Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-cjp7w Total loading time: 0 Render date: 2024-07-06T08:15:26.586Z Has data issue: false hasContentIssue false

7 - Knowledge Base for Forests in Cooling and Warming

Published online by Cambridge University Press:  22 June 2020

William J. Manning
Affiliation:
University of Massachusetts, Amherst
Get access

Summary

Previous chapters considered the nature and cause of global warming; the key role of carbon dioxide; the importance of the biogeochemical factors photosynthesis and BVOCs, and the biogeophysical factors albedo, evapotranspiration, and ozone; and how their interactions affect forests, atmospheric temperatures, and climate. Warming temperatures, forest fires, insect infestations, drought, deforestation and land-use change, latitudinal forest locations, and species composition all affect these interactions.

Type
Chapter
Information
Trees and Global Warming
The Role of Forests in Cooling and Warming the Atmosphere
, pp. 217 - 279
Publisher: Cambridge University Press
Print publication year: 2020

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

Alkama, R. and Cescatti, A. 2016. Biophysical climate impacts of recent changes in global forest cover. Science 351: 600603.Google Scholar
Arneth, A., Sitch, S., Pongratz, J. et al. 2017. Historical carbon dioxide emissions caused by land-use changes are possibly larger than assumed. Nature Geoscience 10: 7984. doi: 10.1038/NGEO2882.CrossRefGoogle Scholar
Arp, W. J. 1991. Effects of source–sink relations on photosynthetic activity acclimation to elevated CO2. Plant, Cell and Environment 14: 869875.CrossRefGoogle Scholar
Ashraf, M. I., Meng, F.-R., Bourque, C. P.-A. and MacLean, D. A. 2015. A novel modelling approach for predicting forest growth and yield under climate change PLoS One 10: e0132066. doi: 10.1371/journal.pone.132066.CrossRefGoogle ScholarPubMed
Asshoff, R., Zotz, G. and Korner, C. 2006. Growth and phenology of mature temperate forest trees in elevated CO2. Global Change Biology 12: 848861. doi: 10.1111/j.1365-2486.01133.x.CrossRefGoogle Scholar
Baccini, A., Walker, W., Carvalho, L. et al. 2017. Tropical forests are a net carbon source based on aboveground measurements of gain and loss. Science. doi: 10.1126/scienceaam5962.Google Scholar
Bader, M. K.-F., Leuzinger, S., Keel, S. G. et al. 2013. Central European hardwood trees in high-CO2 future: synthesis of an 8-year forest canopy CO2 enrichment project. Journal of Ecology 101: 15091515. doi: 10.1111/1365-2754.12149.CrossRefGoogle Scholar
Bala, G., Caldeira, K., Wickett, M. et al. 2007. Combined climate and carbon-cycle effects of large-scale deforestation. Proceedings of the National Academy of Sciences. doi: 10.1073/pnas.0608998104.Google Scholar
Baldocchi, D. D. 2003. Assessing the eddy covariance technique for evaluating carbon dioxide exchange rates of ecosystems: past, present and future. Global Change Biology 9: 114.CrossRefGoogle Scholar
Baldocchi, D. D. 2014. Measuring fluxes of trace gases and energy between ecosystems and the atmosphere – the state and future of the eddy covariance method. Global Change Biology 20: 36003609. doi: 10.1111/gcb.12649.Google Scholar
Ban-Weiss, G. A., Bala, G., Cao, L., Pongratz, J. and Caldeira, K. 2011. Climate forcing and response to idealized changes in surface latent and sensible heat. Environmental Research Letters 6: 034032.Google Scholar
