Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-29T04:12:13.645Z Has data issue: false hasContentIssue false

25 - Terrestrial Ecosystems and Earth System Models

from Part VI - Terrestrial Forcings and Feedbacks

Published online by Cambridge University Press:  05 November 2015

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

Summary

Chapter Summary

Much of our understanding of how land surface processes and terrestrial ecosystems affect weather, climate, and atmospheric composition comes from numerical models of surface energy fluxes, the hydrologic cycle, and biogeochemical cycles coupled to atmospheric models. Land surface models are coupled to atmospheric models to simulate the absorption of radiation at the land surface, the exchanges of sensible and latent heat between land and atmosphere, storage of heat in soil, and the frictional drag of vegetation and other surface elements on wind. These models were initially developed to provide the surface boundary conditions of radiative and turbulent fluxes required by atmospheric models. They have since evolved to simulate the hydrologic cycle, biogeochemical cycles, and vegetation dynamics so that the land and atmosphere are represented as a coupled system. This chapter reviews the historical development of land surface models. Model evaluation is discussed, as well as application of the models in climate model experiments.

Hydrometeorological Models

Global climate models represent a set of numerical equations that describe the large-scale circulation of the atmosphere and ocean and their physical state, including interactions among oceans, atmosphere, land, and sea ice that affect climate. The land surface fluxes of energy, moisture, and momentum and the associated hydrologic cycle that regulates them have long been represented in global climate models. In these models, absorption of radiation at the surface, the reflection of solar radiation and emission of longwave radiation, sensible and latent heat fluxes, storage of heat in soil, and frictional drag of the surface on wind influence climate. The land surface models used with climate models provide these biogeophysical boundary conditions at the land–atmosphere interface. They partition net radiation at the surface into sensible and latent heat fluxes, soil heat storage, and snow melt. They also partition precipitation into runoff, evaporation, and water storage in snow or soil. The most recent versions of these models simulate biogeochemical cycles (e.g., carbon), wildfires, land use, and land-cover change. The models update the state variables (snow cover, soil moisture, soil temperature, vegetation cover, leaf area index, and carbon pools) that regulate surface fluxes with the atmosphere.

First Generation Models

The first generation of land surface models used aerodynamic bulk transfer equations and simple prescriptions of albedo, surface roughness, and soil water without explicitly representing vegetation or the hydrologic cycle (e.g., Manabe et al. 1965; Williamson et al. 1987).

Type
Chapter
Information
Ecological Climatology
Concepts and Applications
, pp. 453 - 482
Publisher: Cambridge University Press
Print publication year: 2015

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

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, 637–661.CrossRefGoogle ScholarPubMed
Anav, A., Friedlingstein, P., Kidston, M., et al. (2013). Evaluating the land and ocean components of the global carbon cycle in the CMIP5 Earth system models. Journal of Climate, 26, 6801–6843.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, 8037–8052.CrossRefGoogle Scholar
Arora, V. K., and Boer, G. J. (2005). Fire as an interactive component of dynamic vegetation models. Journal of Geophysical Research, 110, G02008, doi:10.1029/2005JG000042.CrossRefGoogle Scholar
Avissar, R. (1992). Conceptual aspects of a statistical-dynamical approach to represent landscape subgrid-scale heterogeneities in atmospheric models. Journal of Geophysical Research, 97D, 2729–2742.Google Scholar
Avissar, R., and Pielke, R. A. (1989). A parameterization of heterogeneous land surfaces for atmospheric numerical models and its impact on regional meteorology. Monthly Weather Review, 117, 2113–2136.2.0.CO;2>CrossRefGoogle Scholar
Barlage, M., and Zeng, X. (2004). The effects of observed fractional vegetation cover on the land surface climatology of the Community Land Model. Journal of Hydrometeorology, 5, 823–830.2.0.CO;2>CrossRefGoogle Scholar
Best, M. J. (2005). Representing urban areas within operational numerical weather prediction models. Boundary-Layer Meteorology, 114, 91–109.CrossRefGoogle Scholar
Best, M. J., Grimmond, C. S. B., and Villani, M. G. (2006). Evaluation of the urban tile in MOSES using surface energy balance observations. Boundary-Layer Meteorology, 118, 503–525.CrossRefGoogle Scholar
Blyth, E., Clark, D. B., Ellis, R., et al. (2011). A comprehensive set of benchmark tests for a land surface model of simultaneous fluxes of water and carbon at both the global and seasonal scale. Geoscientific Model Development, 4, 255–269.CrossRefGoogle Scholar
Boisier, J. P., de Noblet-Ducoudré, N., Pitman, A. J., et al. (2012). Attributing the impacts of land-cover changes in temperate regions on surface temperature and heat fluxes to specific causes: Results from the first LUCID set of simulations. Journal of Geophysical Research, 117, D12116, doi:10.1029/2011JD017106.CrossRefGoogle 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, 2817–2831.Google Scholar
