Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-29T01:31:32.440Z Has data issue: false hasContentIssue false

29 - Carbon Cycle–Climate Feedbacks

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

As shown in previous chapters, numerous climate model experiments demonstrate that vegetation exerts an important feedback on climate through energy and water cycles. In addition to these biogeophysical feedbacks, terrestrial ecosystems are coupled to climate through the carbon cycle. Terrestrial ecosystems absorb a significant portion of the annual emission of CO2 to the atmosphere by human activities. This arises from enhanced photosynthesis as a result of climate change, increasing concentration of CO2 in the atmosphere, or by increasing deposition of nitrogen on land. It is also caused by regrowth of forests following abandonment of farmland. At longer timescales, changes in the biogeography of ecosystems alter carbon storage on land. Climate model simulations show that the terrestrial carbon cycle is a positive climate feedback whereby a warmer climate decreases the capacity of the terrestrial biosphere to storage anthropogenic carbon emissions. The uncertainty in the carbon cycle feedback is of comparable magnitude to the uncertainty arising from physical climate processes and relates in part to plant and microbial physiological responses to temperature and plant demographic processes in response to disturbance and climate change. In addition, nitrogen limits the productivity of many terrestrial ecosystems. This control of the carbon cycle by nitrogen results in prominent carbon–nitrogen interactions including constraints on productivity increases with CO2 enrichment, enhanced productivity with nitrogen deposition, and additional nitrogen mineralization with soil warming.

Glacial–Interglacial Cycles

In the absence of human influences, the concentration of CO2 in the atmosphere is the balance between oceanic and terrestrial processes. Over hundreds of thousands of years, atmospheric CO2 has varied by about 100 ppm during glacial–interglacial cycles (Figure 8.2). Atmospheric CO2 was lower during glacial periods (~180–200 ppm) than during interglacial periods (~270–290 ppm). The causes of this variation are unclear, but likely reflect oceanic processes (Ciais et al. 2013). In particular, the colder glacial ocean held more carbon than during warmer interglacial periods because the solubility of CO2 increases with colder temperature. Changes in ocean circulation also contributed to lower atmospheric CO2, while the lower sea level and increased salinity of the glacial ocean decreased oceanic carbon stocks.

Type
Chapter
Information
Ecological Climatology
Concepts and Applications
, pp. 563 - 593
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

Achard, F., Eva, H. D., Mayaux, P., Stibig, H.-J., and Belward, A. (2004). Improved estimates of net carbon emissions from land cover change in the tropics for the 1990s. Global Biogeochemical Cycles, 18, GB2008, doi:10.1029/2003GB002142.CrossRefGoogle Scholar
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
Allen, C. D., Macalady, A. K., Chenchouni, H., et al. (2010). A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. Forest Ecology and Management, 259, 660–684.CrossRefGoogle Scholar
Allison, S. D., and Martiny, J. B. H. (2008). Resistance, resilience, and redundancy in microbial communities. Proceedings of the National Academy of Sciences USA, 105, 11512–11519.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
Anderegg, W. R. L., Berry, J. A., Smith, D. D., et al. (2012). The roles of hydraulic and carbon stress in a widespread climate-induced forest die-off. Proceedings of the National Academy of Sciences USA, 109, 233–237.CrossRefGoogle Scholar
Arora, V. K., Boer, G. J., Friedlingstein, P., et al. (2013). Carbon–concentration and carbon–climate feedbacks in CMIP5 Earth system models. Journal of Climate, 26, 5289–5314.CrossRefGoogle Scholar
Ballantyne, A. P., Alden, C. B., Miller, J. B., Tans, P. P., and White, J. W. C. (2012). Increase in observed net carbon dioxide uptake by land and oceans during the past 50 years. Nature, 488, 70–72.CrossRefGoogle ScholarPubMed
Barichivich, J., Briffa, K. R., Myneni, R. B., et al. (2013). Large-scale variations in the vegetation growing season and annual cycle of atmospheric CO2 at high northern latitudes from 1950 to 2011. Global Change Biology, 19, 3167–3183.CrossRefGoogle ScholarPubMed
Beer, C., Reichstein, M., Tomelleri, E., et al. (2010). Terrestrial gross carbon dioxide uptake: Global distribution and covariation with climate. Science, 329, 834–838.CrossRefGoogle ScholarPubMed
Betzelberger, A. M., Yendrek, C. R., Sun, J., et al. (2012). Ozone exposure response for U.S. soybean cultivars: Linear reductions in photosynthetic potential, biomass, and yield. Plant Physiology, 160, 1827–1839.CrossRefGoogle ScholarPubMed