Bazzaz, F. A., Miao, S. L. and Wayne, P. M. 1993. CO2-induced growth enhancements of co-occurring tree species decline at different rates. Oecologia 96: 478482.Google Scholar
Betts, A. K. and Ball, J. H. 1997. Albedo over the boreal forest. Journal of Geophysical Research 102: 2890128909.Google Scholar
Betts, R. A. 2000. Offset of the potential carbon sink from boreal forestation by decreases in albedo. Nature 408: 187190.Google Scholar
Bochenek, Z., Ziolkowski, D., Bartold, M., Orlowska, K. and Ochtra, A. 2018. Monitoring forest diversity and the impact of climate on forest environment using high-resolution satellite images. European Journal of Remote Sensing 51: 166181. doi: 10.1080/22797254.2017.1414773.Google Scholar
Bonan, G. B. 2008. Forests and climate change: forcings, feedbacks, and the climate benefits of forests. Science 320: 14441489.Google Scholar
Bortier, K., DeTemmerman, L. and Ceulemans, R. 2000. Effects of ozone exposure in open-top chambers on poplar (Populus nigra) and beech (Fagus sylvatica): a comparison. Environmental Pollution 109: 509516.CrossRefGoogle ScholarPubMed
Bowman, D. M. S., Brienen, R. J. W., Gloor, E., Phillips, O. L. and Prior, L. D. 2013. Detecting trends in tree growth: not so simple. Trends in Plant Science 18: 1117.Google Scholar
Brienen, R. J. W., Phillips, O. L., Felpausch, T. R. et al. 2015. Long-term decline of the Amazon carbon sink. Nature 519: 344348. doi: 10.1038/nature14283.Google Scholar
Bugmann, H. and Bigler, C. 2011. Will CO2 fertilization effect in the forests be offset by reduced tree longevity? Oecologia 165: 533544. doi: 10.1007/s00442-010-1837-4.Google Scholar
Burba, G., Madsen, R. and Feese, K. 2013. Eddy covariance method for CO2 emission measurements in CCUS applications: principles, instrumentation and software. Energy Procedia 40: 329336.Google Scholar
Burton, A. J. 2014. Final Technical Report for the final phase of the AspenFACE Project. www.osti.gov/biblio/1135783 (accessed 03/04/2018).Google Scholar
Butler, S. M., Melio, J. M., Johnson, J. E. et al. 2102 Soil warming alters nitrogen cycling in a New England forest: implications for ecosystem function and structure. Oecologia 168: 819828.Google Scholar
Calatayud, V., Cervo, J. and Sanz, M. J. 2007. Foliar, physiological and growth responses of four maple species exposed to ozone. Water, Air and Soil Pollution 185: 239254 doi: 10.1007/s11270-007-9446-5.Google Scholar
Camarero, J. J., Gazol, A., Galvan, J. D., Sangues-Barreda, G. and Guttiererez, E. 2015. Disparate effects of global-change drivers on mountain conifer forests; warming-induced growth enhancement in young trees vs. CO2 fertilization in old trees in wet sites. Global Change Biology 21: 738749. doi: 10.1111/gcb.12687.Google Scholar
Campioli, M., Malhi, Y., Vicca, S. et al. 2016. Evaluating the convergence between eddy-covariance and biometric methods for assessing carbon budgets of forests. Nature Communications 7: 13717. doi: 10.1038/ncomms13717.Google Scholar
Cao, l., Bala, G., Caldeira, K., Nemani, R. and Ban-Weiss, 2010. Importance of carbon dioxide physical forcing to future climate change. Proceedings of the National Academy of Sciences 107: 95139518. www.pnas.org/cgi/doi/10.1073/pnas.0913000107.Google Scholar
Carriero, G., Emiliani, G., Giovannelli, A. et al. 2015. Effects of long-term ambient ozone exposure on biomass and wood traits in poplar treated with ethylenediurea (EDU). Environmental Pollution 206: 575581. doi: 10.1016/j.envpol.2015.08.014.Google Scholar
Christ, R. 2014 . IPCC Fifth Assessment Report: The Role of Forests. IPCC. www.unece.org.Google Scholar