Bonan, G. B. (2008). Forests and climate change: Forcings, feedbacks, and the climate benefits of forests. Science, 320, 1444–1449.CrossRefGoogle ScholarPubMed
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, 29065–29075.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.CrossRefGoogle 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, 1543–1566.CrossRefGoogle 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.CrossRefGoogle Scholar
Bonan, G. B., Oleson, K. W., Fisher, R. A., Lasslop, G., and Reichstein, M. (2012). Reconciling leaf physiological traits and canopy flux data: Use of the TRY and FLUXNET databases in the Community Land Model version 4. Journal of Geophysical Research, 117, G02026, doi:10.1029/2011JG001913.CrossRefGoogle 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, 957–974.CrossRefGoogle Scholar
Bondeau, A., Smith, P. C., Zaehle, S., et al. (2007). Modelling the role of agriculture for the 20th century global terrestrial carbon balance. Global Change Biology, 13, 679–706.CrossRefGoogle Scholar
Boone, A., Habets, F., Noilhan, J., et al. (2004). The Rhône-aggregation land surface scheme intercomparison project: An overview. Journal of Climate, 17, 187–208.2.0.CO;2>CrossRefGoogle Scholar
Bounoua, L., Collatz, G. J., Sellers, P. J., et al. (1999). Interactions between vegetation and climate: Radiative and physiological effects of doubled atmospheric CO2. Journal of Climate, 12, 309–324.2.0.CO;2>CrossRefGoogle Scholar
Bounoua, L., Collatz, G. J., Los, S. O., et al. (2000). Sensitivity of climate to changes in NDVI. Journal of Climate, 13, 2277–2292.2.0.CO;2>CrossRefGoogle Scholar
Bowling, L. C., Lettenmaier, D. P., Nijssen, B., et al. (2003a). Simulation of high-latitude hydrological processes in the Torne–Kalix basin: PILPS Phase 2(e), 1: Experiment description and summary intercomparisons. Global and Planetary Change, 38, 1–30.CrossRefGoogle Scholar
Bowling, L. C., Lettenmaier, D. P., Nijssen, B., et al. (2003b). Simulation of high-latitude hydrological processes in the Torne–Kalix basin: PILPS Phase 2(e), 3: Equivalent model representation and sensitivity experiments. Global and Planetary Change, 38, 55–71.CrossRefGoogle Scholar
Buermann, W., Dong, J., Zeng, X., Myneni, R. B., and Dickinson, R. E. (2001). Evaluation of the utility of satellite-based vegetation leaf area index data for climate simulations. Journal of Climate, 14, 3536–3551.2.0.CO;2>CrossRefGoogle Scholar
Buermann, W., Wang, Y., Dong, J., et al. (2002). Analysis of a multiyear global vegetation leaf area index data set. Journal of Geophysical Research, 107, 4646, doi:10.1029/2001JD000975.CrossRefGoogle Scholar
Cadule, P., Friedlingstein, P., Bopp, L., et al. (2010). Benchmarking coupled climate–carbon models against long-term atmospheric CO2 measurements. Global Biogeochemical Cycles, 24, GB2016, doi:10.1029/2009GB003556.CrossRefGoogle Scholar
Chase, T. N., Pielke, R. A., Kittel, T. G. F., Nemani, R., and Running, S. W. (1996). Sensitivity of a general circulation model to global changes in leaf area index. Journal of Geophysical Research, 101D, 7393–7408.Google Scholar
Chen, F., Kusaka, H., Bornstein, R., et al. (2011). The integrated WRF/urban modelling system: Development, evaluation, and applications to urban environmental problems. International Journal of Climatology, 31, 273–288.CrossRefGoogle 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, 1989–2014.2.0.CO;2>CrossRefGoogle Scholar
Chen, T. H., Henderson-Sellers, A., Milly, P. C. D., et al. (1997). Cabauw experimental results from the Project for Intercomparison of Land-Surface Parameterization Schemes. Journal of Climate, 10, 1194–1215.2.0.CO;2>CrossRefGoogle 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, 183–203.CrossRefGoogle Scholar
Cox, P. M., Pearson, D., Booth, B. B., et al. (2013). Sensitivity of tropical carbon to climate change constrained by carbon dioxide variability. Nature, 494, 341–344.CrossRefGoogle ScholarPubMed
Craig, S. G., Holmén, K. J., Bonan, G. B., and Rasch, P. J. (1998). Atmospheric CO2 simulated by the National Center for Atmospheric Research Community Climate Model, 1: Mean fields and seasonal cycles. Journal of Geophysical Research, 103D, 13213–13235.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, 141–153.CrossRefGoogle ScholarPubMed
Deardorff, J. W. (1978). Efficient prediction of ground surface temperature and moisture, with inclusion of a layer of vegetation. Journal of Geophysical Research, 83C, 1889–1903.Google Scholar
DeFries, R. S., and Townshend, J. R. G. (1994). NDVI-derived land cover classifications at a global scale. International Journal of Remote Sensing, 15, 3567–3586.CrossRefGoogle Scholar
DeFries, R. S., Field, C. B., Fung, I., et al. (1995). Mapping the land surface for global atmosphere–biosphere models: Toward continuous distributions of vegetation's functional properties. Journal of Geophysical Research, 100D, 20867–20882.Google Scholar
DeFries, R., Hansen, M., Steininger, M., et al. (1997). Subpixel forest cover in Central Africa from multisensor, multitemporal data. Remote Sensing of Environment, 60, 228–246.CrossRefGoogle Scholar