Boisvenue, C., and Running, S. W. (2006). Impacts of climate change on natural forest productivity – evidence since the middle of the 20th century. Global Change Biology, 12, 862–882.CrossRefGoogle 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.CrossRefGoogle Scholar
Booth, B. B. B., Jones, C. D., Collins, M., et al. (2012). High sensitivity of future global warming to land carbon cycle processes. Environmental Research Letters, 7, 024002, doi:10.1088/1748–9326/7/2/024002.CrossRefGoogle Scholar
Booth, B. B. B., Bernie, D., McNeall, D., et al. (2013). Scenario and modelling uncertainty in global mean temperature change derived from emission-driven global climate models. Earth System Dynamics, 4, 95–108.CrossRefGoogle Scholar
Brovkin, V., Lorenz, S. J., Jungclaus, J., et al. (2010). Sensitivity of a coupled climate–carbon cycle model to large volcanic eruptions during the last millennium. Tellus B, 62, 674–681.CrossRefGoogle Scholar
Brovkin, V., Boysen, L., Arora, V. K., et al. (2013). Effect of anthropogenic land-use and land-cover changes on climate and land carbon storage in CMIP5 projections for the twenty-first century. Journal of Climate, 26, 6859–6881.CrossRefGoogle Scholar
Buermann, W., Anderson, B., Tucker, C. J., et al. (2003). Interannual covariability in Northern Hemisphere air temperatures and greenness associated with El Niño–Southern Oscillation and the Arctic Oscillation. Journal of Geophysical Research, 108, 4396, doi:10.1029/2002JD002630.CrossRefGoogle Scholar
Buermann, W., Lintner, B. R., Koven, C. D., et al. (2007). The changing carbon cycle at Mauna Loa Observatory. Proceedings of the National Academy of Sciences USA, 104, 4249–4254.CrossRefGoogle ScholarPubMed
Burke, E. J., Jones, C. D., and Koven, C. D. (2013). Estimating the permafrost-carbon climate response in the CMIP5 climate models using a simplified approach. Journal of Climate, 26, 4897–4909.CrossRefGoogle Scholar
Butterbach-Bahl, K., Nemitz, E., Zaehle, S., et al. (2011). Nitrogen as a threat to the European greenhouse balance. In The European Nitrogen Assessment, ed. Sutton, M. A., Howard, C. M., Erisman, J. W., et al. Cambridge: Cambridge University Press, pp. 434–462.Google Scholar
Choat, B., Jansen, S., Brodribb, T. J., et al. (2012). Global convergence in the vulnerability of forests to drought. Nature, 491, 752–755.CrossRefGoogle ScholarPubMed
Ciais, P., Reichstein, M., Viovy, N., et al. (2005). Europe-wide reduction in primary productivity caused by the heat and drought in 2003. Nature, 437, 529–533.CrossRefGoogle ScholarPubMed
Ciais, P., Rayner, P., Chevallier, F., et al. (2010). Atmospheric inversions for estimating CO2 fluxes: Methods and perspectives. Climatic Change, 103, 69–92.CrossRefGoogle Scholar
Ciais, P., Tagliabue, A., Cuntz, M., et al. (2012). Large inert carbon pool in the terrestrial biosphere during the Last Glacial Maximum. Nature Geoscience, 5, 74–79.Google Scholar
Ciais, P., Sabine, C., Bala, G., et al. (2013). Carbon and other biogeochemical cycles. In Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, ed. Stocker, T. F., Qin, D., Plattner, G.-K., et al. Cambridge: Cambridge University Press, pp. 465–570.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, 184–187.CrossRefGoogle Scholar
Cox, P. M., Betts, R. A., Collins, M., et al. (2004). Amazonian forest dieback under climate–carbon cycle projections for the 21st century. Theoretical and Applied Climatology, 78, 137–156.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
Cramer, W., Bondeau, A., Woodward, F. I., et al. (2001). Global response of terrestrial ecosystem structure and function to CO2 and climate change: Results from six dynamic global vegetation models. Global Change Biology, 7, 357–373.CrossRefGoogle Scholar
DeFries, R. S., Houghton, R. A., Hansen, M. C., et al. (2002). Carbon emissions from tropical deforestation and regrowth based on satellite observations for the 1980s and 1990s. Proceedings of the National Academy of Sciences USA, 99, 14256–14261.CrossRefGoogle ScholarPubMed
Denman, K. L., Brasseur, G., Chidthaisong, A., et al. (2007). Couplings between changes in the climate system and biogeochemistry. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. Solomon, S., Qin, D., Manning, M., et al. Cambridge: Cambridge University Press, 499–587.Google Scholar
de Vries, W., Solberg, S., Dobbertin, M., et al. (2008). Ecologically implausible carbon response?Nature, 451, E1–E3.CrossRefGoogle ScholarPubMed
de Vries, W., Solberg, S., Dobbertin, M., et al. (2009). The impact of nitrogen deposition on carbon sequestration by European forests and heathlands. Forest Ecology and Management, 258, 1814–1823.CrossRefGoogle Scholar
Donohue, R. J., Roderick, M. L., McVicar, T. R., and Farquhar, G. D. (2013). Impact of CO2 fertilization on maximum foliage cover across the globe's warm, arid environments. Geophysical Research Letters, 40, 3031–3035, doi:10.1002/grl.50563.CrossRefGoogle Scholar