Chung, H., Muraoka, H., Nakamura, M. et al. 2013. Experimental warming studies on tree species and forest ecosystems: A literature review. Plant Research 126: 447460. doi: 10.1007/s10265-013-0565-3.Google Scholar
Climate Focus. 2015. Forest and land use in the Paris Agreement. www.climatefocus.com.Google Scholar
Cole, C. T., Anderson, J. E., Lindroth, R. L. and Walker, D. M. 2010. Rising concentrations of atmospheric CO2 have increased growth in natural stands of quaking aspen (Populus tremuloides). Global Change Biology 16: 21862197. doi: 10.1111/j.1365-2486.2009.02103.x.CrossRefGoogle Scholar
Crowther, T. W., Anderson, J. E. and Bradford, M. A. 2015. Tree mapping density at a global scale. Nature 525: 201205. doi: 10.1038/nature14967.Google Scholar
Davis, D. D. and Skelly, J. M. 1992. Foliar sensitivity of eight eastern hardwood tree species to ozone. Water, Air and Soil Pollution 62: 269277.CrossRefGoogle Scholar
de Dios, V., Mereed, T. E., Ferrio, J. P., Tissue, D. T. and Voltas, J. 2016. Intraspecific variation in juvenile tree growth under elevated CO2 alone or with O3: a meta-analysis. Tree Physiology 36: 682693. doi: 10.1093/treephys/tpw026.Google Scholar
Doutriaux-Boucher, M., Webb, M. J., Gregory, J. M. and Boucher, O. 2009. Carbon dioxide induced stomatal closure increases radiative forcing via rapid reduction in low cloud. Geophysical Research Letters 36: L02703. doi: 10.1020/2008GK036273.2009.Google Scholar
Drake, J. E., Macdonald, C. A., Tjoelker, M. G. et al. 2016. Short-term carbon cycling responses of a mature eucalypt woodland to gradual stepwise enrichment of atmospheric CO2 concentration. Global Change Biology 22: 380390. doi: 10.1111/gcb.13109.Google Scholar
Ellison, D., Morris, C. E., Locatelli, B. et al. 2017. Trees, forests and water: Cool insights for a hot world. Global Environmental Change 43: 5161. doi: 10.1016/j.glovencha.2017.02.0002.Google Scholar
Frank, D. C., Poulter, B., Saurer, M. et al. 2015. Water-use efficiency and transpiration across European forests during the Anthropocene. Nature Climate Change 5 : 579583. doi: 10.1038/nclimate2614.Google Scholar
Gazol, A., Camarero, J.J., Gutieirrez, E. et al. 2015. Distinct effects of climate warming on populations of silver fir (Abies alba) across Europe. Journal of Biogeography 42: 11501162. doi: 10.1111/jbi.12512.CrossRefGoogle Scholar
Gerhart, L. M. and Ward, J. K. 2010 Plant responses to low [CO2] of the past. New Phytologist 188: 674695. doi: 10.1111/j.1469-8137.2010.03441.x.Google Scholar
Gibbard, S., Caldeira, K., Bala, G., Phillips, T. J. and Wickett, M. 2005. Climate effects of global land cover change. Geophysical Research Letters 32: L23705. doi: 10.1029/2005GL0245550.2005.Google Scholar
Gimeno, T. E., McVicar, T. R., O’Grady, A. P., Tissue, D. T. and Ellsworth, D. S. 2018. Elevated CO2 did not affect the hydrological balance of a mature Eucalyptus woodland. Global Change Biology. doi: 10.1111/gcb.14139.Google Scholar
Global Forest Watch. 2015. New satellite data reveal massive tree cover loss in Russia and Canada. www.wri.org/news/2015/04/release-new-satellite-data-reveal-massive-tree-loss-russia-and-canada (accessed 07/03/2018).Google Scholar
Groenendijk, P., van der Sleen, P., Vlam, M. et al. 2015. No evidence for consistent long-term growth stimulation of 13 tropical tree species: results from tree-ring analysis. Global Change Biology 21: (10). https://doi.org/10.1111/gcb.12955.CrossRefGoogle ScholarPubMed