DeFries, R. S., Hansen, M., Townshend, J. R. G., and Sohlberg, R. (1998). Global land cover classifications at 8 km spatial resolution: The use of training data derived from Landsat Imagery in decision tree classifiers. International Journal of Remote Sensing, 19, 3141–3168.Google Scholar
DeFries, R. S., Townshend, J. R. G., and Hansen, M. C. (1999). Continuous fields of vegetation characteristics at the global scale at 1-km resolution. Journal of Geophysical Research, 104D, 16911–16923.Google Scholar
DeFries, R. S., Hansen, M. C., and Townshend, J. R. G. (2000a). Global continuous fields of vegetation characteristics: A linear mixture model applied to multi-year 8 km AVHRR data. International Journal of Remote Sensing, 21, 1389–1414.CrossRefGoogle Scholar
DeFries, R. S., Hansen, M. C., Townshend, J. R. G., Janetos, A. C., and Loveland, T. R. (2000b). A new global 1-km dataset of percentage tree cover derived from remote sensing. Global Change Biology, 6, 247–254.CrossRefGoogle 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, 1759–1779.CrossRefGoogle ScholarPubMed
De Kauwe, M. G., Medlyn, B. E., Zaehle, S., et al. (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, 883–899.CrossRefGoogle Scholar
Demuzere, M., Oleson, K., Coutts, A. M., Pigeon, G., and van Lipzig, N. P. M. (2013). Simulating the surface energy balance over two contrasting urban environments using the Community Land Model Urban. International Journal of Climatology, 33, 3182–3205.CrossRefGoogle Scholar
Denning, A. S., Fung, I. Y., and Randall, D. (1995). Latitudinal gradient of atmospheric CO2 due to seasonal exchange with land biota. Nature, 376, 240–243.CrossRefGoogle Scholar
Denning, A. S., Collatz, G. J., Zhang, C., et al. (1996a). Simulations of terrestrial carbon metabolism and atmospheric CO2 in a general circulation model. Part 1: Surface carbon fluxes. Tellus B, 48, 521–542.Google Scholar
Denning, A. S., Randall, D. A., Collatz, G. J., and Sellers, P. J. (1996b). Simulations of terrestrial carbon metabolism and atmospheric CO2 in a general circulation model. Part 2: Simulated CO2 concentrations. Tellus B, 48, 543–567.Google Scholar
de Noblet-Ducoudré, N., Boisier, J.-P., Pitman, A., et al. (2012). Determining robust impacts of land-use-induced land cover changes on surface climate over North America and Eurasia: Results from the first set of LUCID experiments. Journal of Climate, 25, 3261–3281.CrossRefGoogle Scholar
Desborough, C. E., Pitman, A. J., and Irannejad, P. (1996). Analysis of the relationship between bare soil evaporation and soil moisture simulated by 13 land surface schemes for a simple non-vegetated site. Global and Planetary Change, 13, 47–56.CrossRefGoogle Scholar
Dickinson, R. E. (1983). Land surface processes and climate-surface albedos and energy balance. Advances in Geophysics, 25, 305–353.Google Scholar
Dickinson, R. E., and Henderson-Sellers, A. (1988). Modelling tropical deforestation: A study of GCM land-surface parameterizations. Quarterly Journal of the Royal Meteorological Society, 114, 439–462.CrossRefGoogle 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, Colorado: 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, Colorado: National Center for Atmospheric Research.Google Scholar
Dickinson, R. E., Oleson, K. W., Bonan, G., et al. (2006). The Community Land Model and its climate statistics as a component of the Community Climate System Model. Journal of Climate, 19, 2302–2324.CrossRefGoogle Scholar
Dietze, M. C., Serbin, S. P., Davidson, C., et al. (2014). A quantitative assessment of a terrestrial biosphere model's data needs across North American biomes. Journal of Geophysical Research: Biogeosciences, 119, 286–300, doi:10.1002/2013JG002392.Google Scholar
Dirmeyer, P. A. (2011). A history and review of the Global Soil Wetness Project (GSWP). Journal of Hydrometeorology, 12, 729–749.CrossRefGoogle Scholar
Dorman, J. L., and Sellers, P. J. (1989). A global climatology of albedo, roughness length and stomatal resistance for atmospheric general circulation models as represented by the simple biosphere model (SiB). Journal of Applied Meteorology, 28, 833–855.2.0.CO;2>CrossRefGoogle Scholar
Drewniak, B., Song, J., Prell, J., Kotamarthi, V. R., and Jacob, R. (2013). Modeling agriculture in the Community Land Model. Geoscientific Model Development, 6, 495–515.CrossRefGoogle Scholar
Ducharne, A., Koster, R. D., Suarez, M. J., and Kumar, P. (1999). A catchment-based land surface model for GCMs and the framework for its evaluation. Physics and Chemistry of the Earth, Part B: Hydrology, Oceans and Atmosphere, 24, 769–773.CrossRefGoogle 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, 24823–24838.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, 816–831.2.0.CO;2>CrossRefGoogle 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, 530–543.2.0.CO;2>CrossRefGoogle Scholar
Fisher, R., McDowell, N., Purves, D., et al. (2010). Assessing uncertainties in a second-generation dynamic vegetation model caused by ecological scale limitations. New Phytologist, 187, 666–681.CrossRefGoogle 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, 603–628.CrossRefGoogle Scholar
Foley, J. A., Levis, S., Prentice, I. C., Pollard, D., and Thompson, S. L. (1998). Coupling dynamic models of climate and vegetation. Global Change Biology, 4, 561–579.CrossRefGoogle Scholar
