Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-26T04:31:28.870Z Has data issue: false hasContentIssue false

31 - Aerosols, Chemistry, and Climate

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

Aerosols affect the radiative balance of the atmosphere by absorbing and scattering radiation, by altering cloud albedo, and through precipitation. Their primary radiative effect is to increase planetary albedo, and aerosols have a negative radiative forcing. However, black carbon (commonly called soot) is an absorbing aerosol that heats the atmosphere, and deposition on snow and ice decreases surface albedo (a positive radiative forcing). Aerosols also have indirect effects by altering biogeochemical cycles. Mineral aerosols affect climate directly by altering the radiative balance of the atmosphere and indirectly by fertilizing ecosystems. Dust emissions may initiate a positive land–atmosphere feedback that enhances drought. Fires influence climate through emissions of long-lived greenhouse gases, organic and black carbon aerosols, and short-lived reactive gases. These latter emissions produce ozone, alter the oxidation capacity of the troposphere through the hydroxyl (OH) radical, and thereby affect the concentration of methane (CH4). The net radiative forcing of fires is the balance of these biogeochemical emissions and also biogeophysical effects from changes in surface albedo and energy fluxes. It is estimated the fires provide a negative radiative forcing. They may also decrease precipitation in a positive feedback whereby biomass burning promotes drought and greater susceptibility to fire. Plants emit numerous biogenic volatile organic compounds (BVOCs). Oxidation of these BVOCs in the presence of nitrogen oxides (NOx) forms ozone. Emissions of BVOCs also reduce the oxidation capacity of the atmosphere (OH), decreasing the atmospheric sink for CH4 and increasing its lifetime in the atmosphere. Chemical transformations also produce secondary organic aerosols. Emissions of BVOCs are thought to provide a negative radiative forcing, but this is likely to diminish in the future because of human activities. Chemistry–climate interactions from short-lived climate forcers (NOx, BVOCs, ozone, CH4, and secondary organic aerosols) are now recognized as being significant and comparable in magnitude to other climate forcings. The aerosol effects of dust, fire, and BVOCs, as well as their chemistry–climate interactions, are important feedbacks with climate change.

Type
Chapter
Information
Ecological Climatology
Concepts and Applications
, pp. 606 - 627
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

Akagi, S. K., Yokelson, R. J., Wiedinmyer, C., et al. (2011). Emission factors for open and domestic biomass burning for use in atmospheric models. Atmospheric Chemistry and Physics, 11, 4039–4072.CrossRefGoogle Scholar
Amin, H., Atkins, P. T., Russo, R. S., et al. (2012). Effect of bark beetle infestation on secondary organic aerosol precursor emissions. Environmental Science and Technology, 46, 5696–5703.CrossRefGoogle ScholarPubMed
Andreae, M. O., and Crutzen, P. J. (1997). Atmospheric aerosols: Biogeochemical sources and role in atmospheric chemistry. Science, 276, 1052–1058.CrossRefGoogle Scholar
Andreae, M. O., and Merlet, P. (2001). Emissions of trace gases and aerosols from biomass burning. Global Biogeochemical Cycles, 15, 955–966, doi:10.1029/2000GB001382.CrossRefGoogle Scholar
Andreae, M. O., and Rosenfeld, D. (2008). Aerosol–cloud–precipitation interactions, Part 1: The nature and sources of cloud-active aerosols. Earth-Science Reviews, 89, 13–41.CrossRefGoogle Scholar
Andreae, M. O., Rosenfeld, D., Artaxo, P., et al. (2004). Smoking rain clouds over the Amazon. Science, 303, 1337–1342.CrossRefGoogle ScholarPubMed