Grulke, N. E. and Miller, P. R. 1994. Changes in gas exchange characteristics during the life span of giant sequoia: implications for response to current and future concentrations of atmospheric ozone. Tree Physiology 14: 659668.Google Scholar
Hall, M., Rantfors, M., Slaney, M., Linder, S. and Wallin, G. 2009. Carbon dioxide exchange of buds and developing shoots of boreal Norway spruce exposed to elevated or ambient CO2 concentration and temperature in whole tree chambers. Tree Physiology 29: 467481. doi: 10.1093/treephys/tpn047.Google Scholar
Hattenschwiler, S. and Korner, C. 2000. Tree seedling responses to in situ CO2 enrichment differ among species and depend on understory light availability. Global Change Biology 6: 213226.Google Scholar
Haworth, M., Hoshika, Y. and Killi, D. 2016. Has the impact of rising CO2 on plants been exaggerated by meta-analysis of free air CO2 enrichment studies? Frontiers in Plant Science 7: article 1153.CrossRefGoogle ScholarPubMed
Heagle, A. S., Body, D. E. and Heck, W. W. 1973. An open-top field chamber to assess the impact of air pollution on plants. Journal of Environmental Quality 2: 365368.Google Scholar
Hickler, T., Smith, B., Prentice, C. et al. 2008. CO2 fertilization in temperate FACE experiments is not representative of boreal and tropical forests. Global Change Biology 14: 15311542. doi: 10.1111/j.1365-2486.2008.01598.x.Google Scholar
Holmes, C. D. 2014. Air pollution and forest water use. Nature 507: E1E2. doi: 10.1038/nature13113.Google Scholar
Hopkinson, C., Chasmer, L., Barr, A. G. et al. 2016. Monitoring boreal forest biomass and carbon storage change by integrating airborne laser scanning, biometry and eddy covariance data. Remote Sensing of the Environment 181: 8295. doi: 10.1016/jrse2016.04.010.Google Scholar
Hoshika, Y., Pecori, F., Conese, I. et al. 2013. Effects of a three-year exposure to ambient ozone on biomass allocation in poplar using ethylenediurea. Environmental Pollution 189: 299303.Google Scholar
Iverson, L., Prasad, A. and Matthews, S. 2008. Modeling potential climate change impacts on the trees of the northeastern United States. Mitigation and Adaptation Strategies for Global Change 13: 517540. doi: 10.1007/s11027-007-9129-y.Google Scholar
Jasechko, S., Sharp, Z. D., Gibson, J. L. et al. 2013. Terrestrial water fluxes dominated by transpiration. Nature 496: 347350. doi: 10.1038/nature11983.Google Scholar
Karnosky, D., Tallis, M., Darbah, J. and Taylor, G. 2007. Direct effects of elevated CO2 on forest tree productivity. In: Forestry and Climate Change, eds Innes, J., Hickey, G. and Hoen, H.. Cambridge, MA: CABI Publishing, pp. 136142.Google Scholar
Keane, K. D. and Manning, W. J. 1988. Effects of ozone and simulated acid rain on birch seedling growth and mycorrhizal associations. Environmental Pollution 52: 5565.Google Scholar
Keeling, R. 2008. Recording Earth’s vital signs. Science 319: 17711772. doi: 1.1126/science.1156761.Google Scholar
Keenan, T. F., Hollinger, D. Y., Bohrer, G. et al. 2013. Increase in forest water-use efficiency as atmospheric carbon dioxide concentrations rise. Nature 499. doi: 10.1038/nature 12291.Google Scholar
Keenan, T. F., Hollinger, D. Y., Bohreer, G. et al. 2014. Reply to Holmes: Comment on ‘Increase in forest water-use efficiency as atmospheric carbon dioxide concentrations rise’ [Nature. 2013] ‘Air pollution and forest water use’ [Nature. 2014]. Nature 507: E2E3. doi: 10.1038/nature13114.Google Scholar
