fisher, J. B., Brzostek, E. R., and Phillips, R. P. (2016). Carbon cost of plant nitrogen acquisition: Global carbon cycle impact from an improved plant nitrogen cycle in the Community Land Model. Global Change Biology, 22, 12991314.Google Scholar
Shinozaki, K., Yoda, K., Hozumi, K., and Kira, T. (1964). A quantitative analysis of plant form – the pipe model theory. I. Basic analyses. Japanese Journal of Ecology, 14, 97105.Google Scholar
Shiraiwa, T., and Sinclair, T. R. (1993). Distribution of nitrogen among leaves in soybean canopies. Crop Science, 33, 804808.Google Scholar
Shugart, H. H. (1984). A Theory of Forest Dynamics: The Ecological Implications of Forest Succession Models. New York, NY: Springer-Verlag.Google Scholar
Shugart, H. H., Crow, T. R., and Jett, J. M. (1973). Forest succession models: A rationale and methodology for modeling forest succession over large regions. Forest Science, 19, 203212.Google Scholar
Shugart, H. H., and West, D. C. (1977). Development of an Appalachian deciduous forest succession model and its application to assessment of the impact of the chestnut blight. Journal of Environmental Management, 5, 161179.Google Scholar
Shugart, H. H., and Woodward, F. I. (2011). Global Change and the Terrestrial Biosphere: Achievements and Challenges. Chichester: Wiley-Blackwell.Google Scholar
Shuman, J. K., Shugart, H. H., and Krankina, O. N. (2013). Assessment of carbon stores in tree biomass for two management scenarios in Russia. Environmental Research Letters, 8, 045019, doi:10.1088/1748-9326/8/4/045019.Google Scholar
Shuman, J. K., Shugart, H. H., and Krankina, O. N. (2014). Testing individual-based models of forest dynamics: Issues and an example from the boreal forests of Russia. Ecological Modelling, 293, 102110.Google Scholar
Shuman, J. K., Tchebakova, N. M., Parfenova, E. I., et al. (2015). Forest forecasting with vegetation models across Russia. Canadian Journal of Forest Research, 45, 175184.Google Scholar
Shuttleworth, W. J. (1976). A one-dimensional theoretical description of the vegetation-atmosphere interaction. Boundary-Layer Meteorology, 10, 273302.Google Scholar
Shuttleworth, W. J. (1978). A simplified one-dimensional theoretical description of the vegetation-atmosphere interaction. Boundary-Layer Meteorology, 14, 327.Google Scholar
Shuttleworth, W. J. (1988). Macrohydrology – the new challenge for process hydrology. Journal of Hydrology, 100, 3156.Google Scholar
Shuttleworth, W. J. (2012). Terrestrial Hydrometeorology. Chichester: Wiley-Blackwell.Google Scholar
Shuttleworth, W. J., and Gurney, R. J. (1990). The theoretical relationship between foliage temperature and canopy resistance in sparse canopies. Quarterly Journal of the Royal Meteorological Society, 116, 497519.Google Scholar
Shuttleworth, W. J., and Wallace, J. S. (1985). Evaporation from sparse crops – an energy combination theory. Quarterly Journal of the Royal Meteorological Society, 111, 839855.Google Scholar
Sierra, C. A., and Müller, M. (2015). A general mathematical framework for representing soil organic matter dynamics. Ecological Monographs, 85, 505524.Google Scholar
Sierra, C. A., Müller, M., and Trumbore, S. E. (2012). Models of soil organic matter decomposition: the SOILR package, version 1.0. Geoscientific Model Development, 5, 10451060.Google Scholar
Sinclair, T. R., Murphy, C. E., and Knoerr, K. R. (1976). Development and evaluation of simplified models for simulating canopy photosynthesis and transpiration. Journal of Applied Ecology, 13, 813829.Google Scholar
Siqueira, M. B., and Katul, G. G. (2010). An analytical model for the distribution of CO2 sources and sinks, fluxes, and mean concentration within the roughness sub-layer. Boundary-Layer Meteorology, 135, 3150.Google Scholar
Sitch, S., Smith, B., Prentice, I. C., et al. (2003). Evaluation of ecosystem dynamics, plant geography and terrestrial carbon cycling in the LPJ dynamic global vegetation model. Global Change Biology, 9, 161185.Google Scholar
Sitch, S., Cox, P. M., Collins, W. J., and Huntingford, C. (2007). Indirect radiative forcing of climate change through ozone effects on the land-carbon sink. Nature, 448, 791794.Google Scholar
Sivapalan, M., Beven, K., and Wood, E. F. (1987). On hydrologic similarity: 2. A scaled model of storm runoff production. Water Resources Research, 23, 22662278.Google Scholar
Skillman, J. B. (2008). Quantum yield variation across the three pathways of photosynthesis: Not yet out of the dark. Journal of Experimental Botany, 59, 16471661.Google Scholar
Smith, B., Prentice, I. C., and Sykes, M. T. (2001). Representation of vegetation dynamics in the modelling of terrestrial ecosystems: Comparing two contrasting approaches within European climate space. Global Ecology and Biogeography, 10, 621637.Google Scholar
Smith, B., Wårlind, D., Arneth, A., et al. (2014). Implications of incorporating N cycling and N limitations on primary production in an individual-based dynamic vegetation model. Biogeosciences, 11, 20272054.Google Scholar
Smith, N. G., and Dukes, J. S. (2013). Plant respiration and photosynthesis in global-scale models: Incorporating acclimation to temperature and CO2. Global Change Biology, 19, 4563.Google Scholar
Smith, N. G., Malyshev, S. L., Shevliakova, E., et al. (2016). Foliar temperature acclimation reduces simulated carbon sensitivity to climate. Nature Climate Change, 6, 407411.Google Scholar
Smits, K. M., Ngo, V. V., Cihan, A., Sakaki, T., and Illangasekare, T. H. (2012). An evaluation of models of bare soil evaporation formulated with different land surface boundary conditions and assumptions. Water Resources Research, 48, W12526, doi:10.1029/2012WR012113.Google Scholar
