Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-26T04:54:56.724Z Has data issue: false hasContentIssue false

4 - Biogeochemical and Biogeophysical Factors that Affect Trees

Published online by Cambridge University Press:  22 June 2020

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

Summary

To assess the effects of trees and forests on the temperature of the atmosphere, it is essential to review how they function in relation to biogeochemical factors, such as photosynthesis and release of biogenic hydrocarbons (BVOCs), and biophysical factors, such as albedo, deforestation, and land-use change, evapotranspiration, and ozone. Although biogeochemical and biophysical factors can be considered separately, their interactive roles determine whether trees and forests cool or warm the atmosphere.

Type
Chapter
Information
Trees and Global Warming
The Role of Forests in Cooling and Warming the Atmosphere
, pp. 81 - 159
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

Agren, G. I., Axelsson, B., Flower-Ellis, J. G. K. et al. 1980. Annual budget for a young Scots pine. Ecological Bulletin 32: 307313.Google Scholar
Ainsworth, E. A. and Rogers, A. 2007. The response of photosynthesis and stomatal conductance to rising CO2: mechanisms and environmental interactions. Plant, Cell and Environment 30: 258270. doi: 10.1111/j.1365-3040.2007.016441.x.Google Scholar
Ainsworth, E. A., Yenderek, C. R., Sitch, S., Collins, W. J. and Emberson, L. D. 2012. The effects of tropospheric ozone on net primary productivity and implications for climate change. Annual Review of Plant Biology 63: 637661. doi: 10.1146/annurev-arplant=042110-103829.Google Scholar
Aschan, G. and Pfanz, H. 2003. Non-foliar photosynthesis – a strategy for additional carbon acquisition. Flora 198: 8197.Google Scholar
Aschan, G., Wittmann, C. and Pfanz, H. 2001. Age-dependent bark photosynthesis of aspen twigs. Trees 15: 431437. doi: 10.1007/s00460100120.Google Scholar
Avila-Lovera, E. and Ezcurra, E. 2016. Stem-succulent trees from the Old and New World tropics. In Tropical Tree Physiology, eds Santiago, Louis S. and Goldstein, Guillermo. Berlin: Springer, pp. 4565.Google Scholar
Avila-Lovera, E., Herrera, A. and Tezara, W. 2014. Contribution of stem CO2 fixation to whole-plant carbon balance in nonsucculent species. Photosynthetica 52: 315. 10.1007/s11099-014-0004-2.Google Scholar
Aydin, Y. M., Yaman, B., Koca, H. et al. 2014. Comparison of biogenic volatile organic compound emissions from broad leaved and coniferous trees in Turkey. WIT Transactions on Ecology and the Environment 181: 647658.Google Scholar
Bartelink, H. H. 1997. Allometric relationships for biomass and leaf area of beech (Fagus sylvatica L.). Annals of Forest Science 54: 3950.CrossRefGoogle Scholar
Bartholomay, G. A., Eckert, R. T. and Smith, K. T. 1997. Reductions in tree-ring widths of white pine following ozone exposure at Acadia National Park, Maine, U.S.A. Canadian Journal of Forest Research 36: 361368.Google Scholar
Bassow, St. and Bazzaz, F. A. 1998. How environmental conditions affect canopy leaf-level photosynthesis in four deciduous tree species. Ecology 79: 26602675.Google Scholar
Bela, M. M., Longo, K. M., Freitas, S. R. et al. 2015. Ozone-production and transport over the Amazon Basin during the dry-to-wet and wet-to-dry transition seasons. Atmospheric Chemistry and Physics 15: 757782. doi: 10.5194/acp-15-757/2015.Google Scholar
Benjamin, M. T., Sudol, M., Bloch, L. and Winer, A. M. 1996. Low-emitting urban forests: a taxonomic methodology. Atmospheric Environment 30: 14371452. doi: 10.1016/1352-2310(95)00439-4.Google Scholar
Betts, R. A. 2000. Offset of the potential carbon sink from boreal forestation by decreases in surface albedo. Nature 408: 187190. doi: 10.1038/35041445.Google Scholar
Bloemen, J., Teskey, R. O., McGuire, M. A., Aubrey, D. P. and Steppe, K. 2016. Root xylem CO2 flux: an important but unaccounted-for component of root respiration. Trees 30: 343352. doi: 10.1007/s00468-015-1185-4N.CrossRefGoogle Scholar
Bonan, G. B. 1992. Effects of boreal forest vegetation on global climate. Nature 359: 716718.Google Scholar
Bonan, G. B. 2008. Forests and climate change: forcings, feedbacks, and the climate benefits of forests. Science 320: 14441449. doi: 10.1126/science.1155121.Google Scholar
Bond, B. 2000. Age-related changes in photosynthesis of woody plants. Trends in Plant Science 5: 349353. doi: 10.1016/S1360–1385(00)01691-5.Google Scholar
Breda, N. J. J. 2003. Ground-based measurements of leaf area index: a review of methods, instruments, and current controversies. Journal of Experimental Botany 54: 24032407. doi: 10.1093/jxb/erg263.Google Scholar