King, J. S., Kubiske, M. E., Pregitzer, K. S. et al. 2005. Tropospheric O3 compromises net primary production in young stands of trembling aspen, paper birch, and sugar maple in response to atmospheric CO2. New Phytologist 168: 623636.Google Scholar
Kirschbaum, M. U. F. and Lambie, S. M. 2015. Re-analysis of plant CO2 responses during the exponential growth phase: interactions with light, temperature, nutrients and water availability. Functional Plant Biology 42: 9891000. doi: 10.1071/FP15103.Google Scholar
Kirschbaum, M. U. F. and McMillan, A. M. S. 2018. Warming and elevated CO2 have opposing influences on transpiration. Which is more important? Current Forestry Reports 4: 5171. doi: 10.1007/s40725–018-0073-8.Google Scholar
Klein, T., Bader, M.-F., Leuzinger, S. et al. 2016. Growth and carbon relations of mature Picea abies trees under 5 years of free-air CO2 enrichment. Journal of Ecology 104: 17201733.Google Scholar
Korner, C. 2017. A matter of tree longevity. Science 355: 130131. doi: 10.1126/science.aaa/2449.Google Scholar
Kovenock, M. and Swann, A. L. S. 2018. Leaf trait acclimation amplifies simulated climate warming in response to elevated carbon dioxide. Global Biogeochemical Cycles 32. doi: 10.1029/2018GB005883.Google Scholar
Kubiske, M. E., Quinn, V. S., Marquardt, P. E. and Karnosky, D. F. 2007. Effects of elevated CO2 and/or O3 on intra- and interspecific competitive ability of aspen. Plant Biology 9: 342355. doi: 10.1055/s-2006-924760.Google Scholar
Lamba, S., Hall, M., Rantfors, M. et al. 2017. Physiological acclimation dampens initial effects of elevated temperature and atmospheric CO2 concentration in mature boreal Norway spruce. Plant, Cell and Environment. doi: 10.1111/pce.13097.Google Scholar
Lapola, D. M., Oyama, M. D., Nobre, C. A. and Sampaio, G. 2008. A new world natural map for global change studies. Annals of the Brazilian Academy of Sciences 80: 397408.Google Scholar
Lawrence, D. M., Thornton, P. M., Oleson, W. and Bonan, G. B. 2006. The partitioning of evapotranspiration, soil evaporation and canopy evaporation in GCM: impacts on land–atmosphere interaction. Journal of Hydrometerology 8. doi: 10.1175/JHM596.1.Google Scholar
Lee, X., Goulden, M. L., Hollinger, D. Y. et al. 2011. Observed increase in local cooling effect of deforestation at higher latitudes. Nature 479: 384387 doi: 10.1038/nature10588.CrossRefGoogle ScholarPubMed
Leggett, L. M. W. and Ball, D. A. 2015. Granger causality from changes in level of atmospheric CO2 to global surface temperature and the El Nino–Southern Oscillation, and a candidate mechanism in global photosynthesis. Atmospheric Chemistry and Physics 15: 1157111592. doi: 10.5194/acp-15-11571-2015.Google Scholar
Leggett, L. M. W. and Ball, D. A. 2018. Evidence that the global transpiration makes a substantial contribution to the global atmospheric temperature slowdown. Theoretical and Applied Climatology doi: 10.1007/s00704-18-2387-7.Google Scholar
Leuzinger, S. and Korner, C. 2007. Water savings in mature deciduous forest trees infer elevated CO2. Global Change Biology 13: 24982508. doi: 10.1111/j.1365.2007.01467.x.Google Scholar
Li, P., Feng, Z., Catalayud, V. et al. 2017. A meta-analysis on growth, physiological, and biochemical responses of woody plant species to ground-level ozone highlights the role of plant functional types. Plant, Cell and Environment 40: 23692380. doi: 10.1111/pce.13043.Google Scholar
Li, Y., Zhao, M., Motesharrei, S., Mu, Q., Kalnay, E. and Li, S. 2015. Local cooling and warming effects on forests based on satellite observations. Nature Communications. doi: 10.1038/ncomm7603.Google Scholar