Dragoni, D., Schmid, H. P., Wayson, C. A., et al. (2011). Evidence of increased net ecosystem productivity associated with a longer vegetated season in a deciduous forest in south-central Indiana, USA. Global Change Biology, 17, 886–897.CrossRefGoogle Scholar
Felzer, B., Reilly, J., Melillo, J., et al. (2005). Future effects of ozone on carbon sequestration and climate change policy using a global biogeochemical model. Climatic Change, 73, 345–373.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
Fisk, J. P., Hurtt, G. C., Chambers, J. Q., et al. (2013). The impacts of tropical cyclones on the net carbon balance of eastern US forests (1851–2000). Environmental Research Letters, 8, 045017, doi:10.1088/1748–9326/8/4/045017.CrossRefGoogle Scholar
Flückiger, J., Monnin, E., Stauffer, B., et al. (2002). High-resolution Holocene N2O ice core record and its relationship with CH4 and CO2. Global Biogeochemical Cycles, 16, 1010, doi:10.1029/2001GB001417.CrossRefGoogle Scholar
Fog, K. (1988). The effect of added nitrogen on the rate of decomposition of organic matter. Biological Reviews, 63, 433–462.CrossRefGoogle Scholar
Frey, S. D., Knorr, M., Parrent, J. L., and Simpson, R. T. (2004). Chronic nitrogen enrichment affects the structure and function of the soil microbial community in temperate hardwood and pine forests. Forest Ecology and Management, 196, 159–171.CrossRefGoogle Scholar
Friedlingstein, P., Bopp, L., Ciais, P., et al. (2001). Positive feedback between future climate change and the carbon cycle. Geophysical Research Letters, 28, 1543–1546.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
Friend, A. D., Lucht, W., Rademacher, T. T., et al. (2014). Carbon residence time dominates uncertainty in terrestrial vegetation responses to future climate and atmospheric CO2. Proceedings of the National Academy of Sciences USA, 111, 3280–3285.CrossRefGoogle ScholarPubMed
Frölicher, T. L., Joos, F., Raible, C. C., and Sarmiento, J. L. (2013). Atmospheric CO2 response to volcanic eruptions: The role of ENSO, season, and variability. Global Biogeochemical Cycles, 27, 239–251, doi:10.1002/gbc.20028.CrossRefGoogle Scholar
Galloway, J. N., Dentener, F. J., Capone, D. G., et al. (2004). Nitrogen cycles: Past, present, and future. Biogeochemistry, 70, 153–226.CrossRefGoogle Scholar
Gasser, T., and Ciais, P. (2013). A theoretical framework for the net land-to-atmosphere CO2 flux and its implications in the definition of “emissions from land-use change”. Earth System Dynamics, 4, 171–186.CrossRefGoogle Scholar
Gerber, S., Hedin, L. O., Keel, S. G., Pacala, S. W., and Shevliakova, E. (2013). Land use change and nitrogen feedbacks constrain the trajectory of the land carbon sink. Geophysical Research Letters, 40, 5218–5222, doi:10.1002/grl.50957.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
Graven, H. D., Keeling, R. F., Piper, S. C., et al. (2013). Enhanced seasonal exchange of CO2 by northern ecosystems since 1960. Science, 341, 1085–1089.CrossRefGoogle ScholarPubMed
Gray, J. M., Frolking, S., Kort, E. A., et al. (2014). Direct human influence on atmospheric CO2 seasonality from increased cropland productivity. Nature, 515, 398–401.CrossRefGoogle ScholarPubMed
Gu, L., Baldocchi, D. D., Wofsy, S. C., et al. (2003). Response of a deciduous forest to the Mount Pinatubo eruption: Enhanced photosynthesis. Science, 299, 2035–2038.CrossRefGoogle ScholarPubMed
Gurney, K. R., and Eckels, W. J. (2011). Regional trends in terrestrial carbon exchange and their seasonal signatures. Tellus B, 63, 328–339.CrossRefGoogle Scholar
Gurney, K. R., Law, R. M., Denning, A. S., et al. (2004). Transcom 3 inversion intercomparison: Model mean results for the estimation of seasonal carbon sources and sinks. Global Biogeochemical Cycles, 18, GB1010, doi:10.1029/2003GB002111.CrossRefGoogle Scholar
Hashimoto, H., Nemani, R. R., White, M. A., et al. (2004). El Niño–Southern Oscillation-induced variability in terrestrial carbon cycling. Journal of Geophysical Research, 109, D23110, doi:10.1029/2004JD004959.CrossRefGoogle Scholar
Hicke, J. A., Allen, C. D., Desai, A. R., et al. (2012). Effects of biotic disturbances on forest carbon cycling in the United States and Canada. Global Change Biology, 18, 7–34.CrossRefGoogle Scholar
Higgins, P. A. T., and Harte, J. (2006). Biophysical and biogeochemical responses to climate change depend on dispersal and migration. BioScience, 56, 407–417.CrossRefGoogle Scholar
Hobbie, S. E. (2008). Nitrogen effects on decomposition: A five-year experiment in eight temperate sites. Ecology, 89, 2633–2644.CrossRefGoogle ScholarPubMed
Hobbie, S. E., Eddy, W. C., Buyarski, C. R., et al. (2012). Response of decomposing litter and its microbial community to multiple forms of nitrogen enrichment. Ecological Monographs, 82, 389–405.CrossRefGoogle Scholar