Sokolov, A. P., Kicklighter, D. W., Melillo, J. M., et al. (2008). Consequences of considering carbon–nitrogen interactions on the feedbacks between climate and the terrestrial carbon cycle. Journal of Climate, 21, 37763796.Google Scholar
Sollins, P., Brown, A. T., and Swartzman, G. L. (1979). CONIFER: A Model of Carbon and Water Flow through a Coniferous Forest – Revised Documentation, Coniferous Forest Biome Bulletin No. 15. Seattle, WA: University of Washington.Google Scholar
Solomon, A. M. (1986). Transient response of forests to CO2-induced climate change: Simulation modeling experiments in eastern North America. Oecologia, 68, 567579.Google Scholar
Solomon, A. M., Delcourt, H. R., West, D. C., and Blasing, T. J. (1980). Testing a simulation model for reconstruction of prehistoric forest-stand dynamics. Quaternary Research, 14, 275293.Google Scholar
Solomon, A. M., West, D. C., and Solomon, J. A. (1981). Simulating the role of climate change and species immigration in forest succession. In Forest Succession: Concepts and Application, ed. West, D. C., Shugart, H. H., and Botkin, D. B.. New York, NY: Springer-Verlag, pp. 154177.Google Scholar
Sperry, J. S., Adler, F. R., Campbell, G. S., and Comstock, J. P. (1998). Limitation of plant water use by rhizosphere and xylem conductance: Results from a model. Plant, Cell and Environment, 21, 347359.Google Scholar
Sperry, J. S., Venturas, M. D., Anderegg, W. R. L., et al. (2017). Predicting stomatal responses to the environment from the optimization of photosynthetic gain and hydraulic cost. Plant, Cell and Environment, 40, 816830.Google Scholar
Spitters, C. J. T. (1986). Separating the diffuse and direct component of global radiation and its implications for modeling canopy photosynthesis. Part II. Calculation of canopy photosynthesis. Agricultural and Forest Meteorology, 38, 231242.Google Scholar
Stamm, J. F., Wood, E. F., and Lettenmaier, D. P. (1994). Sensitivity of a GCM simulation of global climate to the representation of land-surface hydrology. Journal of Climate, 7, 12181239.Google Scholar
Staudt, K., Falge, E., Pyles, R. D., Paw U, K. T., and Foken, T. (2010). Sensitivity and predictive uncertainty of the ACASA model at a spruce forest site. Biogeosciences, 7, 36853705.Google Scholar
Staudt, K., Serafimovich, A., Siebicke, L., Pyles, R. D., and Falge, E. (2011). Vertical structure of evapotranspiration at a forest site (a case study). Agricultural and Forest Meteorology, 151, 709729.Google Scholar
Steele, M. J., Coutts, M. P., and Yeoman, M. M. (1989). Developmental changes in Sitka spruce as indices of physiological age I. Changes in needle morphology. New Phytologist, 113, 367375.Google Scholar
Stewart, J. B. (1977). Evaporation from the wet canopy of a pine forest. Water Resources Research, 13, 915921.Google Scholar
Stieglitz, M., Rind, D., Famiglietti, J., and Rosenzweig, C. (1997). An efficient approach to modeling the topographic control of surface hydrology for regional and global climate modeling. Journal of Climate, 10, 118137.Google Scholar
Stöckli, R., Rutishauser, T., Dragoni, D., et al. (2008). Remote sensing data assimilation for a prognostic phenology model. Journal of Geophysical Research, 113, G04021, doi:10.1029/2008JG000781.Google Scholar
Stöckli, R., Rutishauser, T., Baker, I., Liniger, M. A., and Denning, A. S. (2011). A global reanalysis of vegetation phenology. Journal of Geophysical Research, 116, G03020, doi:10.1029/2010JG001545.Google Scholar
Stoy, P. C., Katul, G. G., Siqueira, M. B., et al. (2006). Separating the effects of climate and vegetation on evapotranspiration along a successional chronosequence in the southeastern US. Global Change Biology, 12, 21152135.Google Scholar
Stoy, P. C., Mauder, M., Foken, T., et al. (2013). A data-driven analysis of energy balance closure across FLUXNET research sites: The role of landscape scale heterogeneity. Agricultural and Forest Meteorology, 171/172, 137152.Google Scholar
Strigul, N., Pristinski, D., Purves, D., Dushoff, J., and Pacala, S. (2008). Scaling from trees to forests: Tractable macroscopic equations for forest dynamics. Ecological Monographs, 78, 523545.Google Scholar
Stroud, C., Makar, P., Karl, T., et al. (2005). Role of canopy-scale photochemistry in modifying biogenic-atmosphere exchange of reactive terpene species: Results from the CELTIC field study. Journal of Geophysical Research, 110, D17303, doi:10.1029/2005JD005775.Google Scholar
Su, H.-B., Paw U, K. T., and Shaw, R. H. (1996). Development of a coupled leaf and canopy model for the simulation of plant–atmosphere interaction. Journal of Applied Meteorology, 35, 733748.Google Scholar
Suits, N. S., Denning, A. S., Berry, J. A., et al. (2005). Simulation of carbon isotope discrimination of the terrestrial biosphere. Global Biogeochemical Cycles, 19, GB1017, doi:10.1029/2003GB002141.Google Scholar
Sun, Y., Gu, L., Dickinson, R. E., et al. (2014a). Asymmetrical effects of mesophyll conductance on fundamental photosynthetic parameters and their relationships estimated from leaf gas exchange measurements. Plant, Cell and Environment, 37, 978994.Google Scholar
Sun, Y., Gu, L., Dickinson, R. E., (2014b). Impact of mesophyll diffusion on estimated global land CO2 fertilization. Proceedings of the National Academy of Sciences USA, 111, 1577415779.Google Scholar
Sutton, M. A., Reis, S., Riddick, S. N., et al. (2013).Towards a climate-dependent paradigm of ammonia emission and deposition. Philosophical Transactions of the Royal Society B, 368, 20130166, doi:10.1098/rstb.2013.0166.Google Scholar