Arneth, A., Miller, P. A., Scholze, M., et al. (2007). CO2 inhibition of global terrestrial isoprene emissions: Potential implications for atmospheric chemistry. Geophysical Research Letters, 34, L18813, doi:10.1029/2007GL030615.CrossRefGoogle Scholar
Arneth, A., Monson, R. K., Schurgers, G., Niinemets, Ü., and Palmer, P. I. (2008). Why are estimates of global terrestrial isoprene emissions so similar (and why is this not so for monoterpenes)?Atmospheric Chemistry and Physics, 8, 4605–4620.CrossRefGoogle Scholar
Arneth, A., Harrison, S. P., Zaehle, S., et al. (2010). Terrestrial biogeochemical feedbacks in the climate system. Nature Geoscience, 3, 525–532.CrossRefGoogle Scholar
Ashworth, K., Folberth, G., Hewitt, C. N., and Wild, O. (2012). Impacts of near-future cultivation of biofuel feedstocks on atmospheric composition and local air quality. Atmospheric Chemistry and Physics, 12, 919–939.CrossRefGoogle Scholar
Ashworth, K., Wild, O., and Hewitt, C. N. (2013). Impacts of biofuel cultivation on mortality and crop yields. Nature Climate Change, 3, 492–496.CrossRefGoogle Scholar
Atkinson, R. (2000). Atmospheric chemistry of VOCs and NOx. Atmospheric Environment, 34, 2063–2101.CrossRefGoogle Scholar
Atkinson, R., and Arey, J. (2003). Gas-phase tropospheric chemistry of biogenic volatile organic compounds: A review. Atmospheric Environment, 37(S2), S197–S219.CrossRefGoogle Scholar
Beerling, D. J., and Osborne, C. P. (2006). The origin of the savanna biome. Global Change Biology, 12, 2023–2031.CrossRefGoogle Scholar
Beerling, D. J., Fox, A., Stevenson, D. S., and Valdes, P. J. (2011). Enhanced chemistry–climate feedbacks in past greenhouse worlds. Proceedings of the National Academy of Sciences USA, 108, 9770–9775.CrossRefGoogle ScholarPubMed
Berg, A. R., Heald, C. L., Huff Hartz, K. E., et al. (2013). The impact of bark beetle infestations on monoterpene emissions and secondary organic aerosol formation in western North America. Atmospheric Chemistry and Physics, 13, 3149–3161.CrossRefGoogle Scholar
Bond, T. C., Doherty, S. J., Fahey, D. W., et al. (2013). Bounding the role of black carbon in the climate system: A scientific assessment. Journal of Geophysical Research: Atmospheres, 118, 5380–5552, doi:10.1002/jgrd.50171.Google Scholar
Boucher, O., Randall, D., Artaxo, P., et al. (2013). Clouds and aerosols. 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. 571–657.Google Scholar
Bowman, D. M. J. S., Balch, J. K., Artaxo, P., et al. (2009). Fire in the Earth System. Science, 324, 481–484.CrossRefGoogle ScholarPubMed
Boy, M., Sogachev, A., Lauros, J., et al. (2011). SOSA – a new model to simulate the concentrations of organic vapours and sulphuric acid inside the ABL – Part 1: Model description and initial evaluation. Atmospheric Chemistry and Physics, 11, 43–51.CrossRefGoogle Scholar
Carslaw, K. S., Boucher, O., Spracklen, D. V., et al. (2010). A review of natural aerosol interactions and feedbacks within the Earth system. Atmospheric Chemistry and Physics, 10, 1701–1737.CrossRefGoogle Scholar
Chameides, W. L., Lindsay, R. W., Richardson, J., and Kiang, C. S. (1988). The role of biogenic hydrocarbons in urban photochemical smog: Atlanta as a case study. Science, 241, 1473–1475.CrossRefGoogle ScholarPubMed
Chen, J., Avise, J., Guenther, A., et al. (2009). Future land use and land cover influences on regional biogenic emissions and air quality in the United States. Atmospheric Environment, 43, 5771–5780.CrossRefGoogle Scholar
Ciccioli, P., Centritto, M., and Loreto, F. (2014). Biogenic volatile organic compound emissions from vegetation fires. Plant, Cell and Environment, 37, 1810–1825.CrossRefGoogle ScholarPubMed
Cook, B. I., Miller, R. L., and Seager, R. (2008). Dust and sea surface temperature forcing of the 1930s “Dust Bowl” drought. Geophysical Research Letters, 35, L08710, doi:10.1029/2008GL033486.CrossRefGoogle Scholar