Bruggemann, N., Gessler, A., Keel, S. G., et al. 2011. Carbon allocation and carbon isotope fluxes in the plant–soil–atmosphere continuum: a review. Biogeosciences 8: 34573489. doi: 10.5194/bg-8-3457/2011.Google Scholar
Chameides, W. L., Lindsay, W. R., Richardson, J. and Kiang, C. S. 1988. The role of biogenic hydrocarbons in urban photochemical smog: Atlanta as a case study. Science 241: 14731475.CrossRefGoogle ScholarPubMed
Chen, G., Yang, Y. and Robinson, D. 2014. Allometric constraints on, and trade-offs in, ground carbon allocation and their control of soil respiration across global forest ecosystems. Global Change Biology 20: 16741682 doi: 10.1111/gcb.1294.CrossRefGoogle Scholar
Cherubini, F., Vezhapparambu, S., Bogren, W., Astrup, R. and Stromman, A. H. 2017. Spatial, seasonal, and topographical patterns of surface albedo in Norwegian forests and cropland. International Journal of Remote Sensing 38: 45654586.Google Scholar
Cieslik, S., Omasa, K. and Paoletti, E. 2009. Why and how terrestrial plants exchange gases with air. Plant Biology 11: 2434.Google Scholar
Climate Data Information. Albedo. 2010–2015. www.climatedata.info/forcing/albedo (accessed 11/02/2018).Google Scholar
Cooley, D. R. and Manning, W. J. 1987. The impact of ozone on assimilate partitioning in plants: a review. Environmental Pollution 47: 95113. doi: qo.1016/0269-7491 (87)90040-6.Google Scholar
Cornic, G. 2000. Drought stress inhibits photosynthesis by decreasing stomatal aperture – not by affecting ATP synthesis. Trends in Plant Science 5: 187188.Google Scholar
Cotrufo, M. F., Ineson, P. and Scott, A. 1998. Elevated CO2 reduces the nitrogen concentration of plant tissues. Global Change Biology 4: 4354.Google Scholar
Craine, J. M., Elmore, A. J., Wang, L. et al. 2018. Isotopic evidence for oligotrophication of terrestrial ecosystems. Nature Ecology and Evolution: 2: 17351744. doi: 10.1038/s41559–018-0694-0.Google Scholar
Dandois, P. and Ellis, E. C. 2010. Remote sensing of vegetation structure using computer vision. Remote Sensing 2: 11571176. doi: 10.3390/rs2041157.Google Scholar
de Dios, V. R., Mereed, T. E., Ferrio, J. P., Tissue, D. T. and Voltas, J. 2016. Intraspecific variation in juvenile tree growth under elevated CO2 alone and with O3: a meta-analysis. Tree Physiology 36: 682693. doi: 10.1093/treephys/twp026.Google Scholar
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 turn-over processes at two temperate forest free-air CO2 enrichment sites. New Phytologist 203: 883899. doi: 10.1111/nph.12847.Google Scholar
de Wit, H. A., Bryn, H. A., Hofgaard, A. et al. 2014. Climate warming feedback from mountain birch forest expansion: reduce albedo dominates carbon uptake. Global Change Biology 20: 23442355. doi: 10.1111/gcb.12483.Google Scholar
Dicke, M. and Loreto, F. 2010. Induced plant volatiles: from genes to climate change. Trends in Plant Science 15: 115117. doi: 10.016/j.tplants.2010.01.007.Google Scholar
Dickson, R. E. 1989. Carbon and nitrogen allocation in trees. Annals of Forest Science 46: supplement 631s–647s.Google Scholar
Dietz, M. C., Sala, A., Carbone, M. S. et al. 2014. Nonstructural carbon in woody plants. Annual Review of Plant Biology 65: 667687. doi: 10.1146/annurev-arplant-050213-040054.Google Scholar
Dillway, D. N. and Kruger, E. L. 2010. Thermal acclimation of photosynthesis: a comparison of boreal and temperate tree species along a latitudinal transect. Plant, Cell and Environment 33: 888899. doi: 10.1111/j.1365-3040.2010,02114.x.Google Scholar
Dittmann, S., Thiessen, E. and Hartung, E. 2017. Applicability of different non-invasive methods for tree mass estimation: a review. Forest Ecology and Management 398: 208215. doi: 10.1016/jforeco.2017.05.013.Google Scholar
Donovan, R. G., Stewart, H. E., Owen, S. M., Mackenzie, A. R. and Hewitt, C. N. 2005. Development and application of an urban tree air quality score for photochemical pollution episodes, using the Birmingham, United Kingdom area as a case study. Environmental Science and Technology 39: 67306738. doi: 10.1021/es050581y.CrossRefGoogle ScholarPubMed
Drake, J. E., Raetz, L. M., Davis, S. C. and Delucia, E. H. 2010. Hydraulic limitation not declining nitrogen availability causes the age-related photosynthetic decline in loblolly pine (Pinus taeda L.). Plant, Cell and Environment 33: 17561766.Google Scholar
Duursma, R. A., Kolari, P., Permaki, M. et al. 2009. Contributions of climate, leaf area index and leaf physiology to variations in gross primary production of six coniferous forests across Europe: a model-based analysis. Tree Physiology 29: 621629. doi: 10.1093/treephys/tpp010.CrossRefGoogle ScholarPubMed