Sutton, M. A., Schjørring, J. K., and Wyers, G. P. (1995). Plant–atmosphere exchange of ammonia. Philosophical Transactions of the Royal Society London A, 351, 261278.Google Scholar
Swenson, S. C., and Lawrence, D. M. (2012). A new fractional snow-covered area parameterization for the Community Land Model and its effect on the surface energy balance. Journal of Geophysical Research, 117, D21107, doi:10.1029/2012JD018178.Google Scholar
Swenson, S. C., and Lawrence, D. M. (2014). Assessing a dry surface layer-based soil resistance parameterization for the Community Land Model using GRACE and FLUXNET-MTE data. Journal of Geophysical Research: Atmospheres, 119, 1029910312, doi:10.1002/2014JD022314.Google Scholar
Tague, C. L., and Band, L. E. (2004). RHESSys: Regional Hydro-Ecologic Simulation System – an object-oriented approach to spatially distributed modeling of carbon, water, and nutrient cycling. Earth Interactions, 8(19), 142.Google Scholar
Takata, K., Emori, S., and Watanabe, T. (2003). Development of the minimal advanced treatments of surface interaction and runoff. Global and Planetary Change, 38, 209222.Google Scholar
Tang, J. Y., and Riley, W. J. (2013). A new top boundary condition for modeling surface diffusive exchange of a generic volatile tracer: Theoretical analysis and application to soil evaporation. Hydrology and Earth System Sciences, 17, 873893.Google Scholar
Ter-Mikaelian, M. T., and Korzukhin, M. D. (1997). Biomass equations for sixty-five North American tree species. Forest Ecology and Management, 97, 124.Google Scholar
Thom, A. S. (1971). Momentum absorption by vegetation. Quarterly Journal of the Royal Meteorological Society, 97, 414428.Google Scholar
Thom, A. S. (1972). Momentum, mass and heat exchange of vegetation. Quarterly Journal of the Royal Meteorological Society, 98, 124134.Google Scholar
Thom, A. S. (1975). Momentum, mass and heat exchange of plant communities. In Vegetation and the Atmosphere: vol. 1. Principles, ed. Monteith, J. L.. New York, NY: Academic Press, pp. 57109.Google Scholar
Thomas, R. Q., Bonan, G. B., and Goodale, C. L. (2013a). Insights into mechanisms governing forest carbon response to nitrogen deposition: A model–data comparison using observed responses to nitrogen addition. Biogeosciences, 10, 38693887.Google Scholar
Thomas, R. Q., and Williams, M. (2014). A model using marginal efficiency of investment to analyze carbon and nitrogen interactions in terrestrial ecosystems (ACONITE Version 1). Geoscientific Model Development, 7, 20152037.Google Scholar
Thomas, R. Q., Zaehle, S., Templer, P. H., and Goodale, C. L. (2013b). Global patterns of nitrogen limitation: Confronting two global biogeochemical models with observations. Global Change Biology, 19, 29862998.Google Scholar
Thomas, S. C., and Winner, W. E. (2000). A rotated ellipsoidal angle density function improves estimation of foliage inclination distributions in forest canopies. Agricultural and Forest Meteorology, 100, 1924.Google Scholar
Thomsen, J. E., Bohrer, G., Matheny, A. M., et al. (2013). Contrasting hydraulic strategies during dry soil conditions in Quercus rubra and Acer rubrum in a sandy site in Michigan. Forests, 4, 11061120.Google Scholar
Thonicke, K., Spessa, A., Prentice, I. C., et al. (2010). The influence of vegetation, fire spread and fire behaviour on biomass burning and trace gas emissions: Results from a process-based model. Biogeosciences, 7, 19912011.Google Scholar
Thonicke, K., Venevsky, S., Sitch, S., and Cramer, W. (2001). The role of fire disturbance for global vegetation dynamics: Coupling fire into a Dynamic Global Vegetation Model. Global Ecology and Biogeography, 10, 661677.Google Scholar
Thornley, J. H. M. (2002). Instantaneous canopy photosynthesis: Analytical expressions for sun and shade leaves based on exponential light decay down the canopy and an acclimated non-rectangular hyperbola for leaf photosynthesis. Annals of Botany, 89, 451458.Google Scholar
Thornton, P. E., Doney, S. C., Lindsay, K., et al. (2009). Carbon–nitrogen interactions regulate climate–carbon cycle feedbacks: Results from an atmosphere–ocean general circulation model. Biogeosciences620992120.Google Scholar
Thornton, P. E., Lamarque, J.-F., Rosenbloom, N. A., and Mahowald, N. M. (2007). Influence of carbon–nitrogen cycle coupling on land model response to CO2 fertilization and climate variability. Global Biogeochemical Cycles, 21, GB4018, doi:10.1029/2006GB002868.Google Scholar
Thornton, P. E., Law, B. E., Gholz, H. L., et al. (2002). Modeling and measuring the effects of disturbance history and climate on carbon and water budgets in evergreen needleleaf forests. Agricultural and Forest Meteorology, 113, 185222.Google Scholar
Tjoelker, M. G., Oleksyn, J., and Reich, P. B. (2001). Modelling respiration of vegetation: Evidence for a general temperature-dependent Q10. Global Change Biology, 7, 223230.Google Scholar
Todini, E. (2007). Hydrological catchment modelling: Past, present and future. Hydrology and Earth System Sciences, 11, 468482.Google Scholar
Toon, O. B., McKay, C. P., Ackerman, T. P., and Santhanam, K. (1989). Rapid calculation of radiative heating rates and photodissociation rates in inhomogeneous multiple scattering atmospheres. Journal of Geophysical Research, 94D, 1628716301.Google Scholar
Trofymow, J. A., Moore, T. R., Titus, B., et al. (2002). Rates of litter decomposition over 6 years in Canadian forests: Influence of litter quality and climate. Canadian Journal of Forest Research, 32, 789804.Google Scholar
Tuzet, A., Perrier, A., and Leuning, R. (2003). A coupled model of stomatal conductance, photosynthesis and transpiration. Plant, Cell and Environment, 26, 10971116.Google Scholar