Hoffman, F. M., Randerson, J. T., Arora, V. K., et al. (2014). Causes and implications of persistent atmospheric carbon dioxide biases in Earth System Models. Journal of Geophysical Research: Biogeosciences, 119, 141–162, doi:10.1002/2013JG002381.Google Scholar
Högberg, P. (2007). Nitrogen impacts on forest carbon. Nature, 447, 781–782.CrossRefGoogle ScholarPubMed
Holland, E. A., Braswell, B. H., Lamarque, J.-F., et al. (1997). Variations in the predicted spatial distribution of atmospheric nitrogen deposition and their impact on carbon uptake by terrestrial ecosystems. Journal of Geophysical Research, 102D, 15849–15866.Google Scholar
Holland, E. A., Braswell, B. H., Sulzman, J., and Lamarque, J.-F. (2005). Nitrogen deposition onto the United States and western Europe: Synthesis of observations and models. Ecological Applications, 15, 38–57.CrossRefGoogle Scholar
Houghton, R. A. (2003). Revised estimates of the annual net flux of carbon to the atmosphere from changes in land use and land management 1850–2000. Tellus B, 55, 378–390.Google Scholar
Houghton, R. A. (2010). How well do we know the flux of CO2 from land-use change?Tellus B, 62, 337–351.CrossRefGoogle Scholar
Houghton, R. A. (2013). Keeping management effects separate from environmental effects in terrestrial carbon accounting. Global Change Biology, 19, 2609–2612.CrossRefGoogle ScholarPubMed
Houghton, R. A., Hobbie, J. E., Melillo, J. M., et al. (1983). Changes in the carbon content of terrestrial biota and soils between 1860 and 1980: A net release of CO2 to the atmosphere. Ecological Monographs, 53, 235–262.CrossRefGoogle Scholar
Houghton, R. A., House, J. I., Pongratz, J., et al. (2012). Carbon emissions from land use and land-cover change. Biogeosciences, 9, 5125–5142.CrossRefGoogle Scholar
Huntingford, C., Lowe, J. A., Booth, B. B. B., et al. (2009). Contributions of carbon cycle uncertainty to future climate projection spread. Tellus B, 61, 355–360.CrossRefGoogle 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, 1208–1229.CrossRefGoogle 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.CrossRefGoogle Scholar
Janssens, I. A., Dieleman, W., Luyssaert, S., et al. (2010). Reduction of forest soil respiration in response to nitrogen deposition. Nature Geoscience, 3, 315–322.CrossRefGoogle Scholar
Jeong, S.-J., Ho, C.-H., Gim, H.-J., and Brown, M. E. (2011). Phenology shifts at start vs. end of growing season in temperate vegetation over the Northern Hemisphere for the period 1982–2008. Global Change Biology, 17, 2385–2399.CrossRefGoogle Scholar
Jones, C. D., and Cox, P. M. (2001). Modeling the volcanic signal in the atmospheric CO2 record. Global Biogeochemical Cycles, 15, 453–465.CrossRefGoogle Scholar
Jones, C. D., Cox, P. M., and Huntingford, C. (2006). Climate–carbon cycle feedbacks under stabilization: Uncertainty and observational constraints. Tellus B, 58, 603–613.CrossRefGoogle Scholar
Jones, C., Robertson, E., Arora, V., et al. (2013). Twenty-first-century compatible CO2 emissions and airborne fraction simulated by CMIP5 Earth system models under four Representative Concentration Pathways. Journal of Climate, 26, 4398–4413.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
Karnosky, D. F., Pregitzer, K. S., Zak, D. R., et al. (2005). Scaling ozone responses of forest trees to the ecosystem level in a changing climate. Plant, Cell and Environment, 28, 965–981.CrossRefGoogle Scholar
Keenan, T. F., Hollinger, D. Y., Bohrer, G., et al. (2013). Increase in forest water-use efficiency as atmospheric carbon dioxide concentrations rise. Nature, 499, 324–327.CrossRefGoogle ScholarPubMed
Keenan, T. F., Gray, J., Friedl, M. A., et al. (2014). Net carbon uptake has increased through warming-induced changes in temperate forest phenology. Nature Climate Change, 4, 598–604.CrossRefGoogle Scholar
Klein Goldewijk, K. (2001). Estimating global land use change over the past 300 years: The HYDE Database. Global Biogeochemical Cycles, 15, 417–433.Google Scholar
Klein Goldewijk, K., Beusen, A., van Drecht, G., and de Vos, M. (2011). The HYDE 3.1 spatially explicit database of human-induced global land-use change over the past 12,000 years. Global Ecology and Biogeography, 20, 73–86.CrossRefGoogle Scholar
Knorr, M., Frey, S. D., and Curtis, P. S. (2005). Nitrogen additions and litter decomposition: A meta-analysis. Ecology, 86, 3252–3257.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, 108, 14769–14774.CrossRefGoogle ScholarPubMed
Koven, C. D., Riley, W. J., and Stern, A. (2013). Analysis of permafrost thermal dynamics and response to climate change in the CMIP5 Earth System Models. Journal of Climate, 26, 1877–1900.CrossRefGoogle Scholar
Kurz, W. A., Dymond, C. C., Stinson, G., et al. (2008a). Mountain pine beetle and forest carbon feedback to climate change. Nature, 452, 987–990.CrossRefGoogle ScholarPubMed