Cook, B. I., Miller, R. L., and Seager, R. (2009). Amplification of the North American “Dust Bowl” drought through human-induced land degradation. Proceedings of the National Academy of Sciences USA, 106, 4997–5001.CrossRefGoogle ScholarPubMed
Cook, B. I., Seager, R., Miller, R. L., and Mason, J. A. (2013). Intensification of North American megadroughts through surface and dust aerosol forcing. Journal of Climate, 26, 4414–4430.CrossRefGoogle Scholar
Creamean, J. M., Suski, K. J., Rosenfeld, D., et al. (2013). Dust and biological aerosols from the Sahara and Asia influence precipitation in the western U.S.Science, 339, 1572–1578.CrossRefGoogle ScholarPubMed
Crutzen, P. J., and Andreae, M. O. (1990). Biomass burning in the tropics: Impact on atmospheric chemistry and biogeochemical cycles. Science, 250, 1669–1678.CrossRefGoogle ScholarPubMed
D'Odorico, P., Bhattachan, A., Davis, K. F., Ravi, S., and Runyan, C. W. (2013). Global desertification: Drivers and feedbacks. Advances in Water Resources, 51, 326–344.Google Scholar
Field, J. P., Belnap, J., Breshears, D. D., et al. (2010). The ecology of dust. Frontiers in Ecology and the Environment, 8, 423–430.CrossRefGoogle Scholar
Flanner, M. G., Zender, C. S., Randerson, J. T., and Rasch, P. J. (2007). Present-day climate forcing and response from black carbon in snow. Journal of Geophysical Research, 112, D11202, doi:10.1029/2006JD008003.CrossRefGoogle Scholar
Flanner, M. G., Zender, C. S., Hess, P. G., et al. (2009). Springtime warming and reduced snow cover from carbonaceous particles. Atmospheric Chemistry and Physics, 9, 2481–2497.CrossRefGoogle Scholar
Forkel, R., Klemm, O., Graus, M., et al. (2006). Trace gas exchange and gas phase chemistry in a Norway spruce forest: A study with a coupled 1-dimensional canopy atmospheric chemistry emission model. Atmospheric Environment, 40, S28–S42.CrossRefGoogle Scholar
Formenti, P., Andreae, M. O., Lang, L., et al. (2001). Saharan dust in Brazil and Suriname during the Large-Scale Biosphere-Atmosphere Experiment in Amazonia (LBA) – Cooperative LBA Regional Experiment (CLAIRE) in March 1998. Journal of Geophysical Research, 106D, 14919–14934.Google Scholar
Forster, P., Ramaswamy, V., Artaxo, P., et al. (2007). Changes in atmospheric constituents and in radiative forcing. 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, pp. 129–234.Google Scholar
Fuentes, J. D., Lerdau, M., Atkinson, R., et al. (2000). Biogenic hydrocarbons in the atmospheric boundary layer: A review. Bulletin of the American Meteorological Society, 81, 1537–1575.2.3.CO;2>CrossRefGoogle Scholar
Ganzeveld, L., and Lelieveld, J. (2004). Impact of Amazonian deforestation on atmospheric chemistry. Geophysical Research Letters, 31, L06105, doi:10.1029/2003GL019205.CrossRefGoogle Scholar
Ganzeveld, L. N., Lelieveld, J., Dentener, F. J., Krol, M. C., and Roelofs, G.-J. (2002a). Atmosphere–biosphere trace gas exchanges simulated with a single-column model. Journal of Geophysical Research, 107, doi:10.1029/2001JD000684.CrossRefGoogle Scholar
Ganzeveld, L. N., Lelieveld, J., Dentener, F. J., et al. (2002b). Global soil-biogenic NOx emissions and the role of canopy processes. Journal of Geophysical Research, 107, doi:10.1029/2001JD001289.CrossRefGoogle Scholar
Ganzeveld, L., Bouwman, L., Stehfest, E., et al. (2010). Impact of future land use and land cover changes on atmospheric chemistry–climate interactions. Journal of Geophysical Research, 115, D23301, doi:10.1029/2010JD014041.CrossRefGoogle Scholar
Gedney, N., Huntingford, C., Weedon, G. P., et al. (2014). Detection of solar dimming and brightening effects on Northern Hemisphere river flow. Nature Geoscience, 7, 796–800.CrossRefGoogle Scholar