Twine, T. E., Kustas, W. P., Norman, J. M., et al. (2000). Correcting eddy-covariance flux underestimates over a grassland. Agricultural and Forest Meteorology, 103, 279300.Google Scholar
Tyree, M. T. (1988). A dynamic model for water flow in a single tree: Evidence that models must account for hydraulic architecture. Tree Physiology, 4, 195217.Google Scholar
Tyree, M. T., and Ewers, F. W. (1991). The hydraulic architecture of trees and other woody plants. New Phytologist, 119, 345360.Google Scholar
Unger, N. (2013). Isoprene emission variability through the twentieth century. Journal of Geophysical Research: Atmospheres, 118, 1360613613, doi:10.1002/2013JD020978.Google Scholar
Unger, N., Harper, K., Zheng, Y., et al. (2013). Photosynthesis-dependent isoprene emission from leaf to planet in a global carbon–chemistry–climate model. Atmospheric Chemistry and Physics, 13, 1024310269.Google Scholar
Urbanski, S., Barford, C., Wofsy, S., et al. (2007). Factors controlling CO2 exchange on timescales from hourly to decadal at Harvard Forest. Journal of Geophysical Research, 112, G02020, doi:10.1029/2006JG000293.Google Scholar
Valente, F., David, J. S., and Gash, J. H. C. (1997). Modelling interception loss for two sparse eucalypt and pine forests in central Portugal using reformulated Rutter and Gash analytical models. Journal of Hydrology, 190, 141162.Google Scholar
Val Martin, M., Heald, C. L., and Arnold, S. R. (2014). Coupling dry deposition to vegetation phenology in the Community Earth System Model: Implications for the simulation of surface O3. Geophysical Research Letters, 41, 29882996, doi:10.1002/2014GL059651.Google Scholar
van den Honert, T. H. (1948). Water transport in plants as a catenary process. Discussions of the Faraday Society, 3, 146153.Google Scholar
van den Hurk, B. J. J. M., and McNaughton, K. G. (1995). Implementation of near-field dispersion in a simple two-layer surface resistance model. Journal of Hydrology, 166, 293311.Google Scholar
van den Hurk, B. J. J. M., Viterbo, P., Beljaars, A. C. M., and Betts, A. K. (2000). Offline Validation of the ERA40 Surface Scheme, Technical Memorandum Number 295. Reading: European Centre for Medium-Range Weather Forecasts.Google Scholar
van der Velde, I. R., Miller, J. B., Schaefer, K., et al. (2013). Biosphere model simulations of interannual variability in terrestrial 13C/12C exchange. Global Biogeochemical Cycles, 27, 637649, doi:10.1002/gbc.20048.Google Scholar
van der Werf, G. R., Randerson, J. T., Giglio, L., et al. (2010). Global fire emissions and the contribution of deforestation, savanna, forest, agricultural, and peat fires (1997–2009). Atmospheric Chemistry and Physics, 10, 1170711735.Google Scholar
van Dijk, A. I. J. M., Gash, J. H. van Gorsel, E., et al. (2015). Rainfall interception and the coupled surface water and energy balance. Agricultural and Forest Meteorology, 214/215, 402415.Google Scholar
van Genuchten, M. T. (1980). A closed-form equation for predicting the hydraulic conductivity of unsaturated soils. Soil Science Society of America Journal, 44, 892898.Google Scholar
Verhoef, A., De Bruin, H. A. R., and van den Hurk, B. J. J. M. (1997). Some practical notes on the parameter kB–1 for sparse vegetation. Journal of Applied Meteorology, 36, 560572.Google Scholar
Verhoef, A., and Egea, G. (2014). Modeling plant transpiration under limited soil water: Comparison of different plant and soil hydraulic parameterizations and preliminary implications for their use in land surface models. Agricultural and Forest Meteorology, 191, 2232.Google Scholar
Verseghy, D. L. (1991). CLASS – A Canadian land surface scheme for GCMs. I. Soil model. International Journal of Climatology, 11, 111133.Google Scholar
Verseghy, D. L., McFarlane, N. A., and Lazare, M. (1993). CLASS – A Canadian land surface scheme for GCMs, II. Vegetation model and coupled runs. International Journal of Climatology, 13, 347370.Google Scholar
Vico, G., Manzoni, S., Palmroth, S., Weih, M., and Katul, G. (2013). A perspective on optimal leaf stomatal conductance under CO2 and light co-limitations. Agricultural and Forest Meteorology, 182/183, 191199.Google Scholar
Vidale, P. L., and Stöckli, R. (2005). Prognostic canopy air space solutions for land surface exchanges. Theoretical and Applied Climatology, 80, 245257.Google Scholar
Vogel, S. (2009). Leaves in the lowest and highest winds: Temperature, force and shape. New Phytologist, 183, 1326.Google Scholar
von Caemmerer, S. (2000). Biochemical Models of Leaf Photosynthesis. Collingwood, Victoria: CSIRO Publishing.Google Scholar
von Caemmerer, S. (2013). Steady-state models of photosynthesis. Plant, Cell and Environment, 36, 16171630.Google Scholar
von Caemmerer, S., and Evans, J. R. (2015). Temperature responses of mesophyll conductance differ greatly between species. Plant, Cell and Environment, 38, 629637.Google Scholar
von Caemmerer, S., Evans, J. R., Hudson, G. S., and Andrews, T. J. (1994). The kinetics of ribulose-1,5-bisphosphate carboxylase/oxygenase in vivo inferred from measurements of photosynthesis in leaves of transgenic tobacco. Planta, 195, 8897.Google Scholar
von Caemmerer, S., Farquhar, G., and Berry, J. (2009). Biochemical model of C3 photosynthesis. In Photosynthesis in silico: Understanding Complexity from Molecules to Ecosystems, ed. Laisk, A., Nedbal, L. and Govindjee, . Dordrecht: Springer, pp. 209230.Google Scholar
Vrettas, M. D., and Fung, I. Y. (2015). Toward a new parameterization of hydraulic conductivity in climate models: Simulation of rapid groundwater fluctuations in Northern California. Journal of Advances in Modeling Earth Systems, 7, doi:10.1002/2015MS000516.Google Scholar