Kurz, W. A., Stinson, G., Rampley, G. J., Dymond, C. C., and Neilson, E. T. (2008b). Risk of natural disturbances makes future contribution of Canada's forests to the global carbon cycle highly uncertain. Proceedings of the National Academy of Sciences USA, 105, 1551–1555.CrossRefGoogle ScholarPubMed
Lamarque, J.-F., Bond, T. C., Eyring, V., et al. (2010). Historical (1850–2000) gridded anthropogenic and biomass burning emissions of reactive gases and aerosols: Methodology and application. Atmospheric Chemistry and Physics, 10, 7017–7039.CrossRefGoogle Scholar
Lamarque, J.-F., Kyle, G. P., Meinshausen, M., et al. (2011). Global and regional evolution of short-lived radiatively-active gases and aerosols in the Representative Concentration Pathways. Climatic Change, 109, 191–212.CrossRefGoogle Scholar
Lamarque, J.-F., Dentener, F., McConnell, J., et al. (2013). Multi–model mean nitrogen and sulfur deposition from the Atmospheric Chemistry and Climate Model Intercomparison Project (ACCMIP): Evaluation of historical and projected future changes. Atmospheric Chemistry and Physics, 13, 7997–8018.CrossRefGoogle Scholar
Lambert, F. H., Harris, G. R., Collins, M., et al. (2013). Interactions between perturbations to different Earth system components simulated by a fully-coupled climate model. Climate Dynamics, 41, 3055–3072.CrossRefGoogle Scholar
Lawrence, D. M., and Slater, A. G. (2005). A projection of severe near-surface permafrost degradation during the 21st century. Geophysical Research Letters, 32, L24401, doi:10.1029/2005GL025080.CrossRefGoogle Scholar
LeBauer, D. S., and Treseder, K. K. (2008). Nitrogen limitation of net primary productivity in terrestrial ecosystems is globally distributed. Ecology, 89, 371–379.CrossRefGoogle ScholarPubMed
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
Le Quéré, C., Andres, R. J., Boden, T., et al. (2013). The global carbon budget 1959–2011. Earth System Science Data, 5, 165–185.CrossRefGoogle Scholar
Levis, S. (2010). Modeling vegetation and land use in models of the Earth System. WIREs Climate Change, 1, 840–856.CrossRefGoogle Scholar
Lewis, S. L., Lopez-Gonzalez, G., Sonké, B., et al. (2009). Increasing carbon storage in intact African tropical forests. Nature, 457, 1003–1006.CrossRefGoogle ScholarPubMed
Lewis, S. L., Brando, P. M., Phillips, O. L., van der Heijden, G. M. F., and Nepstad, D. (2011). The 2010 Amazon drought. Science, 331, 554.CrossRefGoogle ScholarPubMed
Liu, L., and Greaver, T. L. (2009). A review of nitrogen enrichment effects on three biogenic GHGs: The CO2 sink may be largely offset by stimulated N2O and CH4 emission. Ecology Letters, 12, 1103–1117.CrossRefGoogle ScholarPubMed
Liu, X., Zhang, Y., Han, W., et al. (2013). Enhanced nitrogen deposition over China. Nature, 494, 459–462.CrossRefGoogle ScholarPubMed
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
Lombardozzi, D., Bonan, G. B., and Nychka, D. W. (2014). The emerging anthropogenic signal in land–atmosphere carbon-cycle coupling. Nature Climate Change, 4, 796–800.CrossRefGoogle Scholar
Lombardozzi, D., Levis, S., Bonan, G., Hess, P. G., and Sparks, J. P. (2015). The influence of chronic ozone exposure on global carbon and water cycles. Journal of Climate, 28, 292–305.CrossRefGoogle Scholar
Luo, Y. (2007). Terrestrial carbon-cycle feedback to climate warming. Annual Review of Ecology, Evolution, and Systematics, 38, 683–712.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
MacFarling Meure, C., Etheridge, D., Trudinger, C., et al. (2006). Law Dome CO2, CH4 and N2O ice core records extended to 2000 years BP. Geophysical Research Letters, 33, L14810, doi:10.1029/2006GL026152.CrossRefGoogle Scholar
Magnani, F., Mencuccini, M., Borghetti, M., et al. (2007). The human footprint in the carbon cycle of temperate and boreal forests. Nature, 447, 849–851.CrossRefGoogle ScholarPubMed
Matthews, H. D. (2005). Decrease of emissions required to stabilize atmospheric CO2 due to positive carbon cycle–climate feedbacks. Geophysical Research Letters, 32, L21707, doi:10.1029/2005GL023435.CrossRefGoogle Scholar
Matthews, H. D. (2006). Emissions targets for CO2 stabilization as modified by carbon cycle feedbacks. Tellus B, 58, 591–602.CrossRefGoogle Scholar
McDowell, N. G., and Sevanto, S. (2010). The mechanisms of carbon starvation: How, when, or does it even occur at all?New Phytologist, 186, 264–266.CrossRefGoogle ScholarPubMed
McDowell, N., Pockman, W. T., Allen, C. D., et al. (2008). Mechanisms of plant survival and mortality during drought: Why do some plants survive while others succumb to drought?New Phytologist, 178, 719–739.CrossRefGoogle ScholarPubMed
McDowell, N. G., Fisher, R. A., Xu, C., et al. (2013). Evaluating theories of drought-induced vegetation mortality using a multimodel–experiment framework. New Phytologist, 200, 304–321.CrossRefGoogle ScholarPubMed