Ginoux, P., Prospero, J. M., Gill, T. E., Hsu, N. C., and Zhao, M. (2012). Global-scale attribution of anthropogenic and natural dust sources and their emission rates based on MODIS Deep Blue aerosol products. Reviews of Geophysics, 50, RG3005, doi:10.1029/2012RG000388.CrossRefGoogle Scholar
Goldstein, A. H., Koven, C. D., Heald, C. L., and Fung, I. Y. (2009). Biogenic carbon and anthropogenic pollutants combine to form a cooling haze over the southeastern United States. Proceedings of the National Academy of Sciences USA, 106, 8835–8840.CrossRefGoogle Scholar
Grote, R., and Niinemets, Ü. (2008). Modeling volatile isoprenoid emissions – a story with split ends. Plant Biology, 10, 8–28.CrossRefGoogle ScholarPubMed
Guenther, A., Hewitt, C. N., Erickson, D., et al. (1995). A global model of natural volatile organic compound emissions. Journal of Geophysical Research, 100D, 8873–8892.Google Scholar
Guenther, A., Karl, T., Harley, P., et al. (2006). Estimates of global terrestrial isoprene emissions using MEGAN (Model of Emissions of Gases and Aerosols from Nature). Atmospheric Chemistry and Physics, 6, 3181–3210.CrossRefGoogle Scholar
Guenther, A. B., Jiang, X., Heald, C. L., et al. (2012). The Model of Emissions of Gases and Aerosols from Nature version 2.1 (MEGAN2.1): An extended and updated framework for modeling biogenic emissions. Geoscientific Model Development, 5, 1471–1492.CrossRefGoogle Scholar
Hansen, J., and Nazarenko, L. (2004). Soot climate forcing via snow and ice albedos. Proceedings of the National Academy of Sciences USA, 101, 423–428.CrossRefGoogle ScholarPubMed
Hardacre, C. J., Palmer, P. I., Baumanns, K., Rounsevell, M., and Murray-Rust, D. (2013). Probabilistic estimation of future emissions of isoprene and surface oxidant chemistry associated with land-use change in response to growing food needs. Atmospheric Chemistry and Physics, 13, 5451–5472.CrossRefGoogle Scholar
Harrison, S. P., Morfopoulos, C., Dani, K. G. S., et al. (2013). Volatile isoprenoid emissions from plastid to planet. New Phytologist, 197, 49–57.CrossRefGoogle ScholarPubMed
Heald, C. L., Henze, D. K., Horowitz, L. W., et al. (2008). Predicted change in global secondary organic aerosol concentrations in response to future climate, emissions, and land use change. Journal of Geophysical Research, 113, D05211, doi:10.1029/2007JD009092.CrossRefGoogle Scholar
Heald, C. L., Wilkinson, M. J., Monson, R. K., et al. (2009). Response of isoprene emission to ambient CO2 changes and implications for global budgets. Global Change Biology, 15, 1127–1140.CrossRefGoogle Scholar
Herwitz, S. R., Muhs, D. R., Prospero, J. M., Mahan, S., and Vaughn, B. (1996). Origin of Bermuda's clay-rich Quaternary paleosols and their paleoclimatic significance. Journal of Geophysical Research, 101D, 23389–23400.Google Scholar
Hoelzemann, J. J., Schultz, M. G., Brasseur, G. P., Granier, C., and Simon, M. (2004). Global Wildland Fire Emission Model (GWEM): Evaluating the use of global area burnt satellite data. Journal of Geophysical Research, 109, D14S04, doi:10.1029/2003JD003666.CrossRefGoogle Scholar
Huneeus, N., Schulz, M., Balkanski, Y., et al. (2011). Global dust model intercomparison in AeroCom phase I. Atmospheric Chemistry and Physics, 11, 7781–7816.CrossRefGoogle Scholar
Jaffe, D. A., and Wigder, N. L. (2012). Ozone production from wildfires: A critical review. Atmospheric Environment, 51, 1–10.CrossRefGoogle Scholar
Jickells, T. D., An, Z. S., Andersen, K. K., et al. (2005). Global iron connections between desert dust, ocean biogeochemistry, and climate. Science, 308, 67–71.CrossRefGoogle ScholarPubMed
Jimenez, J. L., Canagaratna, M. R., Donahue, N. M., et al. (2009). Evolution of organic aerosols in the atmosphere. Science, 326, 1525–1529.CrossRefGoogle ScholarPubMed