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.gloenvcha.2017.01.02.Google Scholar
Ellsworth, D. S. and Reich, P. B. 1993. Canopy structure and vertical patterns of photosynthesis and related leaf traits in a deciduous forest. Oecologia 96: 169178.Google Scholar
Ellsworth, D. S., Anderson, I. C., Crous, K. Y. et al. 2017. Elevated CO2 does not increase eucalypt forest productivity on a low phosphorus soil. Nature Climate Change 6: 279281.Google Scholar
Epron, D., Bahn, M., Derien, D. et al. 2012. Pulse-labelling trees to study carbon allocation dynamics: a review of methods, current knowledge and future prospects. Tree Physiology 32: 776798. doi: 10.1093/treephys/tps057.Google Scholar
Ericsson, T., Rytter, L. and Vapaavuor, E. 1996. Physiology of carbon allocation in trees. Biomass and Bioenergy 11: 115127.CrossRefGoogle Scholar
Fan, H., McGuire, M. A. and Teskey, R. O. 2017. Effects of stem size on stem respiration and its flux components in yellow-poplar (Liriodendron tulipifera L.) leaves. Tree Physiology. doi: 10.1093/treephys/tpx084.Google Scholar
Farmer, G. T. and Cook, J. 2013. Earth’s albedo, radiative forcing and climate change. In: Climate Change Science: A Modern Synthesis, Vol. 1, Physical Climate. Dordrecht: Springer, pp. 217229.Google Scholar
Felzer, B. S., Cronin, T., Reilly, J. M., Melillo, J. M. and Wang, X. 2007. Impacts of ozone on trees and crops. Comptes Rendus Geoscience 339: 784798. doi: 10.1016/j.crte.2007.08.008.Google Scholar
Fernandez-Martinez, S., Vicca, I. A., Janssens, J. et al. 2014. Nutrient availability as the key regulator of global carbon balance. Nature Climate Change 4: 471478. doi: 10.1038/nclimate2177.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. doi: 10.1038/nclimate2614.Google Scholar
Gedney, N., Cox, P. M., Betts, R. A. et al. 2006. Detection of direct carbon dioxide effect in continental river runoff records. Nature Geoscience 439: 835838. doi: 10.1038/nature04504.Google Scholar
Ghannoum, O., Phillips, N. G., Sears, M. E. et al. 2010. Photosynthetic responses of two eucalypts to industrial-age changes in atmospheric CO2 and temperature. Plant, Cell and Environment 33: 16711681. doi: 10.1111.j.1365-3040.2010.02172.x.CrossRefGoogle ScholarPubMed
Greenberg, J. P., Guenther, A. B., Petron, G. et al. 2004. Biogenic VOC emissions from forested Amazonian landscapes. Global Change Biology 10: 651662. doi: 10.1111/j.1529-8817.2003.00758.x.Google Scholar
Griffin, K. L. and Prager, C. M. 2017. Where does all the carbon go? Thermal acclimation of respiration and increased photosynthesis in trees at the temperate-boreal ecotone. Tree Physiology 37: 281284. doi: 10.1093/treephys/tpw133.Google Scholar
Guan, K., Pan, M., Li, H. et al. 2015. Photosythetic seasonality of global tropical forests constrained by hydroclimate. Nature Geoscience 8. doi: 10.038/ngeo2382.Google Scholar
Hantson, S., Knorr, W., Schurgers, G., Pugh, T. A. M. and Arneth, A. 2017. Global isoprene and monoterpene emissions under changing climate, vegetation, CO2 and land use. Atmospheric Environment 155: 3545. doi: 10.1016/j.atmosenv.2017.02.010.Google Scholar
Helmisaari, H.-S., Makkonen, K., Kellomaki, S., Valtonen, E. and Malkonen, E. 2002. Below-ground and above-ground biomass, production and nitrogen use in Scots pine stands in eastern Finland. Forest Ecology and Management 165: 317326.Google Scholar
Herrick, J. D. and Thomas, R. B. 2003. Leaf senescence and late-season net photosynthesis of sun and shade leaves of overstory sweetgum (Liquidambar styraciflua) grown in elevated and ambient carbon dioxide concentrations. Tree Physiology 23: 109118.Google Scholar
Hirosaka, K., Nabeshima, E. and Hiua, T. 2007. Seasonal changes in the temperature of photosynthesis in canopy leaves of Quercus crispula in a cool-temperature forest. Tree Physiology 27: 10351041.Google Scholar
Hogberg, P., Nordgren, A. and Agren, G. I. 2002. Carbon allocation between tree root growth and root respiration. Oecologia 132: 579581 doi: 10.1007/s00442-002-0983-8.Google Scholar
Hogberg, P., Hogberg, M. N., Gottlicher, S. G. et al. 2008. High temporal resolution tracing of photosynthate carbon from the tree canopy to forest soil microorganisms. New Phytologist 177:220228. doi: 10.1111/j.1469-8137.2007. 02238.x.Google Scholar
Holtta, T., Mencuccini, M. and Nikinmaa, E. 2014. Ecophysiological aspects of phloem transport in trees. In: Trees in a Changing Environment. Plant Ecophysiology No. 9, eds. Tausz, M. and Grulke, N.. Dordrecht: Springer Science+Business Media, Chapter 2.Google Scholar