Menzel, A., Sparks, T. H., Estrella, N., et al. (2006). European phenological response to climate change matches the warming pattern. Global Change Biology, 12, 1969–1976.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
Metsaranta, J. M., Kurz, W. A., Neilson, E. T., and Stinson, G. (2010). Implications of future disturbance regimes on the carbon balance of Canada's managed forest (2010–2100). Tellus B, 62, 719–728.Google Scholar
Monnin, E., Steig, E. J., Siegenthaler, U., et al. (2004). Evidence for substantial accumulation rate variability in Antarctica during the Holocene, through synchronization of CO2 in the Taylor Dome, Dome C and DML ice cores. Earth and Planetary Science Letters, 224, 45–54.CrossRefGoogle Scholar
Morgan, P. B., Mies, T. A., Bollero, G. A., Nelson, R. L., and Long, S. P. (2006). Season-long elevation of ozone concentration to projected 2050 levels under fully open-air conditions substantially decreases the growth and production of soybean. New Phytologist, 170, 333–343.CrossRefGoogle ScholarPubMed
Myneni, R. B., Keeling, C. D., Tucker, C. J., Asrar, G., and Nemani, R. R. (1997). Increased plant growth in the northern high latitudes from 1981 to 1991. Nature, 386, 698–702.CrossRefGoogle Scholar
Nadelhoffer, K. J., Emmett, B. A., Gundersen, P., et al. (1999). Nitrogen deposition makes a minor contribution to carbon sequestration in temperate forests. Nature, 398, 145–148.CrossRefGoogle Scholar
Negrón-Juárez, R., Baker, D. B., Zeng, H., Henkel, T. K., and Chambers, J. Q. (2010). Assessing hurricane‐induced tree mortality in U.S. Gulf Coast forest ecosystems. Journal of Geophysical Research, 115, G04030, doi:10.1029/2009JG001221.CrossRefGoogle Scholar
Nemani, R. R., Keeling, C. D., Hashimoto, H., et al. (2003). Climate-driven increases in global terrestrial net primary production from 1982 to 1999. Science, 300, 1560–1563.CrossRefGoogle ScholarPubMed
Norby, R. J., and Zak, D. R. (2011). Ecological lessons from free-air CO2 enrichment (FACE) experiments. Annual Review of Ecology, Evolution, and Systematics, 42, 181–203.CrossRefGoogle Scholar
Ollinger, S. V., Aber, J. D., Reich, P. B., and Freuder, R. J. (2002). Interactive effects of nitrogen deposition, tropospheric ozone, elevated CO2 and land use history on the carbon dynamics of northern hardwood forests. Global Change Biology, 8, 545–562.CrossRefGoogle Scholar
Page, S. E., Siegert, F., Rieley, J. O., et al. (2002).The amount of carbon released from peat and forest fires in Indonesia during 1997. Nature, 420, 61–65.CrossRefGoogle ScholarPubMed
Pan, Y., Birdsey, R., Hom, J., and McCullough, K. (2009). Separating effects of changes in atmospheric composition, climate and land-use on carbon sequestration of U.S. Mid-Atlantic temperate forests. Forest Ecology and Management, 259, 151–164.CrossRefGoogle Scholar
Pan, Y., Birdsey, R. A., Fang, J., et al. (2011). A large and persistent carbon sink in the world's forests. Science, 333, 988–993.CrossRefGoogle ScholarPubMed
Peylin, P., Law, R. M., Gurney, K. R., et al. (2013). Global atmospheric carbon budget: Results from an ensemble of atmospheric CO2 inversions. Biogeosciences, 10, 6699–6720.CrossRefGoogle Scholar
Phillips, O. L., Malhi, Y., Higuchi, N., et al. (1998). Changes in the carbon balance of tropical forests: Evidence from long-term plots. Science, 282, 439–442.CrossRefGoogle ScholarPubMed
Phillips, O. L., Aragão, L. E. O. C., Lewis, S. L., et al. (2009). Drought sensitivity of the Amazon rainforest. Science, 323, 1344–1347.CrossRefGoogle ScholarPubMed
Piao, S., Friedlingstein, P., Ciais, P., Zhou, L., and Chen, A. (2006). Effect of climate and CO2 changes on the greening of the Northern Hemisphere over the past two decades. Geophysical Research Letters, 33, L23402, doi:10.1029/2006GL028205.CrossRefGoogle Scholar
Piao, S., Ciais, P., Friedlingstein, P., et al. (2008). Net carbon dioxide losses of northern ecosystems in response to autumn warming. Nature, 451, 49–52.CrossRefGoogle ScholarPubMed
Piao, S., Ciais, P., Friedlingstein, P., et al. (2009). Spatiotemporal patterns of terrestrial carbon cycle during the 20th century. Global Biogeochemical Cycles, 23, GB4026, doi:10.1029/2008GB003339.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
Pinder, R. W., Bettez, N. D., Bonan, G. B., et al. (2013). Impacts of human alteration of the nitrogen cycle in the US on radiative forcing. Biogeochemistry, 114, 25–40.CrossRefGoogle Scholar
Pongratz, J., Reick, C., Raddatz, T., and Claussen, M. (2008). A reconstruction of global agricultural areas and land cover for the last millennium. Global Biogeochemical Cycles, 22, GB3018, doi:10.1029/2007GB003153.CrossRefGoogle Scholar
Pongratz, J., Reick, C. H., Raddatz, T., and Claussen, M. (2009). Effects of anthropogenic land cover change on the carbon cycle of the last millennium. Global Biogeochemical Cycles, 23, GB4001, doi:10.1029/2009GB003488.CrossRefGoogle Scholar