Keenan, T., Niinemets, Ü., Sabate, S., Gracia, C., and Peñuelas, J. (2009). Process based inventory of isoprenoid emissions from European forests: Model comparisons, current knowledge and uncertainties. Atmospheric Chemistry and Physics, 9, 4053–4076.CrossRefGoogle Scholar
Koren, I., Kaufman, Y. J., Remer, L. A., and Martins, J. V. (2004). Measurement of the effect of Amazon smoke on inhibition of cloud formation. Science, 303, 1342–1345.CrossRefGoogle ScholarPubMed
Kulmala, M., Suni, T., Lehtinen, K. E. J., et al. (2004). A new feedback mechanism linking forests, aerosols, and climate. Atmospheric Chemistry and Physics, 4, 557–562.CrossRefGoogle Scholar
Kulmala, M., Nieminen, T., Nikandrova, A., et al. (2014). CO2-induced terrestrial climate feedback mechanism: From carbon sink to aerosol source and back. Boreal Environment Research, 19 (suppl. B), 122–131.Google Scholar
Kurtén, T., Kulmala, M., Dal Maso, M., et al. (2003). Estimation of different forest-related contributions to the radiative balance using observations in southern Finland. Boreal Environment Research, 8, 275–285.Google Scholar
Laothawornkitkul, J., Taylor, J. E., Paul, N. D., and Hewitt, C. N. (2009). Biogenic volatile organic compounds in the Earth system. New Phytologist, 183, 27–51.CrossRefGoogle ScholarPubMed
Lathière, J., Hauglustaine, D. A., De Noblet-Ducoudré, N., Krinner, G., and Folberth, G. A. (2005). Past and future changes in biogenic volatile organic compound emissions simulated with a global dynamic vegetation model. Geophysical Research Letters, 32, L20818, doi:10.1029/2005GL024164.CrossRefGoogle Scholar
Lathière, J., Hauglustaine, D. A., Friend, A. D., et al. (2006). Impact of climate variability and land use changes on global biogenic volatile organic compound emissions. Atmospheric Chemistry and Physics, 6, 2129–2146.CrossRefGoogle Scholar
Lathière, J., Hewitt, C. N., and Beerling, D. J. (2010). Sensitivity of isoprene emissions from the terrestrial biosphere to 20th century changes in atmospheric CO2 concentration, climate, and land use. Global Biogeochemical Cycles, 24, GB1004, doi:10.1029/2009GB003548.CrossRefGoogle Scholar
Lee, Y. H., Lamarque, J.-F., Flanner, M. G., et al. (2013). Evaluation of preindustrial to present-day black carbon and its albedo forcing from Atmospheric Chemistry and Climate Model Intercomparison Project (ACCMIP). Atmospheric Chemistry and Physics, 13, 2607–2634.CrossRefGoogle Scholar
Lihavainen, H., Kerminen, V.-M., Tunved, P., et al. (2009). Observational signature of the direct radiative effect by natural boreal forest aerosols and its relation to the corresponding first indirect effect. Journal of Geophysical Research, 114, D20206, doi:10.1029/2009JD012078.CrossRefGoogle Scholar
Loreto, F., and Schnitzler, J.-P. (2010). Abiotic stresses and induced BVOCs. Trends in Plant Science, 15, 154–166.CrossRefGoogle ScholarPubMed
Mahowald, N. (2011). Aerosol indirect effect on biogeochemical cycles and climate. Science, 334, 794–796.CrossRefGoogle ScholarPubMed
Mahowald, N. M., and Luo, C. (2003). A less dusty future?Geophysical Research Letters, 30, 1903, doi:10.1029/2003GL017880.CrossRefGoogle Scholar
Mahowald, N., Kohfeld, K., Hansson, M., et al. (1999). Dust sources and deposition during the last glacial maximum and current climate: A comparison of model results with paleodata from ice cores and marine sediments. Journal of Geophysical Research, 104D, 15895–15916.Google Scholar
Mahowald, N. M., Baker, A. R., Bergametti, G., et al. (2005). Atmospheric global dust cycle and iron inputs to the ocean. Global Biogeochemical Cycles, 19, GB4025, doi:10.1029/2004GB002402.CrossRefGoogle Scholar