Waggoner, P. E. (1975). Micrometeorological models. In Vegetation and the Atmosphere: vol. 1. Principles, ed. Monteith, J. L.. New York, NY: Academic Press, pp. 205228.Google Scholar
Waggoner, P. E., Furnival, G. M., and Reifsnyder, W. E. (1969). Simulation of the microclimate in a forest. Forest Science, 15, 3745.Google Scholar
Waggoner, P. E., and Reifsnyder, W. E. (1968). Simulation of the temperature, humidity and evaporation profiles in a leaf canopy. Journal of Applied Meteorology, 7, 400409.Google Scholar
Waggoner, P. E., and Stephens, G. R. (1970). Transition probabilities for a forest. Nature, 225, 11601161.Google Scholar
Walcroft, A. S., Brown, K. J., Schuster, W. S. F., et al. (2005). Radiative transfer and carbon assimilation in relation to canopy architecture, foliage area distribution and clumping in a mature temperate rainforest canopy in New Zealand. Agricultural and Forest Meteorology, 135, 326339.Google Scholar
Walker, A. P., Quaife, T., van Bodegom, P. M., et al. (2017). The impact of alternative trait-scaling hypotheses for the maximum photosynthetic carboxylation rate (Vcmax) on global gross primary production. New Phytologist, 215, 13701386.Google Scholar
Wang, A., Li, K. Y., and Lettenmaier, D. P. (2008). Integration of the variable infiltration capacity model soil hydrology scheme into the community land model. Journal of Geophysical Research, 113, D09111, doi:10.1029/2007JD009246.Google Scholar
Wang, G. (2011). Assessing the potential hydrological impacts of hydraulic redistribution in Amazonia using a numerical modeling approach. Water Resources Research, 47, W02528, doi:10.1029/2010WR009601.Google Scholar
Wang, H., Prentice, I. C., and Davis, T. W. (2014). Biophsyical constraints on gross primary production by the terrestrial biosphere. Biogeosciences, 11, 59876001.Google Scholar
Wang, H., Prentice, I. C., Keenan, T. F., et al. (2017). Towards a universal model for carbon dioxide uptake by plants. Nature Plants, 3, 734741.Google Scholar
Wang, Y.-P. (2000). A refinement to the two-leaf model for calculating canopy photosynthesis. Agricultural and Forest Meteorology, 101, 143150.Google Scholar
Wang, Y.-P. (2003). A comparison of three different canopy radiation models commonly used in plant modelling. Functional Plant Biology, 30, 143152.Google Scholar
Wang, Y. P., and Jarvis, P. G. (1990). Description and validation of an array model – MAESTRO. Agricultural and Forest Meteorology, 51, 257280.Google Scholar
Wang, Y. P., Jarvis, P. G., and Benson, M. L. (1990). Two-dimensional needle-area density distribution within the crowns of Pinus radiata. Forest Ecology and Management, 32, 217237.Google Scholar
Wang, Y. P., Jiang, J., Chen-Charpentier, B., et al. (2016). Responses of two nonlinear microbial models to warming and increased carbon input. Biogeosciences, 13, 887902.Google Scholar
Wang, Y. P., Kowalczyk, E., Leuning, R., et al. (2011). Diagnosing errors in a land surface model (CABLE) in the time and frequency domains. Journal of Geophysical Research, 116, G01034, doi:10.1029/2010JG001385.Google Scholar
Wang, Y. P., Law, R. M., and Pak, B. (2010). A global model of carbon, nitrogen and phosphorus cycles for the terrestrial biosphere. Biogeosciences, 7, 22612282.Google Scholar
Wang, Y.-P., and Leuning, R. (1998). A two-leaf model for canopy conductance, photosynthesis and partitioning of available energy. I: Model description and comparison with a multi-layered model. Agricultural and Forest Meteorology, 91, 89111.Google Scholar
Wania, R., Ross, I., and Prentice, I. C. (2010). Implementation and evaluation of a new methane model within a dynamic global vegetation model: LPJ-WHyMe v1.3.1. Geoscientific Model Development, 3, 565584.Google Scholar
Wania, R., Melton, J. R., Hodson, E. L., et al. (2013). Present state of global wetland extent and wetland methane modelling: Methodology of a model inter-comparison project (WETCHIMP). Geoscientific Model Development, 6, 617641.Google Scholar
Waring, R. H., and Running, S. W. (1976). Water uptake, storage and transpiration by conifers: A physiological model. In Water and Plant Life: Problems and Modern Approaches, ed. Lange, O. L., Kappen, L., and Schulze, E.-D.. Berlin: Springer-Verlag, pp. 189202.Google Scholar
Waring, R. H., and Running, S. W. (2007). Forest Ecosystems: Analysis at Multiple Scales, 3rd edn. Amsterdam: Elsevier.Google Scholar
Warrick, A. W. (1991). Numerical approximations of darcian flow through unsaturated soil. Water Resources Research, 27, 12151222.Google Scholar
Warrick, A. W. (2003). Soil Water Dynamics. Oxford: Oxford University Press.Google Scholar
Webb, E. K., Pearman, G. I., and Leuning, R. (1980). Correction of flux measurements for density effects due to heat and water vapour transfer. Quarterly Journal of the Royal Meteorological Society, 106, 85100.Google Scholar
Weiler, M., and Beven, K. (2015). Do we need a Community Hydrological Model? Water Resources Research, 51, 77777784, doi:10.1002/2014WR016731.Google Scholar
Weng, E. S., Malyshev, S., Lichstein, J. W., et al. (2015). Scaling from individual trees to forests in an Earth system modeling framework using a mathematically tractable model of height-structured competition. Biogeosciences, 12, 26552694.Google Scholar
Werger, M. J. A., and Hirose, T. (1991). Leaf nitrogen distribution and whole canopy photosynthetic carbon gain in herbaceous stands. Vegetatio, 97, 1120.Google Scholar
Wesely, M. L. (1989). Parameterization of surface resistances to gaseous dry deposition in regional-scale numerical models. Atmospheric Environment, 23, 12931304.Google Scholar