Holtum, J. A. M. and Winter, K. 2010. Elevated CO2 and forest vegetation: more a water issue than a carbon issue? Functional Plant Biology 37: 694702. doi: 10.1071/FP10001.Google Scholar
Hovi, A., Liang, J., Korhonen, L., Kobayashi, H. and Rautiainen, M. 2016. Quantifying the missing link between forest productivity and albedo in the boreal zone. Biogeosciences 13: 60156030. doi: 10.5194/bg-13-2016.Google Scholar
Jach, M. E. and Ceulemans, R. 2000. Effects of season, needle age and elevated atmospheric CO2 on photosynthesis in Scots pine (Pinus sylvestris). Tree Physiology 20: 145157. doi: 10.1093/treephys/20.3.145.CrossRefGoogle ScholarPubMed
Janssens, I. A. and Luyssaert, S. 2009. Nitrogen’s carbon bonus. Nature Geoscience 2: 318319.Google Scholar
Jasechko, S., Sharp, Z. D, Gibson, J. et al. 2013. Terrestrial water fluxes dominated by transpiration. Nature 496: 347351. doi: 10.1038/nature11983.CrossRefGoogle ScholarPubMed
Jensen, A. M., Warren, J. M., Hanson, P. J., Childs, J. and Wullschleger, S. D. 2015. Needle age and season influence photosynthetic temperature response and total annual carbon uptake in mature Picea mariana trees. Annals of Botany 116: 821832. doi: 10.1093/aob/mcv115.Google Scholar
Jiang, L., Tian, D., Ma, S. et al. 2018. The response of tree growth to nitrogen and phosphorus additions in a tropical montane rainforest. Science of the Total Environment 618: 10641070. doi: 10.1016/j.sciotenv.2017.09.099.Google Scholar
Jokinen, T., Berndt, T., Makkonen, R. et al. 2015. Production of extremely low volatile organic compounds from biogenic emissions: measured yields and atmospheric limitations. Proceedings of the National Academy of Sciences 112: 71237128. doi: 10.1073/pnas.1423977112.Google Scholar
Jolivet, Y., Bagard, M., Cabane, M. et al. 2016. Deciphering the ozone-induced changes in cellular processes: a prerequisite for risk assessment at the tree and forest levels. Annals of Forest Science 73: 923943. doi: 10.1007/s13595-016-580-3.Google Scholar
Jurik, T. W. 1986. Seasonal patterns of leaf photosynthetic capacity in successional northern hardwood tree species. American Journal of Botany 73: 131138.Google Scholar
Kankare, V., Holopainen, M., Vastaranta, M. et al. 2013. Individual tree biomass estimation using terrestrial laser scanning. ISPRS Journal of Photogrammetry and Remote Sensing 75: 6475. doi: 10.1016/j.isprsjprs2012.10.003.CrossRefGoogle Scholar
Karnosky, D. F., Zak, D. R., Pregitzer, K. S. et al. 2003. Tropospheric O3 moderates responses of temperate hardwood forests to elevated CO2. A synthesis of molecular to ecosystem results from the Aspen Face project. Functional Ecology 17: 289305.Google Scholar
Keeling, C. D., Chin, J. F. S. and Whorf, T. P. 1996. Increased activity of northern vegetation inferred from atmosphere CO2 measurements. Nature 382: 146.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: 324327. doi: 10.1038./nature12291.Google Scholar
King, J. S., Kubiske, M. E., Pregitzer, K. S. et al. 2005. Tropospheric ozone compromises net primary production in young stands of trembling aspen, paper birch and sugar maple in response to elevated atmospheric ozone. New Phytologist 168: 623636. doi: 10.1111/j.1469-8137.2005.0157.x.Google Scholar
Kirschbaum, M. U. F., Whitehead, D., Dean, S. M. et al. 2011. Implications of albedo changes following afforestation on the benefits of forests as carbon sinks. Biogeosciences 8: 36873696. doi: 10.5194/bg-8-3687/2011.Google Scholar
Klein, T. and Hoch, G. 2015. Tree carbon allocation dynamics determined using a carbon mass balance approach. New Phytologist 205: 147159. doi: 10.1111/nph.12993.Google Scholar
Klingberg, J., Engardt, M., Uddling, J., Karlsson, P. E. and Pleijel, H. 2011. Ozone risk for vegetation in the future climate of Europe based on stomatal ozone uptake calculations. Tellus 63A: 174187. doi: 10.1111/j.1600-0870.2010.00465.x.Google Scholar
Konôpka, B., Pajtik, J., Moravcik, M. and Lukac, M. 2010. Biomass partitioning and growth efficiency in four naturally regenerated forest tree species. Basic and Applied Ecology 11: 234243. doi: 10.1016/j.baae.2010.02.004.Google Scholar
Korner, C. 2003. Ecological impacts of atmospheric CO2 enrichment on terrestrial ecosystems. Philosphical Transactions of the Royal Society A 361 (1810): 20232041. https://doi.org/10.1098/rsts.2003.1241.CrossRefGoogle ScholarPubMed
Kozlowski, T. T. 1992. Carbohydrate sources and sinks in woody plants. Botanical Review 58: 107222.Google Scholar
Kramer, P. and Kozlowski, T. T. 1979. Physiology of Woody Plants. New York: Academic Press.Google Scholar
Krupa, S. V. and Manning, W. J. 1988. Atmospheric ozone: formation and effects on vegetation. Environmental Pollution 50: 101137.Google Scholar