Pongratz, J., Reick, C. H., Houghton, R. A., and House, J. I. (2014). Terminology as a key uncertainty in net land use and land cover change carbon flux estimates. Earth System Dynamics, 5, 177–195.CrossRefGoogle Scholar
Prentice, I. C., Harrison, S. P., and Bartlein, P. J. (2011). Global vegetation and terrestrial carbon cycle changes after the last ice age. New Phytologist, 189, 988–998.CrossRefGoogle ScholarPubMed
Ramankutty, N., and Foley, J. A. (1999). Estimating historical changes in global land cover: Croplands from 1700 to 1992. Global Biogeochemical Cycles, 13, 997–1027.CrossRefGoogle Scholar
Randerson, J. T., Thompson, M. V., Conway, T. J., Fung, I. Y., and Field, C. B. (1997). The contribution of terrestrial sources and sinks to trends in the seasonal cycle of atmospheric carbon dioxide. Global Biogeochemical Cycles, 11, 535–560.CrossRefGoogle Scholar
Randerson, J. T., Field, C. B., Fung, I. Y., and Tans, P. P. (1999). Increases in early season ecosystem uptake explain recent changes in the seasonal cycle of atmospheric CO2 at high northern latitudes. Geophysical Research Letters, 26, 2765–2768.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
Raupach, M. R., Canadell, J. G., and Le Quéré, C. (2008). Anthropogenic and biophysical contributions to increasing atmospheric CO2 growth rate and airborne fraction. Biogeosciences, 5, 1601–1613.CrossRefGoogle Scholar
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
Reichstein, M., Bahn, M., Ciais, P., et al. (2013). Climate extremes and the carbon cycle. Nature, 500, 287–295.CrossRefGoogle ScholarPubMed
Richardson, A. D., Black, T. A., Ciais, P., et al. (2010). Influence of spring and autumn phenological transitions on forest ecosystem productivity. Philosophical Transactions of the Royal Society B, 365, 3227–3246.CrossRefGoogle ScholarPubMed
Sarmiento, J. L., Gloor, M., Gruber, N., et al. (2010). Trends and regional distributions of land and ocean carbon sinks. Biogeosciences, 7, 2351–2367.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
Schimel, D., Stephens, B. B., and Fisher, J. B. (2015). Effect of increasing CO2 on the terrestrial carbon cycle. Proceedings of the National Academy of Sciences USA, 112, 436–441.CrossRefGoogle ScholarPubMed
Schuur, E. A. G., Bockheim, J., Canadell, J. G., et al. (2008). Vulnerability of permafrost carbon to climate change: Implications for the global carbon cycle. BioScience, 58, 701–714.CrossRefGoogle Scholar
Schuur, E. A. G., Abbott, B. W., Bowden, W. B., et al. (2013). Expert assessment of vulnerability of permafrost carbon to climate change. Climatic Change, 119, 359–374.CrossRefGoogle Scholar
Schwartz, M. D., Ahas, R., and Aasa, A. (2006). Onset of spring starting earlier across the Northern Hemisphere. Global Change Biology, 12, 343–351.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
Shevliakova, E., Stouffer, R. J., Malyshev, S., et al. (2013). Historical warming reduced due to enhanced land carbon uptake. Proceedings of the National Academy of Sciences USA, 110, 16730–16735.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., and Lawrence, D. M. (2013). Diagnosing present and future permafrost from climate models. Journal of Climate, 26, 5608–5623.CrossRefGoogle Scholar
Slayback, D. A., Pinzon, J. E., Los, S. O., and Tucker, C. J. (2003). Northern hemisphere photosynthetic trends 1982–99. Global Change Biology, 9, 1–15.CrossRefGoogle Scholar
Smith, N. G., and Dukes, J. S. (2013). Plant respiration and photosynthesis in global-scale vegetation models: Incorporating acclimation to temperature and CO2. Global Change Biology, 19, 45–63.CrossRefGoogle ScholarPubMed
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, 3776–3796.CrossRefGoogle Scholar
Stephens, B. B., Gurney, K. R., Tans, P. P., et al. (2007). Weak northern and strong tropical land carbon uptake from vertical profiles of atmospheric CO2. Science, 316, 1732–1735.CrossRefGoogle ScholarPubMed
Sutton, M. A., Simpson, D., Levy, P. E., et al. (2008). Uncertainties in the relationship between atmospheric nitrogen deposition and forest carbon sequestration. Global Change Biology, 14, 2057–2063.CrossRefGoogle Scholar
Thomas, R. Q., Canham, C. D., Weathers, K. C., and Goodale, C. L. (2010). Increased tree carbon storage in response to nitrogen deposition in the US. Nature Geoscience, 3, 13–17.Google Scholar
Thomas, R. Q., Bonan, G. B., and Goodale, C. L. (2013). 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
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
Thornton, P. E., Doney, S. C., Lindsay, K., et al. (2009). Carbon–nitrogen interactions regulate climate–carbon cycle feedbacks: Results from an atmosphere–ocean general circulation model. Biogeosciences, 6, 2099–2120.CrossRefGoogle Scholar