Foley, J. A., Levis, S., Costa, M. H., Cramer, W., and Pollard, D. (2000). Incorporating dynamic vegetation cover within global climate models. Ecological Applications, 10, 1620–1632.CrossRefGoogle Scholar
Friedl, M. A., McIver, D. K., Hodges, J. C. F., et al. (2002). Global land cover mapping from MODIS: Algorithms and early results. Remote Sensing of Environment, 83, 287–302.CrossRefGoogle 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, 3337–3353.CrossRefGoogle 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, 511–526.CrossRefGoogle Scholar
Gedney, N., Cox, P. M., and Huntingford, C. (2004). Climate feedback from wetland methane emissions. Geophysical Research Letters, 31, L20503, doi:10.1029/2004GL020919.CrossRefGoogle Scholar
Gerber, S., Hedin, L. O., Oppenheimer, M., Pacala, S. W., and Shevliakova, E. (2010). Nitrogen cycling and feedbacks in a global dynamic land model. Global Biogeochemical Cycles, 24, GB1001, doi:10.1029/2008GB003336.CrossRefGoogle Scholar
Gervois, S., de Noblet-Ducoudré, N., Viovy, N., et al. (2004). Including croplands in a global biosphere model: Methodology and evaluation at specific sites. Earth Interactions, 8, 1–25.2.0.CO;2>CrossRefGoogle Scholar
Gervois, S., Ciais, P., de Noblet-Ducoudré, N., et al. (2008). Carbon and water balance of European croplands throughout the 20th century. Global Biogeochemical Cycles, 22, GB2022, doi:10.1029/2007GB003018.CrossRefGoogle Scholar
Goll, D. S., Brovkin, V., Parida, B. R., et al. (2012). Nutrient limitation reduces land carbon uptake in simulations with a model of combined carbon, nitrogen and phosphorus cycling. Biogeosciences, 9, 3547–3569.CrossRefGoogle Scholar
Grimmond, C. S. B., Blackett, M., Best, M. J., et al. (2010). The international urban energy balance models comparison project: First results from phase 1. Journal of Applied Meteorology and Climatology, 49, 1268–1292.CrossRefGoogle Scholar
Grimmond, C. S. B., Blackett, M., Best, M. J., et al. (2011). Initial results from Phase 2 of the international urban energy balance model comparison. International Journal of Climatology, 31, 244–272.CrossRefGoogle 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, 8873–8892.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, 3181–3210.CrossRefGoogle 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, 1471–1492.CrossRefGoogle Scholar
Gutman, G., and Ignatov, A. (1998). The derivation of the green vegetation fraction from NOAA/AVHRR data for use in numerical weather prediction models. International Journal of Remote Sensing, 19, 1533–1543.CrossRefGoogle Scholar
Hansen, M. C., and DeFries, R. S. (2004). Detecting long-term global forest change using continuous fields of tree-cover maps from 8-km Advanced Very High Resolution Radiometer (AVHRR) data for the years 1982–1999. Ecosystems, 7, 695–716.CrossRefGoogle Scholar
Hansen, M. C., DeFries, R. S., Townshend, J. R. G., and Sohlberg, R. (2000). Global land cover classification at 1km spatial resolution using a classification tree approach. International Journal of Remote Sensing, 21, 1331–1364.CrossRefGoogle Scholar
Hansen, M. C., DeFries, R. S., Townshend, J. R. G., et al. (2002). Towards an operational MODIS continuous field of percent tree cover algorithm: examples using AVHRR and MODIS data. Remote Sensing of Environment, 83, 303–319.CrossRefGoogle Scholar
Hansen, M. C., DeFries, R. S., Townshend, J. R. G., et al. (2003). Global percent tree cover at a spatial resolution of 500 meters: First results of the MODIS vegetation continuous fields algorithm. Earth Interactions, 7, 1–15.2.0.CO;2>CrossRefGoogle Scholar
Hararuk, O., Xia, J., and Luo, Y. (2014). Evaluation and improvement of a global land model against soil carbon data using a Bayesian Markov chain Monte Carlo method. Journal of Geophysical Research: Biogeosciences, 119, 403–417, doi:10.1002/2013JG002535.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.CrossRefGoogle Scholar
Henderson-Sellers, A. (1993). Continental vegetation as a dynamic component of a global climate model: A preliminary assessment. Climatic Change, 23, 337–377.CrossRefGoogle Scholar
Henderson-Sellers, A., and McGuffie, K. (1995). Global climate models and “dynamic” vegetation changes. Global Change Biology, 1, 63–75.CrossRefGoogle Scholar
Henderson-Sellers, A., McGuffie, K., and Pitman, A. J. (1996). The Project for Intercomparison of Land-surface Parameterization Schemes (PILPS): 1992 to 1995. Climate Dynamics, 12, 849–859.CrossRefGoogle Scholar
Henderson-Sellers, A., Irannejad, P., McGuffie, K., and Pitman, A. J. (2003). Predicting land-surface climates – better skill or moving targets?Geophysical Research Letters, 30, 1777, doi:10.1029/2003GL017387.CrossRefGoogle Scholar
Jarvis, P. G. (1976). The interpretation of the variations in leaf water potential and stomatal conductance found in canopies in the field. Philosophical Transactions of the Royal Society B, 273, 593–610.CrossRefGoogle Scholar
Jung, M., Reichstein, M., Margolis, H. A., et al. (2011). Global patterns of land–atmosphere fluxes of carbon dioxide, latent heat, and sensible heat derived from eddy covariance, satellite, and meteorological observations. Journal of Geophysical Research, 116, G00J07, doi:10.1029/2010JG001566.CrossRefGoogle Scholar