Kumala, M., Suni, T., Lehtinen, K. E. J. et al. 2004. A new feedback mechanism linking forests, aerosols, and climate. Atmospheric Chemistry and Physics 4: 557562.Google Scholar
Kuusinen, N., Stenberg, P., Korhonen, L., Rautiainen, M. and Tomppo, E. 2015. Structural factors driving boreal forest albedo in Finland. Remote Sensing of the Environment 175: 4351.CrossRefGoogle 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
Larcher, W. 2003. Physiological Plant Ecology. Heidelberg: Springer.Google Scholar
Lemoine, R., La Camera, S., Atanassova, R. et al. 2013. Source to sink transport of sugar and regulation by environmental factors. Frontiers in Plant Science 4: article 272. doi: 10.3389/fpls.203.00272.Google Scholar
Lerdau, M., Litvak, M., Palmer, P. and Monson, R. 1997. Controls over monoterpene emissions from boreal forests. Tree Physiology 17: 563569.Google Scholar
Lin, Y.-S., Medlyn, B. E. and Ellsworth, D. S. 2012. Temperature responses of leaf net photosynthesis: the role of component processes. Tree Physiology 32: 219231. doi: 10.1093/treephys/tpr141.Google Scholar
Liu, Z., Wu, C. and Wang, S. 2017. Predicting forest evapotranspiration by coupling carbon and water cycling based on a critical stomatal conductance model. Journal of Selected Topics in Applied Earth Observations 10: 44694477.Google Scholar
Long, S. P. 2012. Virtual Special Issue on mechanisms of plant response to global atmospheric change. Plant, Cell and Environment 35: 17051706. doi: 1111/j.1365-3040.2012.0258.x.Google Scholar
Loreto, F., Ciccioli, P., Brancaleoni, E., Cecinato, A. and Fratoni, M. 1996. Different sources of reduced carbon contribute to form three classes of terpenoids emitted by Quercus ilex leaves. Proceedings of the National Academy of Sciences 93: 99669969.Google Scholar
Luyssaert, S., Schulze, E.-D., Borner, A. et al. 2008. Old-growth forests as global carbon sinks. Nature 455. doi: 10.10.1038/nature07276.Google Scholar
Manning, W. J. and Feder, W. A. 1980. Biomonitoring Air Pollutants with Plants. London: Applied Science Publishers.Google Scholar
Mao, J., Ribes, A., Yan, B. et al. 2016. Human-induced greening of the northern extratropical land surface. Nature Climate Change 6. doi: 10.1038/nclimate3056.Google Scholar
Markings, S. 2017. The effects of temperature on the rate of photosynthesis. https://sciencing.com/effect-temperature-rate-photosynthesis-19595.html (accessed 13/09/2017).Google Scholar
Matyssek, R. 1986. Carbon, water and nitrogen relations in evergreen and deciduous conifers. Tree Physiology 2: 177187.Google Scholar
McDonald, E. P., Erickson, J. E. and Kruger, E. L. 2002. Research Note: Can decreased transpiration limit plant nitrogen acquisition in elevated CO2? Functional Plant Biology 29:11151120.Google Scholar
McLaughlin, S. B. and Downing, D. J. 1995. Interactive effects of ambient ozone and climate measured on growth of mature forest trees. Nature 374: 252255.Google Scholar
McLaughlin, S. B., McConathy, R. K., Duvick, D. and Mann, L. K. 1982. Effects of chronic air pollution stress on photosynthesis, carbon allocation, and growth of white pine trees. Forest Science 28: 6070.Google Scholar
Mercado, L. M., Belliouin, N., Sitch, S. et al. 2009. Impact of changes in diffuse radiation on the global land carbon sink. Nature 458: 10141017. doi: 1038/nature07949.Google Scholar
Miller, P. R., Arbaugh, M. J. and Temple, P. J. 1997. Ozone and its known and potential effects on forests in the Western United States. In: Forest Decline and Ozone, eds Sanderman, H., Wellburn, A. R. and Heath, R. L.. Ecological Studies Series 127. Springer, pp. 3967.Google Scholar
Morrison, J. I. L. 1998. Stomatal response to increased CO2 concentration. Journal of Experimental Botany 49: 443452.Google Scholar
Moura, B. B., Alves, E. S., Marabesi, M. A. et al. 2018. Ozone affects leaf physiology and causes injury to foliage of native tree species from the tropical Atlantic Forest of Southern Brazil. Science of the Total Environment 610–611: 912925. doi: 10.1016/j.scitoenv.2017.8.130.Google Scholar
Muukkonen, P. and Heiskanen, J. 2005. Estimating biomass for boreal forests using ASTER satellite data combined with standwise forest inventory data. Remote Sensing of the Environment 99: 434447. doi: 10.116/j.rse2005.9.CrossRefGoogle Scholar
Mykleby, P. M., Snyder, P. K. and Twine, T. E. 2017. Quantifying the trade-off between carbon sequestration and albedo in midlatitude and high-altitude North American forests. Geophysical Research Letters 44: 24932501. doi: 10.1002/2016GL071459.CrossRefGoogle Scholar