Mahowald, N. M., Muhs, D. R., Levis, S., et al. (2006). Change in atmospheric mineral aerosols in response to climate: Last glacial period, preindustrial, modern, and doubled carbon dioxide climates. Journal of Geophysical Research, 111, D10202, doi:10.1029/2005JD006653.Google Scholar
Mahowald, N., Jickells, T. D., Baker, A. R., et al. (2008). Global distribution of atmospheric phosphorus sources, concentrations and deposition rates, and anthropogenic impacts. Global Biogeochemical Cycles, 22, GB4026, doi:10.1029/2008GB003240.CrossRefGoogle Scholar
Mahowald, N. M., Engelstaedter, S., Luo, C., et al. (2009). Atmospheric iron deposition: Global distribution, variability, and human perturbations. Annual Review of Marine Science, 1, 245–278CrossRefGoogle ScholarPubMed
Mahowald, N. M., Kloster, S., Engelstaedter, S., et al. (2010). Observed 20th century desert dust variability: Impact on climate and biogeochemistry. Atmospheric Chemistry and Physics, 10, 10875–10893.CrossRefGoogle Scholar
Mahowald, N., Ward, D. S., Kloster, S., et al. (2011). Aerosol impacts on climate and biogeochemistry. Annual Review of Environment and Resources, 36, 45–74.CrossRefGoogle Scholar
Mao, J., Horowitz, L. W., Naik, V., et al. (2013). Sensitivity of tropospheric oxidants to biomass burning emissions: Implications for radiative forcing. Geophysical Research Letters, 40, 1241–1246, doi:10.1002/grl.50210.CrossRefGoogle Scholar
Marcella, M. P., and Eltahir, E. A. B. (2014). The role of mineral aerosols in shaping the regional climate of West Africa. Journal of Geophysical Research: Atmospheres, 119, 5806–5822, doi:10.1002/2012JD019394.Google Scholar
Martin, S. T., Andreae, M. O., Artaxo, P., et al. (2010). Sources and properties of Amazonian aerosol particles. Reviews of Geophysics, 48, RG2002, doi:10.1029/2008RG000280.CrossRefGoogle Scholar
McGee, D., Broecker, W. S., and Winckler, G. (2010). Gustiness: The driver of glacial dustiness?Quaternary Science Reviews, 29, 2340–2350.CrossRefGoogle Scholar
Mulitza, S., Heslop, D., Pittauerova, D., et al. (2010). Increase in African dust flux at the onset of commercial agriculture in the Sahel region. Nature, 466, 226–228.CrossRefGoogle ScholarPubMed
Myhre, G., Shindell, D., Bréon, F.-M., et al. (2013). Anthropogenic and natural radiative forcing. 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. 659–740.Google Scholar
Neff, J. C., Ballantyne, A. P., Farmer, G. L., et al. (2008). Increasing eolian dust deposition in the western United States linked to human activity. Nature Geoscience, 1, 189–195.CrossRefGoogle Scholar
Nicholson, S. E. (2000). Land surface processes and Sahel climate. Reviews of Geophysics, 38, 117–140.CrossRefGoogle Scholar
Okin, G. S., Mahowald, N., Chadwick, O. A., and Artaxo, P. (2004). Impact of desert dust on the biogeochemistry of phosphorus in terrestrial ecosystems. Global Biogeochemical Cycles, 18, GB2005, doi:10.1029/2003GB002145.CrossRefGoogle Scholar
Paasonen, P., Asmi, A., Petäjä, T., et al. (2013). Warming-induced increase in aerosol number concentration likely to moderate climate change. Nature Geoscience, 6, 438–442.CrossRefGoogle Scholar
Pacifico, F., Harrison, S. P., Jones, C. D., and Sitch, S. (2009). Isoprene emissions and climate. Atmospheric Environment, 43, 6121–6135.CrossRefGoogle Scholar
Painter, T. H., Barrett, A. P., Landry, C. C., et al. (2007). Impact of disturbed desert soils on duration of mountain snow cover. Geophysical Research Letters, 34, L12502, doi:10.1029/2007GL030284.CrossRefGoogle Scholar
Painter, T. H., Deems, J. S., Belnap, J., et al. (2010). Response of Colorado River runoff to dust radiative forcing in snow. Proceedings of the National Academy of Sciences USA, 107, 17125–17130.CrossRefGoogle ScholarPubMed