Long, S. P. 2012. Virtual special issue (VSI) on mechanisms of plant response to global atmospheric change. Plant, Cell and Environment 35: 17051706. doi: 10.1111/j.1365-3040.2012.02589.x.Google Scholar
Luyssaert, S. and Cornelissen, J. H. C. 2018. Forests in flux as climate varies. Nature 556: 3536.Google Scholar
Lyons, D. A., Arvanitdis, C., Blight, A. J. et al. 2016. There are no whole truths in meta-analyses: all their truths are half-truths. Global Change Biology 22: 968971. doi: 10.1111/gcb.12989.Google Scholar
Manning, W. J. 2000. Use of protective chemicals to assess the effects of ambient ozone on plants. In: Environmental Pollution and Plant Responses, eds. Agrawal, S. B. and Agrawal, M.. Boca Raton: Lewis, pp. 247258.Google Scholar
Manning, W. J. 2005. Establishing a cause and effect relationship for ambient ozone exposure and tree growth in the forest: progress and an experimental approach. Environmental Pollution 137: 443454. doi: 10/1016/j.envpol.2005.01.031.Google Scholar
Manning, W. J. and Krupa, S. V. 1992. Experimental methodology for studying the effects of ozone on crops and trees. In: Surface Level Ozone Exposures: Effects on Vegetation, Lefohn, ed. A. S. Chelsea. MI: Lewis Publishers, Chapter 4, pp. 93148.Google Scholar
Manning, W. J., Flagler, R. B. and Frenkel, M. 2003. Assessing plant response to ambient ozone: growth of ozone-sensitive loblolly pine seedlings treated with ethylenediurea or sodium erythorbate. Environmental Pollution 126: 7381. doi: 10.1016/S0269-7491 (03)00141-6.CrossRefGoogle ScholarPubMed
Manning, W. J., Paoletti, E., Sandermann, H. Jr. and Ernst, D. 2011. Ethylenediurea (EDU): a research tool for assessment and verification of the effects of ground level ozone on plants under natural conditions. Environmental Pollution 150: 32833293. doi: 10.1016/j.envpol.2011.07.005.Google Scholar
Matyssek, R., Bahnweg, G., Ceulemans, R. et al. 2007. Synopsis of the CASIROZ case study: carbon sink strength of Fagus sylvatica L. in a changing environment – experimental risk assessment of mitigation by chronic ozone impact. Plant Biology 9: 163180. doi: 10.1055/s-2007-964883.Google Scholar
Medlyn, B. E., Duursma, R. A. and Zeppel, J. B. 2011. Forest productivity under climate change: a check-list for evaluating model studies. WIRES Climate Change 2: 332355.Google Scholar
Nabuurs, G.-J., Lindner, M., Verkerk, P. J. et al. 2013. First signs of carbons sink saturation in European forest biomass. Nature Climate Change. doi: 10.1038/nclimate1853.Google Scholar
NASA. 2000. Measuring vegetation (NDVI and EVI). https://earthobservatory.nasa.gov/Features/MeasuringVegetation (accessed 14/10/2017).Google Scholar
NCAR. 2017. Climate Data Guide FLUXNET. https://climatedataguide.ucar.edu/climate-data/fluxnet (accessed 02/09/2017).Google Scholar
Norby, R. J., Edwards, N. T., Riggs, J. S. et al. 1997. Temperature-controlled open-top chambers for global change research. Global Change Biology 3: 259267.Google Scholar
Overdieck, D. 2016. Research methods. In: CO2, Temperature and Trees – Experimental Approaches. Ecological Research Monographs. Dordrecht: Springer, Chapter 2.Google Scholar
Pallardy, S. G. 2008. Photosynthesis. In: Physiology of Woody Plants. New York: Academic Press, pp. 107167.Google Scholar
Peck, S. L. 2000. A tutorial for understanding modeling papers for the nonmodeler. American Entomologist 46: 4049.Google Scholar