Nadelhoffer, K. J., Emmett, B. A., Gundersen, P. et al. 1999. Nitrogen deposition make a minor contribution to carbon sequestration in temperate forests. Nature 398: 145198.Google Scholar
NASA. 2011. Measuring Earth’s albedo: Image of the Day 2011. https://earthobservatory.nasa.gov/IOTD/view.php?id=8449 (accessed 11/02/2018).Google Scholar
Naudts, K., Chen, Y., McGrath, M. J. et al. 2016. Europe’s forest management did not mitigate climate warming. Science 351: 597600.Google Scholar
Niinemets, U., Flexas, J. and Penuelas, J. 2010. Evergreens favored by higher responsiveness to increased CO2. Trends in Ecology and Evolution 26: 136142. doi: 10.1016/j.tree.2010.12.012.Google Scholar
Nowak, D. J. 1996. Estimating leaf area and leaf biomass of open-grown deciduous trees. Forest Science 42: 540547.Google Scholar
Oren, R., Ellsworth, D. S., Johnsen, K. H. et al. 2001. Soil fertility limits carbon sequestration by forest ecosystems in a CO2-enriched atmosphere. Nature 411: 469472. doi: 10.1038/35078064.Google Scholar
Pallardy, S. G. 2008. Photosynthesis. In: Physiology of Woody Plants. New York: Academic Press, pp.107167.Google Scholar
Penuelas, J. and Staudt, M. 2009. BVOCs and global change. Trends in Plant Science 15: 133134.Google Scholar
Perry, D. A., Oren, O. and Hart, S. C. 2008. Forest Ecosystems. Baltimore: Johns Hopkins University Press.Google Scholar
Popescue, S. C. 2007. Estimating biomass of individual pine trees using airborne lidar. Biomass and Bioenergy 31: 646655. doi: 10.1016/j.biombioe.2007.06.022.Google Scholar
Rap, A., Scott, C. E., Reddington, L. et al. 2018. Enhanced global primary production by biogenic aerosol via diffuse radiation fertilization. Nature Geoscience 11: 640644. www.nature.com/articles/s41561-018-0208-3.Google Scholar
Reay, D. S., Detenter, F., Smith, P., Grace, J. and Feeley, R. A. 2008. Global nitrogen deposition and carbon sinks. Nature Geoscience 1: 430437. doi: 10.1038/ngeo230.Google Scholar
Reich, P. B. 1987. Quantifying plant response to ozone: a unifying theory. Tree Physiology 3: 6391.CrossRefGoogle ScholarPubMed
Reich, P. B., Sendell, K. M., Stefansky, A. et al. 2016. Boreal and temperate trees show strong acclimation of respiration to warming. Nature 531: 633636. doi: 10.1038/nature17142.Google Scholar
Rennenberg, H. and Schmidt, S. 2010. Perennial lifestyle – an adaptation to nutrient limitation? Tree Physiology 30: 10471049. doi: 10.1093/treephys/tpq076.Google Scholar
Richards, B. L., Taylor, O. C. and Edmunds, G. F. Jr. 1968. Ozone needle mottle in Southern California. Journal of the Air Pollution Control Association 18: 7377. doi: 10.1080/00022470.1968.10469097.Google Scholar
Roberts, J. M. 2009. The role of forests in the hydrological cycle. In: Forests and Forest Plants Vol. III. UNESCO-EOLSS.Google Scholar
Rogers, A. and Humphries, S. W. 2000. A mechanistic evaluation of photosynthetic acclimation at elevated CO2. Global Change Biology 6: 10051011. doi: 10.1046/j.1365-2486.2000.00375.x.Google Scholar
Royal Society of Chemistry: Advancing the Chemical Sciences. Rate of photosynthesis: limiting factors. www.rsc.org/search-results/?q= rate of photosynthesis limiting factors (accessed 11/06/2017).Google Scholar
Ryan, M. G. and Asao, S. 2014. Phloem transport in trees. Tree Physiology 34: 14 doi: 10.1093/treephys/tpt123.Google Scholar
Ryan, M. G., Gower, S. T., Hubbard, R. M. et al. 1995. Woody tissue maintenance respiration of four conifers in contrasting climates. Oecologia 101: 133140.CrossRefGoogle ScholarPubMed
Ryan, M. G., Lavigne, M. B. and Gower, S. T. 1997. Annual carbon cost of autrophic respiration in boreal forest ecosystems in relation to species and climate. Journal of Geophysical Research 102: No. D24 28.871–28.883.Google Scholar
Sala, A., Woodruff, D. R. and Meinzer, F. C. 2012. Carbon dynamics in trees: feast or famine? Tree Physiology 32: 764775. doi: 10.1093/treephys/tpr143.Google Scholar
Sanderman, H., Wellburn, A. R. and Heath, R. L. (eds) 1997. Forest Decline and Ozone. Dordrecht: Springer Ecological Studies Series 127.CrossRefGoogle Scholar
Sanderson, M. G., Collins, W. J., Hemming, D. L. and Betts, R. A. 2007. Stomatal conductance changes due to increasing carbon dioxide level: projected impact on surface ozone levels. Tellus 59B: 404411. doi: 10.1111/j.1600-0889.2007.00277.x.Google Scholar
Sara, V., Luyssaert, S., Penuelas, J. et al. 2012. Fertile forests produce biomass more efficiently. Ecology Letters 15: 520526. doi: 10.1111/j.1461-1=0248.2012.01775.Google Scholar
Sass-Klassen, U. 2015. Tracking tree carbon gain. Nature Plants 1: 12. doi: 10.1038/nplants.2015.175.CrossRefGoogle Scholar