Kleinen, T., Brovkin, V., and Schuldt, R. J. (2012). A dynamic model of wetland extent and peat accumulation: Results for the Holocene. Biogeosciences, 9, 235–248.CrossRefGoogle 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, 1877–1902.CrossRefGoogle 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, 1578–1591.2.0.CO;2>CrossRefGoogle Scholar
Koster, R. D., and Suarez, M. J. (1992a). Modeling the land surface boundary in climate models as a composite of independent vegetation stands. Journal of Geophysical Research, 97D, 2697–2715.Google Scholar
Koster, R. D., and Suarez, M. J. (1992b). A comparative analysis of two land surface heterogeneity representations. Journal of Climate, 5, 1379–1390.2.0.CO;2>CrossRefGoogle Scholar
Koster, R. D., Suarez, M. J., Ducharne, A., Stieglitz, M., and Kumar, P. (2000). A catchment-based approach to modeling land surface processes in a general circulation model 1: Model structure. Journal of Geophysical Research, 105D, 24809–24822.Google Scholar
Koven, C., Friedlingstein, P., Ciais, P., et al. (2009). On the formation of high-latitude soil carbon stocks: Effects of cryoturbation and insulation by organic matter in a land surface model. Geophysical Research Letters, 36, L21501, doi:10.1029/2009GL040150.CrossRefGoogle Scholar
Koven, C. D., Ringeval, B., Friedlingstein, P., et al. (2011). Permafrost carbon–climate feedbacks accelerate global warming. Proceedings of the National Academy of Sciences USA, 36, 14769–14774.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, 7109–7131.CrossRefGoogle Scholar
Kucharik, C. J. (2003). Evaluation of a process-based agro-ecosystem model (Agro-IBIS) across the U.S. cornbelt: Simulations of the interannual variability in maize yield. Earth Interactions, 7, 1–33.2.0.CO;2>CrossRefGoogle Scholar
Kumar, S., Dirmeyer, P. A., Merwade, V., et al. (2013). Land use/cover change impacts in CMIP5 climate simulations: A new methodology and 21st century challenges. Journal of Geophysical Research: Atmospheres, 118, 6337–6353, doi:10.1002/jgrd.50463.Google Scholar
Lawrence, D. M., Thornton, P. E., Oleson, K. W., and Bonan, G. B. (2007). The partitioning of evapotranspiration into transpiration, soil evaporation, and canopy evaporation in a GCM: Impacts on land–atmosphere interaction. Journal of Hydrometeorology, 8, 862–880.CrossRefGoogle Scholar
Lawrence, D. M., Oleson, K. W., Flanner, M. G., et al. (2011). Parameterization improvements and functional and structural advances in version 4 of the Community Land Model. Journal of Advances in Modeling Earth Systems, 3, doi:10.1029/2011MS000045.Google Scholar
Lawrence, D. M., Oleson, K. W., Flanner, M. G., et al. (2012). The CCSM4 land simulation, 1850–2005: Assessment of surface climate and new capabilities. Journal of Climate, 25, 2240–2260.CrossRefGoogle Scholar
Lawrence, P. J., and Chase, T. N. (2007). Representing a new MODIS consistent land surface in the Community Land Model (CLM 3.0). Journal of Geophysical Research, 112, G01023, doi:10.1029/2006JG000168.CrossRefGoogle Scholar
Le Quéré, C., Raupach, M. R., Canadell, J. G., et al. (2009). Trends in the sources and sinks of carbon dioxide. Nature Geoscience, 2, 831–836.CrossRefGoogle 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.CrossRefGoogle Scholar
Levis, S., Bonan, G. B., Kluzek, E., et al. (2012). Interactive crop management in the Community Earth System Model (CESM1): Seasonal influences on land–atmosphere fluxes. Journal of Climate, 25, 4839–4859.CrossRefGoogle Scholar
Li, B., and Avissar, R. (1994). The impact of spatial variability of land-surface characteristics on land-surface heat fluxes. Journal of Climate, 7, 527–537.2.0.CO;2>CrossRefGoogle Scholar
Li, F., Zeng, X. D., and Levis, S. (2012). A process-based fire parameterization of intermediate complexity in a Dynamic Global Vegetation Model. Biogeosciences, 9, 2761–2780.CrossRefGoogle 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, 2293–2314.CrossRefGoogle Scholar
Liang, X., Wood, E. F., Lettenmaier, D. P., et al. (1998). The Project for Intercomparison of Land-surface Parameterization Schemes (PILPS) phase 2(c) Red–Arkansas River basin experiment, 2: Spatial and temporal analysis of energy fluxes. Global and Planetary Change, 19, 137–159.CrossRefGoogle Scholar
Lohmann, D., Lettenmaier, D. P., Liang, X., et al. (1998). The Project for Intercomparison of Land-surface Parameterization Schemes (PILPS) phase 2(c) Red–Arkansas River basin experiment, 3: Spatial and temporal analysis of water fluxes. Global and Planetary Change, 19, 161–179.CrossRefGoogle 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, 6815–6831.CrossRefGoogle Scholar
Loveland, T. R., Reed, B. C., Brown, J. F., et al. (2000). Development of a global land cover characteristics database and IGBP DISCover from 1km AVHRR data. International Journal of Remote Sensing, 21, 1303–1330.CrossRefGoogle Scholar
Luo, L., Robock, A., Vinnikov, K. Y., et al. (2003). Effects of frozen soil on soil temperature, spring infiltration, and runoff: Results from the PILPS 2(d) experiment at Valdai, Russia. Journal of Hydrometeorology, 4, 334–351.2.0.CO;2>CrossRefGoogle Scholar
Luo, Y. Q., Randerson, J. T., Abramowitz, G., et al. (2012). A framework for benchmarking land models. Biogeosciences, 9, 3857–3874.CrossRefGoogle 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