Painter, T. H., Skiles, S. M., Deems, J. S., Bryant, A. C., and Landry, C. C. (2012). Dust radiative forcing in snow of the Upper Colorado River Basin, 1: A 6 year record of energy balance, radiation, and dust concentrations. Water Resources Research, 48, W07521, doi:10.1029/2012WR011985.CrossRefGoogle Scholar
Park, J.-H., Goldstein, A. H., Timkovsky, J., et al. (2013). Active atmosphere–ecosystem exchange of the vast majority of detected volatile organic compounds. Science, 341, 643–647.CrossRefGoogle ScholarPubMed
Peñuelas, J., and Staudt, M. (2010). BVOCs and global change. Trends in Plant Science, 15, 133–144.CrossRefGoogle ScholarPubMed
Pöschl, U., Martin, S. T., Sinha, B., et al. (2010). Rainforest aerosols as biogenic nuclei of clouds and precipitation in the Amazon. Science, 329, 1513–1516.CrossRefGoogle ScholarPubMed
Prospero, J. M., and Lamb, P. J. (2003). African droughts and dust transport to the Caribbean: Climate change implications. Science, 302, 1024–1027.CrossRefGoogle ScholarPubMed
Prospero, J. M., and Nees, R. T. (1977). Dust concentration in the atmosphere of the equatorial North Atlantic: Possible relationship to the Sahelian drought. Science, 196, 1196–1198.CrossRefGoogle ScholarPubMed
Prospero, J. M., and Nees, R. T. (1986). Impact of the North African drought and El Niño on mineral dust in the Barbados trade winds. Nature, 320, 735–738.CrossRefGoogle Scholar
Prospero, J. M., Glaccum, R. A., and Nees, R. T. (1981). Atmospheric transport of soil dust from Africa to South America. Nature, 289, 570–572.CrossRefGoogle Scholar
Prospero, J. M., Nees, R. T., and Uematsu, M. (1987). Deposition rate of particulate and dissolved aluminum derived from Saharan dust in precipitation at Miami, Florida. Journal of Geophysical Research, 92D, 14723–14731.Google Scholar
Prospero, J. M., Barrett, K., Church, T., et al. (1996). Atmospheric deposition of nutrients to the North Atlantic Basin. Biogeochemistry, 35, 27–73.CrossRefGoogle Scholar
Purves, D. W., Caspersen, J. P., Moorcroft, P. R., Hurtt, G. C., and Pacala, S. W. (2004). Human-induced changes in US biogenic volatile organic compound emissions: Evidence from long-term forest inventory data. Global Change Biology, 10, 1737–1755.CrossRefGoogle Scholar
Ramanathan, V., and Carmichael, G. (2008). Global and regional climate changes due to black carbon. Nature Geoscience, 1, 221–227.CrossRefGoogle Scholar
Ramanathan, V., Crutzen, P. J., Kiehl, J. T., and Rosenfeld, D. (2001). Aerosols, climate, and the hydrological cycle. Science, 294, 2119–2124.CrossRefGoogle ScholarPubMed
Randerson, J. T., Liu, H., Flanner, M. G., et al. (2006). The impact of boreal forest fire on climate warming. Science, 314, 1130–1132.CrossRefGoogle ScholarPubMed
Ravi, S., D'Odorico, P., Breshears, D. D., et al. (2011). Aeolian processes and the biosphere. Reviews of Geophysics, 49, RG3001, doi:10.1029/2010RG000328.CrossRefGoogle Scholar
Ridley, D. A., Heald, C. L., and Prospero, J. M. (2014). What controls the recent changes in African mineral dust aerosol across the Atlantic?Atmospheric Chemistry and Physics, 14, 5735–5747.CrossRefGoogle Scholar
Rosenfeld, D., Rudich, Y., and Lahav, R. (2001). Desert dust suppressing precipitation: A possible desertification feedback loop. Proceedings of the National Academy of Sciences USA, 98, 5975–5980.CrossRefGoogle ScholarPubMed
Rosenfeld, D., Lohmann, U., Raga, G. B., et al. (2008). Flood or drought: How do aerosols affect precipitation?Science, 321, 1309–1313.CrossRefGoogle ScholarPubMed
Rosenfeld, D., Andreae, M. O., Asmi, A., et al. (2014). Global observations of aerosol-cloud-precipitation climate interactions. Reviews of Geophysics, 52, doi:10.1002/2013RG000441.CrossRefGoogle Scholar
Saylor, R. D. (2013). The Atmospheric Chemistry and Canopy Exchange Simulation System (ACCESS): Model description and application to a temperate deciduous forest canopy. Atmospheric Chemistry and Physics, 13, 693–715.CrossRefGoogle Scholar