Saveyn, A., Steppe, K. and Lemeur, R. 2008. Report on non-temperature related variations in CO2 efflux rates from young tree stems in the dormant season. Trees 22: 165174.Google Scholar
Saveyn, A., Steppe, K., Ubierna, N. and Dawson, T. 2010. Woody tissue photosynthesis and its contribution to trunk growth and bud development in young plants. Plant, Cell and Environment 33: 19491958. doi: 10.1111/j.1365-3040.2010.021997.x.Google Scholar
Schlesinger, W. H. and Jasechko, S. 2014. Transpiration in the global water cycle. Agricultural and Forest Meteorology 189–190: 115117.Google Scholar
Schulte-Uebbing, L. and de Vries, W. 2017. Global-scale impacts of nitrogen depositon on tree carbon sequestration in tropical, temporal and boreal forests: a meta-analysis. Global Change Biology. doi: 10.1111/gcb.13862.Google Scholar
Scott, C. E., Monks, S. A., Spracklen, D. V. et al. 2018. Impact on short-lived climate forcers increases projected warming due to deforestation. Nature Communications 9: 17. doi: 10.1038/s41467-017-02412-4.Google Scholar
Sharkey, T. D. and Singass, E. L. 1995. Why plants emit isoprene. Nature 374: 769.Google Scholar
Sharkey, T. D. and Monson, R. K. 2014. The future of isoprene emissions from leaves, canopies and landscapes. Plant, Cell and Environment. doi: 10.1111/pce.12289.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: 9299-9304 July 28 www.pnas.org/cgi/doi/10.1073/pnas.15044418112.Google Scholar
Sillett, S. C., Van Pelt, R., Koch, G. W. et al. 2010. Increasing wood production through old age in tall trees. Forest Ecology and Management 259: 976994. doi: 10.1016/jforeco.2009.12.003.Google Scholar
Simpson, J. R. and McPherson, E. G. 2011. The tree BVOC index. Environmental Pollution 159: 20882093. doi: 10.1016/j.envpol.2011.02.034..Google Scholar
Sitch, S., Cox, P. M., Collins, W. J. and Huntingford, C. 2007. Indirect radiative forcing of climate change through ozone effects on the land-carbon sink. Nature 448: 791794. August doi: 10.1038/nature06059.Google Scholar
Slot, M. and Winter, K. 2017. Photosynthetic acclimation to warming in tropical tree seedlings. Journal of Experimental Botany 68: 22752284. doi: 10.1093/jxb/erx071.Google Scholar
Stephenson, N. L., Das, A. J., Condit, R. et al. 2014. Rate of tree carbon accumulation increases continuously with tree size. Nature 507: 9093. doi: 10.1038/nature12914.Google Scholar
Sun, Z., Niinemets, U., Huve, K., Rasulov, B. and Noe, S. M. 2013. Elevated atmospheric CO2 concentration leads to increased whole-plant isoprene emission in hybrid aspen (Populus tremula x P. tremuloides. New Phytologist 198: 788800. doi: 10.1111/nph.12200.Google Scholar
Sun, Y., Gu, L., Dickinson, R. E. et al. 2014. Impact of mesophyll diffusion on estimated global land CO2 fertilization. Proceedings of the National Academy of Sciences 111: 1577415779. doi: 10.1073/pnas.1418075111.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: 10.1073/pnas.0913846107.Google Scholar
Tang, J., Luyssaert, S., Richardson, A. D., Kutsch, W. and Janssens, I. A. 2014. Steeper declines in forest photosynthesis than respiration explain age-driven decreases in forest growth. Proceedings of the National Academy of Sciences 111: 88568860.CrossRefGoogle ScholarPubMed
Taub, D. R. and Wang, X. 2008. Why are nitrogen concentrations in plant tissues lower under elevated CO2? A critical examination of the hypotheses. Journal of Integrative Biology 50: 13651374. doi: 10.1111/j.1744-7909.2008.00754.x.Google Scholar
Taylor, G., Tallis, M. J., Giardina, C. P. et al. 2007. Future atmospheric CO2 leads to delayed autumnal senescence. Global Change Biology 14: 112. doi: 10.1111/j.1365-2486.2007.0147.x.Google Scholar
Teskey, R. O., Saveyn, A., Steppe, K. and McGuire, M. A. 2008. Origin, fate and significance of CO2 in tree stems. New Phytologist 177: 1732. doi: 10.1111/j.1469-8137.2007.02286.x.Google Scholar
Thomas, R. Q., Canham, C. D., Weathers, K. C. and Goodale, C. L. 2009. Increased tree carbon storage in response to nitrogen deposition in the US. Nature Geoscience 3: 1317. doi: 10.1038/ngeo721.Google Scholar
Thomas, S. C. 2010. Photosynthetic capacity peaks at intermediate size in temperate deciduous trees. Tree Physiology 30: 555573. doi: 1093/treephys/tpq005.Google Scholar
Topping, D., Connolly, P. and McFiggans, G. 2013. Cloud droplet number enhanced by co-condensation of organic vapors. Nature Geoscience 6: 443446. doi: 10.1038/ngeo1809.Google Scholar
Tunved, P., Hansson, H.-C., Kerminen, V.-M. et al. 2006. High natural aerosol loading over boreal forests. Science 312: 262263. doi: 10.1126/science.1123052.Google Scholar