Manabe, S. (1969). Climate and the ocean circulation, I: The atmospheric circulation and the hydrology of the Earth's surface. Monthly Weather Review, 97, 739–774.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, 769–798.2.3.CO;2>CrossRefGoogle Scholar
Marticorena, B., and Bergametti, G. (1995). Modeling the atmospheric dust cycle, 1: Design of a soil-derived dust emission scheme. Journal of Geophysical Research, 100D, 16415–16430.Google Scholar
Masson, V. (2000). A physically-based scheme for the urban energy budget in atmospheric models. Boundary-Layer Meteorology, 94, 357–397.CrossRefGoogle Scholar
Masson, V., Grimmond, C. S. B., and Oke, T. R. (2002). Evaluation of the Town Energy Balance (TEB) scheme with direct measurements from dry districts in two cities. Journal of Applied Meteorology, 41, 1011–1026.Google Scholar
Matthews, E. (1983). Global vegetation and land-use: New high-resolution data bases for climate studies. Journal of Climate and Applied Meteorology, 22, 474–487.2.0.CO;2>CrossRefGoogle 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, 1021–1036.CrossRefGoogle Scholar
Meng, L., Hess, P. G. M., Mahowald, N. M., et al. (2012). Sensitivity of wetland methane emissions to model assumptions: Application and model testing against site observations. Biogeosciences, 9, 2793–2819.CrossRefGoogle Scholar
Mercado, L. M., Bellouin, N., Sitch, S., et al. (2009). Impact of changes in diffuse radiation on the global land carbon sink. Nature, 458, 1014–1017.CrossRefGoogle ScholarPubMed
Myneni, R. B., Nemani, R. R., and Running, S. W. (1997). Estimation of global leaf area index and absorbed par using radiative transfer models. IEEE Transactions on Geoscience and Remote Sensing, 35, 1380–1393.CrossRefGoogle Scholar
Nemani, R. R., Running, S. W., Pielke, R. A., and Chase, T. N. (1996). Global vegetation cover changes from coarse resolution satellite data. Journal of Geophysical Research, 101D, 7157–7162.Google Scholar
Nijssen, B., Bowling, L. C., Lettenmaier, D. P., et al. (2003). Simulation of high latitude hydrological processes in the Torne–Kalix basin: PILPS Phase 2(e). 2: Comparison of model results with observations. Global and Planetary Change, 38, 31–53.CrossRefGoogle Scholar
Oleson, K. W., Bonan, G. B., Feddema, J., Vertenstein, M., and Grimmond, C. S. B. (2008). An urban parameterization for a global climate model, Part I: Formulation and evaluation for two cities. Journal of Applied Meteorology and Climatology, 47, 1038–1060.CrossRefGoogle Scholar
Oleson, K. W., Bonan, G. B., Feddema, J., and Jackson, T. (2011). An examination of urban heat island characteristics in a global climate model. International Journal of Climatology, 31, 1848–1865.CrossRefGoogle Scholar
Olson, J. S., Watts, J. A., and Allison, L. J. (1983). Carbon in Live Vegetation of Major World Ecosystems, ORNL-5862. Oak Ridge, Tennessee: Oak Ridge National Laboratory.Google Scholar
Osborne, T. M., Lawrence, D. M., Challinor, A. J., Slingo, J. M., and Wheeler, T. R. (2007). Development and assessment of a coupled crop–climate model. Global Change Biology, 13, 169–183.CrossRefGoogle Scholar
Osborne, T., Slingo, J., Lawrence, D., and Wheeler, T. (2009). Examining the interaction of growing crops with local climate using a coupled crop–climate model. Journal of Climate, 22, 1393–1411.CrossRefGoogle 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, 4371–4389.CrossRefGoogle Scholar
Piao, S., Sitch, S., Ciais, P., et al. (2013). Evaluation of terrestrial carbon cycle models for their response to climate variability and to CO2 trends. Global Change Biology, 19, 2117–2132.CrossRefGoogle ScholarPubMed
Pitman, A. J. (2003). The evolution of, and revolution in, land surface schemes designed for climate models. International Journal of Climatology, 23, 479–510.CrossRefGoogle Scholar
Pitman, A. J., Henderson-Sellers, A., Desborough, C. E., et al. (1999). Key results and implications from phase 1(c) of the Project for Intercomparison of Land-surface Parametrization Schemes. Climate Dynamics, 15, 673–684.CrossRefGoogle Scholar
Pitman, A. J., de Noblet-Ducoudré, N., Cruz, F. T., et al. (2009). Uncertainties in climate responses to past land cover change: First results from the LUCID intercomparison study. Geophysical Research Letters, 36, L14814, doi:10.1029/2009GL039076.CrossRefGoogle Scholar
Price, J. C. (1992). Estimating vegetation amount from visible and near infrared reflectances. Remote Sensing of Environment, 41, 29–34.CrossRefGoogle Scholar
Qu, W., Henderson-Sellers, A., Pitman, A. J., et al. (1998). Sensitivity of latent heat flux from PILPS land-surface schemes to perturbations of surface air temperature. Journal of the Atmospheric Sciences, 55, 1909–1927.2.0.CO;2>CrossRefGoogle Scholar
Randall, D. A., Dazlich, D. A., Zhang, C., et al. (1996). A revised land surface parameterization (SiB2) for GCMS, Part III: The greening of the Colorado State University General Circulation Model. Journal of Climate, 9, 738–763.2.0.CO;2>CrossRefGoogle 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, 2462–2484.CrossRefGoogle Scholar
Reich, P. B., Rich, R. L., Lu, X., Wang, Y.-P., and Oleksyn, J. (2014). Biogeographic variation in evergreen conifer needle longevity and impacts on boreal forest carbon cycle projections. Proceedings of the National Academy of Sciences USA, 111, 13703–13708.CrossRefGoogle ScholarPubMed