Tucker, C. J., Fung, I. Y., Keeling, C. D., and Gammon, R. H. (1986). Relationship between atmospheric CO2 variations and a satellite-derived vegetation index. Nature, 319, 195–199.CrossRefGoogle Scholar
van der Molen, M. K., Dolman, A. J., Ciais, P., et al. (2011). Drought and ecosystem carbon cycling. Agricultural and Forest Meteorology, 151, 765–773.CrossRefGoogle Scholar
van der Sleen, P., Groenendijk, P., Vlam, M., et al. (2015). No growth stimulation of tropical trees by 150 years of CO2 fertilization but water-use efficiency increased. Nature Geoscience, 8, 24–28.CrossRefGoogle Scholar
van der Werf, G. R., Randerson, J. T., Collatz, G. J., et al. (2004). Continental-scale partitioning of fire emissions during the 1997 to 2001 El Niño/La Niña period. Science, 303, 73–76.CrossRefGoogle ScholarPubMed
van der Werf, G. R., Dempewolf, J., Trigg, S. N., et al. (2008). Climate regulation of fire emissions and deforestation in equatorial Asia. Proceedings of the National Academy of Sciences USA, 105, 20350–20355.CrossRefGoogle ScholarPubMed
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, 11707–11735.CrossRefGoogle Scholar
Vitousek, P. M., Porder, S., Houlton, B. Z., and Chadwick, O. A. (2010). Terrestrial phosphorus limitation: Mechanisms, implications, and nitrogen–phosphorus interactions. Ecological Applications, 20, 5–15.CrossRefGoogle ScholarPubMed
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
Williams, C. A., Collatz, G. J., Masek, J., and Goward, S. N. (2012). Carbon consequences of forest disturbance and recovery across the conterminous United States. Global Biogeochemical Cycles, 26, GB1005, doi:10.1029/2010GB003947.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
Wolkovich, E. M., Cook, B. I., Allen, J. M., et al. (2012). Warming experiments underpredict plant phenological responses to climate change. Nature, 485, 494–497.CrossRefGoogle ScholarPubMed
Wu, C., Chen, J. M., Black, T. A., et al. (2013). Interannual variability of net ecosystem productivity in forests is explained by carbon flux phenology in autumn. Global Ecology and Biogeography, 22, 994–1006.CrossRefGoogle Scholar
Xia, J., and Wan, S. (2008). Global response patterns of terrestrial plant species to nitrogen addition. New Phytologist, 179, 428–439.CrossRefGoogle ScholarPubMed
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. (2013). Terrestrial nitrogen–carbon cycle interactions at the global scale. Philosophical Transactions of the Royal Society B, 368, 20130125, doi:10.1098/rstb.2013.0125.CrossRefGoogle ScholarPubMed
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, 311–320.CrossRefGoogle Scholar
Zaehle, S., Friedlingstein, P., and Friend, A. D. (2010a). Terrestrial nitrogen feedbacks may accelerate future climate change. Geophysical Research Letters, 37, L01401, doi:10.1029/2009GL041345.CrossRefGoogle Scholar
Zaehle, S., Friend, A. D., Friedlingstein, P., et al. (2010b). 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.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
Zeng, H. C., Chambers, J. Q., Negron-Juarez, R. I., et al. (2009). Impacts of tropical cyclones on US forest tree mortality and carbon flux from 1851 to 2000. Proceedings of the National Academy of Sciences USA, 106, 7888–7892.CrossRefGoogle Scholar
Zeng, N., Zhao, F., Collatz, G. J., et al. (2014). Agricultural Green Revolution as a driver of increasing atmospheric CO2 seasonal amplitude. Nature, 515, 394–397.CrossRefGoogle ScholarPubMed
Zhang, L., Jacob, D. J., Knipping, E. M., et al. (2012). Nitrogen deposition to the United States: Distribution, sources, and processes. Atmospheric Chemistry and Physics, 12, 4539–4554.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
Zhao, M., and Running, S. W. (2010). Drought-induced reduction in global terrestrial net primary production from 2000 through 2009. Science, 329, 940–943.CrossRefGoogle ScholarPubMed
Zscheischler, J., Mahecha, M. D., von Buttlar, J., et al. (2014). A few extreme events dominate global interannual variability in gross primary production. Environmental Research Letters, 9, 035001, doi:10.1088/1748–9326/9/3/035001.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.

  • Carbon Cycle–Climate Feedbacks
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Ecological Climatology
  • Online publication: 05 November 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781107339200.030
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.

  • Carbon Cycle–Climate Feedbacks
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Ecological Climatology
  • Online publication: 05 November 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781107339200.030
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.

  • Carbon Cycle–Climate Feedbacks
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Ecological Climatology
  • Online publication: 05 November 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781107339200.030
Available formats
×