Wesely, M. L., and Hicks, B. B. (1977). Some factors that affect the deposition rates of sulfur dioxide and similar gases on vegetation. Journal of the Air Pollution Control Association, 27, 11101116.Google Scholar
Wesely, M. L., and Hicks, B. B. (2000). A review of the current status of knowledge on dry deposition. Atmospheric Environment, 34, 22612282.Google Scholar
West, J. B., Bowen, G. J., Dawson, T. E., and Tu, K. P. (2010). Isoscapes: Understanding Movement, Pattern, and Process on Earth through Isotope Mapping. Dordrecht: Springer.Google Scholar
Whitehead, D. (1998). Regulation of stomatal conductance and transpiration in forest canopies. Tree Physiology, 18, 633644.Google Scholar
Whittaker, R. H., Bormann, F. H., Likens, G. E., and Siccama, T. G. (1974). The Hubbard Brook Ecosystem Study: Forest biomass and production. Ecological Monographs, 44, 233252.Google Scholar
Wieder, W. R., Allison, S. D., Davidson, E. A., et al. (2015a). Explicitly representing soil microbial processes in Earth system models. Global Biogeochemical Cycles, 29, 17821800, doi:10.1002/2015GB005188.Google Scholar
Wieder, W. R., Cleveland, C. C., Lawrence, D. M., and Bonan, G. B. (2015b). Effects of model structural uncertainty on carbon cycle projections: Biological nitrogen fixation as a case study. Environmental Research Letters, 10, 044016, doi:10.1088/1748-9326/10/4/044016.Google Scholar
Wieder, W. R., Cleveland, C. C., Smith, W. K., and Todd-Brown, K. (2015c). Future productivity and carbon storage limited by terrestrial nutrient availability. Nature Geoscience, 8, 441444.Google Scholar
Wieder, W. R., Grandy, A. S., Kallenbach, C. M., and Bonan, G. B. (2014). Integrating microbial physiology and physio-chemical principles in soils with the MIcrobial-MIneral Carbon Stabilization (MIMICS) model. Biogeosciences, 11, 38993917.Google Scholar
Wieder, W. R., Grandy, A. S., Kallenbach, C. M., Taylor, P. G., and Bonan, G. B. (2015d). Representing life in the Earth system with soil microbial functional traits in the MIMICS model. Geoscientific Model Development, 8, 17891808.Google Scholar
Wieder, W. R., Hartman, M. D., Sulman, B. N., et al. (2018). Carbon cycle confidence and uncertainty: Exploring variation among soil biogeochemical models. Global Change Biology, 24, 15631579.Google Scholar
Wiedinmyer, C., Akagi, S. K., Yokelson, R. J., et al. (2011). The Fire INventory from NCAR (FINN): A high resolution global model to estimate the emissions from open burning. Geoscientific Model Development, 4, 625641.Google Scholar
Williams, M., Bond, B. J., and Ryan, M. G. (2001a). Evaluating different soil and plant hydraulic constraints on tree function using a model and sap flow data from ponderosa pine. Plant, Cell and Environment, 24, 679690.Google Scholar
Williams, M., Eugster, W., Rastetter, E. B., McFadden, J. P., and Chapin, F. S. III (2000). The controls on net ecosystem productivity along an Arctic transect: A model comparison with flux measurements. Global Change Biology, 6(S1), 116126.Google Scholar
Williams, M., Malhi, Y., Nobre, A. D., et al. (1998). Seasonal variation in net carbon exchange and evapotranspiration in a Brazilian rain forest: A modelling analysis. Plant, Cell and Environment, 21, 953968.Google Scholar
Williams, M., Law, B. E., Anthoni, P. M., and Unsworth, M. H. (2001b). Use of a simulation model and ecosystem flux data to examine carbon–water interactions in ponderosa pine. Tree Physiology, 21, 287298.Google Scholar
Williams, M., Rastetter, E. B., Fernandes, D. N., et al. (1996). Modelling the soil–plant–atmosphere continuum in a QuercusAcer stand at Harvard Forest: The regulation of stomatal conductance by light, nitrogen and soil/plant hydraulic properties. Plant, Cell and Environment, 19, 911927.Google Scholar
Williams, M., Richardson, A. D., Reichstein, M., et al. (2009). Improving land surface models with FLUXNET data. Biogeosciences, 6, 13411359.Google Scholar
Williamson, D. L., Kiehl, J. T., Ramanathan, V., Dickinson, R. E., and Hack, J. J. (1987). Description of NCAR Community Climate Model (CCM1), Technical Note NCAR/TN-285+STR. Boulder, CO: National Center for Atmospheric Research.Google 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, 118.Google Scholar
Wilson, K. B., Baldocchi, D. D., and Hanson, P. J. (2000). Spatial and seasonal variability of photosynthetic parameters and their relationship to leaf nitrogen in a deciduous forest. Tree Physiology, 20, 565578.Google Scholar
Wilson, T. B., Norman, J. M., Bland, W. L., and Kucharik, C. J. (2003). Evaluation of the importance of Lagrangian canopy turbulence formulations in a soil–plant–atmosphere model. Agricultural and Forest Meteorology, 115, 5169.Google Scholar
Witkowski, E. T. F., and Lamont, B. B. (1991). Leaf specific mass confounds leaf density and thickness. Oecologia, 88, 486493.Google Scholar
Wittig, V. E., Ainsworth, E. A., and Long, S. P. (2007). To what extent do current and projected increases in surface ozone affect photosynthesis and stomatal conductance of trees? A meta-analytic review of the last 3 decades of experiments. Plant, Cell and Environment, 30, 11501162.Google Scholar
Wohlfahrt, G., Bahn, M., Tappeiner, U., and Cernusca, A. (2001). A multi-component, multi-species model of vegetation–atmosphere CO2 and energy exchange for mountain grasslands. Agricultural and Forest Meteorology, 106, 261287.Google Scholar
Wohlfahrt, G., and Cernusca, A. (2002). Momentum transfer by a mountain meadow canopy: A simulation analysis based on Massman’s (1997) model. Boundary-Layer Meteorology, 103, 391407.Google Scholar