Reichstein, M., Ciais, P., Papale, D., et al. (2007). Reduction of ecosystem productivity and respiration during the European summer 2003 climate anomaly: A joint flux tower, remote sensing and modelling analysis. Global Change Biology, 13, 634–651.CrossRefGoogle 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.CrossRefGoogle Scholar
Riley, W. J., Still, C. J., Helliker, B. R., Ribas-Carbo, M., and Berry, J. A. (2003). 18O composition of CO2 and H2O ecosystem pools and fluxes in a tallgrass prairie: Simulations and comparisons to measurements. Global Change Biology, 9, 1567–1581.CrossRefGoogle 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, 1925–1953.CrossRefGoogle Scholar
Ringeval, B., Decharme, B., Piao, S. L., et al. (2012). Modelling sub-grid wetland in the ORCHIDEE global land surface model: Evaluation against river discharges and remotely sensed data. Geoscientific Model Development, 5, 941–962.CrossRefGoogle Scholar
Risi, C., Bony, S., Vimeux, F., and Jouzel, J. (2010). Water-stable isotopes in the LMDZ4 general circulation model: Model evaluation for present-day and past climates and applications to climatic interpretations of tropical isotopic records. Journal of Geophysical Research, 115, D12118, doi:10.1029/2009JD013255.CrossRefGoogle Scholar
Rogers, A. (2014). The use and misuse of Vc,max in Earth System Models. Photosynthesis Research, 119, 15–29.CrossRefGoogle Scholar
Scheiter, S., Langan, L., and Higgins, S. I. (2013). Next-generation dynamic global vegetation models: Learning from community ecology. New Phytologist, 198, 957–969.CrossRefGoogle ScholarPubMed
Schlosser, C. A., Slater, A. G., Robock, A., et al. (2000). Simulations of a boreal grassland hydrology at Valdai, Russia: PILPS Phase 2(d). Monthly Weather Review, 128, 301–321.2.0.CO;2>CrossRefGoogle 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, 505–531.2.0.CO;2>CrossRefGoogle Scholar
Sellers, P. J., Tucker, C. J., Collatz, G. J., et al. (1994). A global 1° by 1° NDVI data set for climate studies, Part 2: The generation of global fields of terrestrial biophysical parameters from the NDVI. International Journal of Remote Sensing, 15, 3519–3545.CrossRefGoogle Scholar
Sellers, P. J., Bounoua, L., Collatz, G. J., et al. (1996a). Comparison of radiative and physiological effects of doubled atmospheric CO2 on climate. Science, 271, 1402–1406.CrossRefGoogle 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, 706–737.Google Scholar
Sellers, P. J., Randall, D. A., Collatz, G. J., et al. (1996c). A revised land surface parameterization (SiB2) for atmospheric GCMs, Part I: Model formulation. Journal of Climate, 9, 676–705.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, 18651–18667.Google Scholar
Shao, Y., and Henderson-Sellers, A. (1996a). Modeling soil moisture: A Project for Intercomparison of Land Surface Parameterization Schemes Phase 2(b). Journal of Geophysical Research, 101D, 7227–7250.Google Scholar
Shao, Y., and Henderson-Sellers, A. (1996b). Validation of soil moisture simulation in landsurface parameterisation schemes with HAPEX data. Global and Planetary Change, 13, 11–46.CrossRefGoogle 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.CrossRefGoogle Scholar
Shukla, J., and Mintz, Y. (1982). Influence of land-surface evapotranspiration on the Earth's climate. Science, 215, 1498–1501.CrossRefGoogle ScholarPubMed
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, 791–794.CrossRefGoogle ScholarPubMed
Sitch, S., Huntingford, C., Gedney, N., et al. (2008). Evaluation of the terrestrial carbon cycle, future plant geography and climate–carbon cycle feedbacks using five Dynamic Global Vegetation Models (DGVMs). Global Change Biology, 14, 2015–2039.CrossRefGoogle Scholar
Slater, A. G., Schlosser, C. A., Desborough, C. E., et al. (2001). The representation of snow in land surface schemes: Results from PILPS 2(d). Journal of Hydrometeorology, 2, 7–25.2.0.CO;2>CrossRefGoogle Scholar
Stöckli, R., Lawrence, D. M., Niu, G.-Y., et al. (2008). Use of FLUXNET in the Community Land Model development. Journal of Geophysical Research, 113, G01025, doi:10.1029/2007JG000562.CrossRefGoogle Scholar
Swenson, S. C., Lawrence, D. M., and Lee, H. (2012). Improved simulation of the terrestrial hydrological cycle in permafrost regions by the Community Land Model. Journal of Advances in Modeling Earth Systems, 4, M08002, doi:10.1029/2012MS000165.CrossRefGoogle Scholar
Tesfa, T. K., Li, H.-Y., Leung, L. R., et al. (2014). A subbasin-based framework to represent land surface processes in an Earth system model. Geoscientific Model Development, 7, 947–963.CrossRefGoogle Scholar
Tegen, I., and Fung, I. (1994). Modeling of mineral dust in the atmosphere: sources, transport, and optical thickness. Journal of Geophysical Research, 99D, 22897–22914.Google Scholar
Tegen, I., Harrison, S. P., Kohfeld, K., et al. (2002). Impact of vegetation and preferential source areas on global dust aerosol: Results from a model study. Journal of Geophysical Research, 107, 4576, doi:10.1029/2001JD000963.CrossRefGoogle 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, 3869–3887.CrossRefGoogle 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, 2986–2998.CrossRefGoogle ScholarPubMed
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, 661–677.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, 1991–2011.CrossRefGoogle 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.CrossRefGoogle Scholar