Turnbull, M. H., Whitehead, D., Tissue, D. T. et al. 2001. Responses of leaf respiration to temperature and leaf characteristics in three deciduous tree species vary with site water availability. Tree Physiology 21: 571578.Google Scholar
Uddling, J., Hogg, A. J., Teclaw, R. M., Carroll, M. A. and Ellsworth, D. S. 2010. Stomatal uptake of O3 in aspen and aspen-birch forests under free-air CO2 and O3 enrichment. Environmental Pollution 158: 20232031. doi: 10.1016/j.envpol.2009.12.001.Google Scholar
Unger, N. 2012. New directions: enduring ozone. Atmospheric Environment 55: 456-458. doi: 10.1016/j.atmosenv.2012.03.036.Google Scholar
Unger, N. 2014a. Human land-use-driven reduction of forest volatiles cools global climate. Nature Climate Change 4: 907910. doi: 10.1038/nclimate2347.Google Scholar
Unger, N. 2014b. On the role of plant volatiles in anthropogenic global climate change. Geophysical Research Letters. 41: 85638569. doi: 10.1002/2014GL061616.Google Scholar
US Geological Survey. 2016. Evapotranspiration – the water cycle. shttps://water.usgs.gov/edu/watercycleevapotranspiration.html (accessed 15/02/2018).Google Scholar
Vandegehuchte, M. W., Bloemen, J., Vergeynst, L. and Stepe, K. 2015. Woody tissue photosynthesis in trees: salve on the wounds of drought. New Phytologist 208: 9981002.Google 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: 2428. doi: 10.1038/ngeo2313.Google Scholar
Wan, W., Manning, W. J., Wang, X. et al. 2014. Ozone and ozone injury on plants in and around Beijing, China. Environmental Pollution 191: 215222. doi: 10.1016/j.envpol.2014.02.035.Google Scholar
Wang, B., Shugart, H. H., Shuman, J. K. and Lerdue, M. T. 2016. Forests and ozone: productivity, carbon storage, and feedbacks. Scientific Reports 6: 17. doi: 10.1038/srep22133.Google Scholar
Wang, M., Shi, S., Lin, F. et al. 2012. Effects of soil water and nitrogen on growth and photosynthetic response of Manchurian ash (Fraxinus mandschurica) seedlings in Northeast China. PLoS One 7: 113.Google Scholar
Wang, M., Schurgers, G., Arneth, A., Eckberg, A. and Hoist, M. 2017. Seasonal variation in biogenic volatile organic emissions from Norway spruce in a Swedish boreal forest. Boreal Environment Research 22: 353-367.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
Wenzel, S., Cox, P. M., Eyring, V. and Friedlingstein, P. 2016. Projected land photosynthesis constrained by changes in the seasonal cycle of atmospheric CO2. Nature 538. doi: 10.1038/nature19772.Google Scholar
Wieder, W. R., Cleveland, C. C., Kolby-Smith, W. and Todd-Brown, K. 2015. Future productivity and carbon storage limited by terrestrial nutrient availability. Nature Geoscience 8: 441444. doi: 10.1038/ngeo2413.Google Scholar
Wilson, B. F. 1984. The Growing Tree. Amherst: University of Massachusetts Press.Google Scholar
Wilson, K. B., Baldocchi, D. D. and Hanson, P. 2001. Leaf age affects seasonal pattern of photosynthetic capacity and net ecosystem exchange of carbon in a deciduous forest. Plant, Cell and Environment 24: 571583. doi: 10.1046/j.0016-8025.2001.00706.x.Google Scholar
Wittig, V. E., Ainsworth, E. A., Naidu, S. A., Karnosky, D. F. and Long, S. P. 2009. Quantifying the impact of current and future tropospheric ozone on tree biomass, growth, physiology and biochemistry. Global Change Biology 15: 396424. doi: 10.1111/j.1365-2486.01774.x.Google Scholar
Wu, J., Albert, L. P., Lopes, A. P. et al. 2016. Leaf development and demography explain photosynthesis seasonality in Amazon evergreen forests. Science 351: 972976.Google Scholar
Yoder, B. J., Ryan, M. G., Waring, R. H., Schoettle, A. W. and Kaufmann, M. R. 1994. Evidence of reduced photosynthesis rates in old trees. Forest Science 40: 513537.Google Scholar
Zeng, Z., Pioa, S., Laurent, Z. X. et al. 2017. Climate mitigation from vegetation biophysical feedbacks during the past three decades. Nature Climate Change 7: 432436. doi: 10. 1038/nclimate3299.Google Scholar
Zhao, D. F., Bucholz, A., Tillmann, R. et al. 2017. Environmental conditions regulate the impact of plants on cloud formation. Nature Communications 8: 14067 doi: 10.1038/ncomms14067.Google Scholar
Zhao, K. and Popescue, S. 2009. Lidar-based mapping of leaf area index and its use for validating GLOBOCARBON satellite LAI product in a temperate forest of the Southern USA. Remote Sensing of the Environment 113: 16281645. doi: 10.1016/j.rse.2009.03.006.CrossRefGoogle Scholar
Zhou, Y.-M., Wang, C.-G., Han, S.-J. et al. 2011. Species-specific and needle age-related responses of photosynthesis in two Pinus species to long-term exposure to elevated CO2 concentration. Trees 25: 163173. doi: 10.1007/s00468-010-0495-9.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
×