Wolf, A., Anderegg, W. R. L., and Pacala, S. W. (2016). Optimal stomatal behavior with competition for water and risk of hydraulic impairment. Proceedings of the National Academy of Sciences USA, 113, E7222E7230.Google Scholar
Wolfe, G. M., and Thornton, J. A. (2011). The Chemistry of Atmosphere–Forest Exchange (CAFE) model – Part 1: Model description and characterization. Atmospheric Chemistry and Physics, 11, 77101.Google Scholar
Wolock, D. M. (1993). Simulating the Variable-Source-Area Concept of Streamflow Generation with the Watershed Model TOPMODEL, Water-Resources Investigations Report 93–4124. Lawrence, KS: U.S. Geological Survey.Google Scholar
Wolz, K. J., Wertin, T. M., Abordo, M., Wang, D., and Leakey, A. D. B. (2017). Diversity in stomatal function is integral to modelling plant carbon and water fluxes. Nature Ecology and Evolution, 1, 12921298.Google Scholar
Wong, S. C., Cowan, I. R., and Farquhar, G. D. (1978). Leaf conductance in relation to assimilation in Eucalyptus pauciflora Sieb. ex Spreng: Influence of irradiance and partial pressure of carbon dioxide. Plant Physiology, 62, 670674.Google Scholar
Wong, S. C., Cowan, I. R., and Farquhar, G. D. (1979). Stomatal conductance correlates with photosynthetic capacity. Nature, 282, 424426.Google Scholar
Wong, T. E., Nusbaumer, J., and Noone, D. C., (2017). Evaluation of modeled land-atmosphere exchanges with a comprehensive water isotope fractionation scheme in version 4 of the Community Land Model. Journal of Advances in Modeling Earth Systems, 9, 9781001, doi:10.1002/2016MS000842.Google Scholar
Wood, E. F., Lettenmaier, D. P., and Zartarian, V. G. (1992). A land-surface hydrology parameterization with subgrid variability for general circulation models. Journal of Geophysical Research, 97D, 27172728.Google Scholar
Woodward, F. I., and Lomas, M. R. (2004). Vegetation dynamics – simulating responses to climatic change. Biological Reviews, 79, 643670.Google Scholar
Woodward, F. I., Smith, T. M., and Emanuel, W. R. (1995). A global land primary productivity and phytogeography model. Global Biogeochemical Cycles, 9, 471490.Google Scholar
Wright, I. J., Reich, P. B., Westoby, M., et al. (2004). The worldwide leaf economics spectrum. Nature, 428, 821827.Google Scholar
Wu, A., Black, A., Verseghy, D. L., and Bailey, W. G. (2001). Comparison of two-layer and single-layer canopy models with Lagrangian and K-theory approaches in modelling evaporation from forests. International Journal of Climatology, 21, 18211839.Google Scholar
Wu, T., Li, W., Ji, J., et al. (2013). Global carbon budgets simulated by the Beijing Climate Center Climate System Model for the last century. Journal of Geophysical Research: Atmospheres, 118, 43264347, doi:10.1002/jgrd.50320.Google Scholar
Wu, Y., Brashers, B., Finkelstein, P. L., and Pleim, J. E. (2003). A multilayer biochemical dry deposition model. 1. Model formulation. Journal of Geophysical Research, 108D, 4013, doi:10.1029/2002JD002293.Google Scholar
Wullschleger, S. D. (1993). Biochemical limitations to carbon assimilation in C3 plants - A retrospective analysis of the A/Ci curves from 109 species. Journal of Experimental Botany, 44, 907920.Google Scholar
Xu, L., and Baldocchi, D. D. (2003). Seasonal trends in photosynthetic parameters and stomatal conductance of blue oak (Quercus douglasii) under prolonged summer drought and high temperature. Tree Physiology, 23, 865877.Google Scholar
Xia, J., Luo, Y., Wang, Y.-P., and Hararuk, O. (2013). Traceable components of terrestrial carbon storage capacity in biogeochemical models. Global Change Biology, 19, 21042116.Google Scholar
Xia, J. Y., Luo, Y. Q., Wang, Y.-P., Weng, E. S., and Hararuk, O. (2012). A semi-analytical solution to accelerate spin-up of a coupled carbon and nitrogen land model to steady state. Geoscientific Model Development, 5, 12591271.Google Scholar
Xia, J., Yuan, W., Wang, Y.-P., and Zhang, Q. (2017). Adaptive carbon allocation by plants enhances the terrestrial carbon sink. Scientific Reports, 7, 3341, doi:10.1038/s41598-017-03574-3.Google Scholar
Xu, C., Fisher, R., Wullschleger, S. D., et al. (2012). Toward a mechanistic modeling of nitrogen limitation on vegetation dynamics. PLoS ONE, 7(5), e37914, doi:10.1371/journal.pone.0037914.Google Scholar
Xu, L., Pyles, R. D., Paw U, K. T., Chen, S. H., and Monier, E. (2014). Coupling the high-complexity land surface model ACASA to the mesoscale model WRF. Geoscientific Model Development, 7, 29172932.Google Scholar
Xu, X., Medvigy, D., Powers, J. S., Becknell, J. M., and Guan, K. (2016). Diversity in plant hydraulic traits explains seasonal and inter-annual variations of vegetation dynamics in seasonally dry tropical forests. New Phytologist, 212, 8095.Google Scholar
Yan, B., and Dickinson, R. E. (2014). Modeling hydraulic redistribution and ecosystem response to droughts over the Amazon basin using Community Land Model 4.0 (CLM4). Journal of Geophysical Research: Biogeosciences, 119, 21302143, doi:10.1002/2014JG002694.Google Scholar
Yan, X., and Shugart, H. H. (2005). FAREAST: A forest gap model to simulate dynamics and patterns of eastern Eurasian forests. Journal of Biogeography, 32, 16411658.Google Scholar
Yang, X., Thornton, P. E., Ricciuto, D. M., and Post, W. M. (2014). The role of phosphorus dynamics in tropical forests – a modeling study using CLM-CNP. Biogeosciences, 11, 16671681.Google Scholar
Yeh, T.-C. J., Gelhar, L. W., and Gutjahr, A. L. (1985a). Stochastic analysis of unsaturated flow in heterogeneous soils: 1. Statistically isotropic media. Water Resources Research, 21, 447456.Google Scholar
Yeh, T.-C. J., Gelhar, L. W., and Gutjahr, A. L. (1985b). Stochastic analysis of unsaturated flow in heterogeneous soils: 2. Statistically anisotropic media with variable α. Water Resources Research, 21, 457464.Google Scholar