Todd-Brown, K. E. O., Randerson, J. T., Post, W. M., et al. (2013). Causes of variation in soil carbon simulations from CMIP5 Earth system models and comparison with observations. Biogeosciences, 10, 1717–1736.CrossRefGoogle 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, 10243–10269.CrossRefGoogle Scholar
Verheijen, L. M., Brovkin, V., Aerts, R., et al. (2013). Impacts of trait variation through observed trait–climate relationships on performance of an Earth system model: a conceptual analysis. Biogeosciences, 10, 5497–5515.CrossRefGoogle Scholar
Wang, Y.-P., Trudinger, C. M., and Enting, I. G. (2009). A review of applications of model–data fusion to studies of terrestrial carbon fluxes at different scales. Agricultural and Forest Meteorology, 149, 1829–1842.CrossRefGoogle 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, 2261–2282.CrossRefGoogle 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., Lu, X. J., Wright, I. J., et al. (2012). Correlations among leaf traits provide a significant constraint on the estimate of global gross primary production. Geophysical Research Letters, 39, L19405, doi:10.1029/2012GL053461.CrossRefGoogle Scholar
Wania, R., Ross, I., and Prentice, I. C. (2009). Integrating peatlands and permafrost into a dynamic global vegetation model, 1: Evaluation and sensitivity of physical land surface processes. Global Biogeochemical Cycles, 23, GB3014, doi:10.1029/2008GB003412.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, 565–584.CrossRefGoogle Scholar
Wania, R., Meissner, K. J., Eby, M., et al. (2012). Carbon–nitrogen feedbacks in the UVic ESCM. Geoscientific Model Development, 5, 1137–1160.CrossRefGoogle 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, 617–641.CrossRefGoogle Scholar
Wei, J., Dirmeyer, P. A., Guo, Z., Zhang, L., and Misra, V. (2010). How much do different land models matter for climate simulation? Part I: Climatology and variability. Journal of Climate, 23, 3120–3134.Google Scholar
Wenzel, S., Cox, P. M., Eyring, V., and Friedlingstein, P. (2014). Emergent constraints on climate-carbon cycle feedbacks in the CMIP5 Earth system models. Journal of Geophysical Research: Biogeosciences, 119, 794–807, doi:10.1002/2013JG002591.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, 3899–3917.CrossRefGoogle 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, 625–641.CrossRefGoogle Scholar
Williams, M., Richardson, A. D., Reichstein, M., et al. (2009). Improving land surface models with FLUXNET data. Biogeosciences, 6, 1341–1359.CrossRefGoogle 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, Colorado: National Center for Atmospheric Research.Google Scholar
Wilson, M. F., and Henderson-Sellers, A. (1985). A global archive of land cover and soil data for use in general circulation climate models. Journal of Climatology, 5, 119–143.CrossRefGoogle 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, 1150–1162.CrossRefGoogle Scholar
Wittig, V. E., Ainsworth, E. A., Naidu, S. L., Karnosky, D. F., and Long, S. P. (2009). Quantifying the impact of current and future tropospheric ozone on tree biomass, growth, physiology and biochemistry: A quantitative meta-analysis. Global Change Biology, 15, 396–424.CrossRefGoogle Scholar
Wood, E. F., Lettenmaier, D. P., Liang, X., et al. (1998). The Project for Intercomparison of Land-surface Parameterization Schemes (PILPS) Phase 2(c) Red–Arkansas River basin experiment, 1: Experiment description and summary intercomparisons. Global and Planetary Change, 19, 115–135.CrossRefGoogle 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, 1667–1681.CrossRefGoogle 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.CrossRefGoogle Scholar
Zaehle, S., Ciais, P., Friend, A. D., and Prieur, V. (2011). Carbon benefits of anthropogenic reactive nitrogen offset by nitrous oxide emissions. Nature Geoscience, 4, 601–605.CrossRefGoogle 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, 803–822.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle Scholar
Zeng, X., Dickinson, R. E., Walker, A., et al. (2000). Derivation and evaluation of global 1-km fractional vegetation cover data for land modeling. Journal of Applied Meteorology, 39, 826–839.2.0.CO;2>CrossRefGoogle Scholar
Zeng, X., Rao, P., DeFries, R. S., and Hansen, M. C. (2003). Interannual variability and decadal trend of global fractional vegetation cover from 1982 to 2000. Journal of Applied Meteorology, 42, 1525–1530.2.0.CO;2>CrossRefGoogle Scholar
Zhang, Q., Wang, Y. P., Pitman, A. J., and Dai, Y. J. (2011). Limitations of nitrogen and phosphorous on the terrestrial carbon uptake in the 20th century. Geophysical Research Letters, 38, L22701, doi:10.1029/2011GL049244.CrossRefGoogle Scholar
Zhang, Q., Wang, Y.P., Matear, R. J., Pitman, A. J., and Dai, Y. J. (2014). Nitrogen and phosphorous limitations significantly reduce future allowable CO2 emissions. Geophysical Research Letters, 41, 632–637, doi:10.1002/2013GL058352.CrossRefGoogle Scholar
Zhou, T., and Luo, Y. (2008). Spatial patterns of ecosystem carbon residence time and NPP-driven carbon uptake in the conterminous United States. Global Biogeochemical Cycles, 22, GB3032, doi:10.1029/2007GB002939.CrossRefGoogle Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×