Peng, S.-S., Pioa, S., Zeng, A. et al. 2014. Afforestation in China cools local land surface temperature. Proceedings of the National Academy of Sciences 111: 29152919. doi: 10.1073/pnas.1315126111.Google Scholar
Penuelas, J., Rutishauser, T. and Filella, I. 2009. Phenology feedbacks on climate change. Science 324: 887888. doi: 10. 1126/science.1173004.Google Scholar
Pepin, S. and Korner, C. 2002. Web-FACE: a new canopy free-air CO2 enrichment system for tall trees in mature forests. Oecologia 133: 19. doi: 10.1007/s00442-002-1088-3.Google Scholar
Peters, R. L., Groenendijk, P., Vlam, M. and Zuidema, P. A. 2015. Detecting long-term growth trends using tree rings: a critical evaluation of methods. Global Change Biology 21: 20402054. doi: 10.111/gcb.2826.Google Scholar
Pye, J. M. 1988. Impact of ozone on the growth and yield of trees: a review. Journal of Environmental Quality 17: 347360.Google Scholar
Reich, P. B. and Lassoie, J. P. 1985. Influence of low concentrations of ozone on growth biomass partitioning and leaf senescence in young hybrid poplar plants. Environmental Pollution 39: 3951. doi: 10.1111/j.1461-0248.2008.01172.x.Google Scholar
Reich, P. B., Sendall, K. M., Rice, K. et al. 2015. Geographic range predicts photosynthetic and growth responses to warming in co-occurring tree species. Nature Climate Change. doi: 10.1038/nclimate2497.Google Scholar
Reich, P. B., Sendall, K. M., Stefanski, A. et al. 2016. Boreal and temperate trees show strong acclimation of respiration to warming. Nature. doi: 10.1038/nature17142.Google Scholar
Rey, A. and Jarvis, , 1998. Long-term photosynthetic acclimation to increased atmospheric CO2 concentration in young birch (Betula pendula) trees. Tree Physiology 18: 441450.Google Scholar
Roberntz, P. and Stockfors, J. 1998. Effects of elevated CO2 concentration and nutrition on net photosynthesis, stomatal conductance and needle respiration of field-grown Norway spruce trees. Tree Physiology 18: 233241.Google Scholar
Rogers, B. M., Solvik, K., Hogg, E. H. et al. 2017. Detecting early warning signals of tree mortality in boreal North America using multiscale satellite data. Global Change Biology. doi: 10.1111/gcb.14107.Google Scholar
Samuelson, L. and Kelly, J. M. 2001. Scaling ozone effects from seedlings to forest trees. New Phytologist 149: 2141.Google Scholar
Schimel, D., Stephens, B. B. and Fisher, J. 2015. Effect of increasing CO2 on the terrestrial carbon cycle. Proceedings of the National Academy of Sciences 112: 436441. doi: 10.1073/pnas.1407302112.Google Scholar
Schultz, N. M., Lawrence, P. J. and Lee, X. 2017. Global satellite data highlights the diurnal asymmetry of the surface temperature response to deforestation. Journal of Geophysical Research: Biogeosciences 122: 903917. doi: 10.1002/2016/jG003653.Google Scholar
Shen, M., Piao, S., Jeong, S.-J. et al. 2015. Evaporative cooling over the Tibetan Plateau induced by vegetation growth. Proceedings of the National Academy of Sciences 112: 92999304. doi: 10.1073/pnas.15044418112.Google Scholar
Sigurdsson, B. D., Medhurst, J. L., Wallin, G., Eggertsson, O. and Linder, S. 2013. Growth of mature boreal Norway spruce was not affected by elevated [CO2] and/or air temperature unless nutrient availability was improved. Tree Physiology 33: 11921205. doi: 10.1093/treephys/tpt043.Google Scholar
Smith, W. K., Reed, C., Cleveland, C. C. et al. 2015. Large divergence of satellite and Earth system model estimates of global terrestrial CO2 fertilization. Nature Climate Change 6: 306310. doi: 10.1038/nclimate2879.Google Scholar