Yeh, T.-C. J., Gelhar, L. W., and Gutjahr, A. L. (1985c). Stochastic analysis of unsaturated flow in heterogeneous soils: 3. Observations and applications. Water Resources Research, 21, 465471.Google Scholar
Yoshimura, K., Miyazaki, S., Kanae, S., and Oki, T. (2006). Iso-MATSIRO, a land surface model that incorporates stable water isotopes. Global and Planetary Change, 51, 90107.Google Scholar
Yuan, H., Dickinson, R. E., Dai, Y., et al. (2014). A 3D canopy radiative transfer model for global climate modeling: Description, validation, and application. Journal of Climate, 27, 11681192.Google Scholar
Yuan, W., Liu, S., Zhou, G., et al. (2007). Deriving a light use efficiency model from eddy covariance flux data for predicting daily gross primary production across biomes. Agricultural and Forest Meteorology, 143, 189207.Google Scholar
Yue, C., Ciais, P., Cadule, P., et al. (2014). Modelling the role of fires in the terrestrial carbon balance by incorporating SPITFIRE into the global vegetation model ORCHIDEE – Part 1: Simulating historical global burned area and fire regimes. Geoscientific Model Development, 7, 27472767.Google Scholar
Zaehle, S., and Dalmonech, D. (2011). Carbon–nitrogen interactions on land at global scales: Current understanding in modelling climate biosphere feedbacks. Current Opinion in Environmental Sustainability, 3, 311320.Google Scholar
Zaehle, S., and Friend, A. D. (2010). Carbon and nitrogen cycle dynamics in the O-CN land surface model: 1. Model description, site-scale evaluation, and sensitivity to parameter estimates. Global Biogeochemical Cycles, 24, GB1005, doi:10.1029/2009GB003521.Google Scholar
Zaehle, S., Friend, A. D., Friedlingstein, P., et al. (2010). Carbon and nitrogen cycle dynamics in the O-CN land surface model: 2. Role of the nitrogen cycle in the historical terrestrial carbon balance. Global Biogeochemical Cycles, 24, GB1006, doi:10.1029/2009GB003522.Google Scholar
Zaehle, S., Medlyn, B. E., De Kauwe, M. G., et al. (2014). Evaluation of 11 terrestrial carbon–nitrogen cycle models against observations from two temperate Free-Air CO2 Enrichment studies. New Phytologist, 202, 803822.Google Scholar
Zender, C. S., Bian, H., and Newman, D. (2003). Mineral Dust Entrainment and Deposition (DEAD) model: Description and 1990s dust climatology. Journal of Geophysical Research, 108, 4416, doi:10.1029/2002JD002775.Google Scholar
Zeppel, M., Macinnis-Ng, C., Palmer, A., et al. (2008). An analysis of the sensitivity of sap flux to soil and plant variables assessed for an Australian woodland using a soil–plant–atmosphere model. Functional Plant Biology35, 509520.Google Scholar
Zhang, L., Brook, J. R., and Vet, R. (2003). A revised parameterization for gaseous dry deposition in air-quality models. Atmospheric Chemistry and Physics, 3, 20672082.Google Scholar
Zhao, M., Heinsch, F. A., Nemani, R. R., and Running, S. W. (2005). Improvements of the MODIS terrestrial gross and net primary production global data set. Remote Sensing of Environment, 95, 164176.Google Scholar
Zhao, R.-J. (1992). The Xinanjiang model applied in China. Journal of Hydrology, 135, 371381.Google Scholar
Zhao, R.-J., and Liu, X.-R. (1995). The Xinanjiang model. In Computer Models of Watershed Hydrology, ed. Singh, V. P.. Highlands Ranch, CO: Water Resources Publications, pp. 215232.Google Scholar
Zhao, R.-J., Zuang, Y.-L., Fang, L. R., Liu, X.-R., and Zhang, Q.-S. (1980). The Xinanjiang model. In Hydrological Forecasting: Proceedings of the Oxford Symposium, 15–18 April 1980, Publication No. 129. Wallingford: International Association of Hydrological Sciences, pp. 351356.Google Scholar
Zheng, Z., and Wang, G. (2007). Modeling the dynamic root water uptake and its hydrological impact at the Reserva Jaru site in Amazonia. Journal of Geophysical Research112, G04012, doi:10.1029/2007JG000413.Google Scholar
Zhou, S., Duursma, R. A., Medlyn, B. E., Kelly, J. W. G., and Prentice, I. C. (2013). How should we model plant responses to drought? An analysis of stomatal and non-stomatal responses to water stress. Agricultural and Forest Meteorology, 182/183, 204214.Google Scholar
Zhu, Q., and Riley, W. J. (2015). Improved modelling of soil nitrogen losses. Nature Climate Change, 5, 705706.Google Scholar
Zhu, Q., Riley, W. J., and Tang, J. (2017). A new theory of plant–microbe nutrient competition resolves inconsistencies between observations and model predictions. Ecological Applications, 27, 875886.Google Scholar
Zhu, Q., Riley, W. J., Tang, J., and Koven, C. D. (2016). Multiple soil nutrient competition between plants, microbes, and mineral surfaces: Model development, parameterization, and example applications in several tropical forests. Biogeosciences, 13, 341363.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Climate Change and Terrestrial Ecosystem Modeling
  • Online publication: 08 February 2019
  • Chapter DOI: https://doi.org/10.1017/9781107339217.023
Available formats
×

Save book to Dropbox

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

  • References
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Climate Change and Terrestrial Ecosystem Modeling
  • Online publication: 08 February 2019
  • Chapter DOI: https://doi.org/10.1017/9781107339217.023
Available formats
×

Save book to Google Drive

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

  • References
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Climate Change and Terrestrial Ecosystem Modeling
  • Online publication: 08 February 2019
  • Chapter DOI: https://doi.org/10.1017/9781107339217.023
Available formats
×