Spracklen, D. V., Bonn, B. and Carslaw, R. S. 2008. Boreal forests, aerosols and the impact on clouds and climate. Philosophical Transactions of the Royal Society A 366 : 46134626.Google Scholar
Sun, Y., Frankenberg, C., Wood, J. D. et al. 2017. OCO-2 advances photosynthesis observation via solar-induced chlorophyll fluorescence. Science 358 eaam5747. doi: 10.1126/science.aam5747.Google Scholar
Swann, A., Fung, I. Y., Levis, S., Bonan, G. B. and Doney, S. C. 2010. Changes in Arctic vegetation amplify high-latitude warming through the greenhouse effect. Proceedings of the National Academy of Sciences 107: 12951300. doi: 20.1073/pnas.0913846107.Google Scholar
Talhelm, A. F., Pregitzer, K. S., Kubiske, M. E. et al. 2014. Elevated carbon dioxide and ozone alter productivity and ecosystem carbon content in northern temperate forests. Global Change Biology 20: 24922504. doi: 10.1111/gcb.12564.Google Scholar
Tang, B., Zhao, X. and Zhao, W. 2018. Local effects of forests on temperatures across Europe. Remote Sensing 10. doi: 10.13390/rs10040529.Google Scholar
Tang, Y., Chen, A. and Zhao, S. 2016. Carbon storage and sequestration of urban street trees in Beijing, China. Frontiers in Ecology and Evolution 4: article 53. doi: 10.3389/fevo.2016.00053.Google Scholar
Tjoelker, M. G., Oleksen, J. and Reich, P. B. 1998. Seedlings of five boreal tree species differ in acclimation of net photosynthesis to elevated CO2 and temperature. Tree Physiology 18: 715726.Google Scholar
Ueyama, M., Iwata, H. and Harazono, . 2013. Autumn warming reduces the CO2 sink of a black spruce forest in interior Alaska based on a nine-year eddy covariance measurement. Global Change Biology 20: (4) https://doi.org/10.1111/gcb.12434.Google Scholar
van der Sleen, P., Groenendijk, P., Viam, M. et al. 2014. No growth stimulation of tropical trees by 150 years of CO2 fertilization but water-use efficiency increased. Nature Geoscience 8: 2428. doi: 10.1038/ngeo2313.Google Scholar
Walberg, E. 2018. Forests in Paris: how trees factor into global climate decisions. Manomet Newsletter. www.manomet.org/publication/forests-in-paris-how-trees-factor-into-global-climate-decisions/ (accessed 08/04/2018).Google Scholar
Warren, J. M., Jensen, A. M., Medlyn, B. E., Norby, R. J. and Tissue, D. T. 2014. Carbon dioxide stimulation of photosynthesis in Liquidambar styraciflua is not sustained during a 12-year field experiment. AoB Plants 7: plu074. doi: 10.1093/aobpla/plu074.Google Scholar
Way, D.A., Oren, R. and Kroner, Y. 2015. The space-time continuum: the effects of elevated CO2 and temperature and the importance of scaling. Plant, Cell and Environment 38: 6. doi: 10.1111/pce.12527.Google Scholar
Wieloch, T., Ehlers, I., Jun, Y. et al. 2018. Intramolecular 13C analysis of tree rings provides multiple plant ecophysiology signals covering decades. Scientific Reports 8: 5048. doi: 10.1038/s41598-018-23422-2.Google 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: 396424. doi: 10.1111/j.1365-24586.2008.01774.x.Google Scholar
Zhang, T., Niinemets, U., Sheffield, J. and Lichstein, J. 2018. Shifts in tree functional composition amplify the response of forest biomass to climate. Nature 556: 99102. doi: 10.1038/nature 26152.Google Scholar
Zhang, Z. 2015. Tree-ring, a key ecological indicator of environment and climate change. Ecological Indicators 51: 107116. doi: 10.1016/j.ecolind.2014.07.042.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×