Schultz, M. G., Heil, A., Hoelzemann, J. J., et al. (2008). Global wildland fire emissions from 1960 to 2000. Global Biogeochemical Cycles, 22, GB2002, doi:10.1029/2007GB003031.CrossRefGoogle Scholar
Scott, C. E., Rap, A., Spracklen, D. V., et al. (2014). The direct and indirect radiative effects of biogenic secondary organic aerosol. Atmospheric Chemistry and Physics, 14, 447–470.CrossRefGoogle Scholar
Sharkey, T. D., and Monson, R. K. (2014). The future of isoprene emission from leaves, canopies and landscapes. Plant, Cell and Environment, 37, 1727–1740.CrossRefGoogle ScholarPubMed
Sharkey, T. D., Wiberley, A. E., and Donohue, A. R. (2008). Isoprene emission from plants: Why and how. Annals of Botany, 101, 5–18.Google Scholar
Spracklen, D. V., Bonn, B., and Carslaw, K. S. (2008). Boreal forests, aerosols and the impacts on clouds and climate. Philosophical Transactions of the Royal Society A, 366, 4613–4626.CrossRefGoogle ScholarPubMed
Squire, O. J., Archibald, A. T., Abraham, N. L., et al. (2014). Influence of future climate and cropland expansion on isoprene emissions and tropospheric ozone. Atmospheric Chemistry and Physics, 14, 1011–1024.CrossRefGoogle Scholar
Stanelle, T., Bey, I., Raddatz, T., Reick, C., and Tegen, I. (2014). Anthropogenically induced changes in twentieth century mineral dust burden and the associated impact on radiative forcing. Journal of Geophysical Research: Atmospheres, 119, 13526–13546, doi:10.1002/2014JD022062.Google Scholar
Swap, R., Garstang, M., Greco, S., Talbot, R., and Kållberg, P. (1992). Saharan dust in the Amazon Basin. Tellus B, 44, 133–149.CrossRefGoogle Scholar
Tai, A. P. K., Mickley, L. J., Heald, C. L., and Wu, S. (2013). Effect of CO2 inhibition on biogenic isoprene emission: Implications for air quality under 2000 to 2050 changes in climate, vegetation, and land use. Geophysical Research Letters, 40, 3479–3483, doi:10.1002/grl.50650.CrossRefGoogle Scholar
Tegen, I., and Fung, I. (1995). Contribution to the atmospheric mineral aerosol load from land surface modification. Journal of Geophysical Research, 100D, 18707–18726.Google Scholar
Tegen, I., Werner, M., Harrison, S. P., and Kohfeld, K. E. (2004). Relative importance of climate and land use in determining present and future global soil dust emission. Geophysical Research Letters, 31, L05105, doi:10.1029/2003GL019216.Google Scholar
Tosca, M. G., Randerson, J. T., Zender, C. S., Flanner, M. G., and Rasch, P. J. (2010). Do biomass burning aerosols intensify drought in equatorial Asia during El Niño?Atmospheric Chemistry and Physics, 10, 3515–3528.CrossRefGoogle Scholar
Tosca, M. G., Randerson, J. T., and Zender, C. S. (2013). Global impact of smoke aerosols from landscape fires on climate and the Hadley circulation. Atmospheric Chemistry and Physics, 13, 5227–5241.CrossRefGoogle Scholar
Tunved, P., Hansson, H.-C., Kerminen, V.-M., et al. (2006). High natural aerosol loading over boreal forests. Science, 312, 261–263.CrossRefGoogle ScholarPubMed
Unger, N. (2013). Isoprene emission variability through the twentieth century. Journal of Geophysical Research: Atmospheres, 118, 13606–13613, doi:10.1002/2013JD020978.Google Scholar
Unger, N. (2014). Human land-use-driven reduction of forest volatiles cools global climate. Nature Climate Change, 4, 907–910.CrossRefGoogle Scholar
Unger, N., and Yue, X. (2014). Strong chemistry–climate feedback in the Pliocene. Geophysical Research Letters, 41, 527–533, doi:10.1002/2013GL058773.CrossRefGoogle Scholar
van der Werf, G. R., Randerson, J. T., Giglio, L., et al. (2010). Global fire emissions and the contribution of deforestation, savanna, forest, agricultural, and peat fires (1997–2009). Atmospheric Chemistry and Physics, 10, 11707–11735.CrossRefGoogle Scholar
Ward, D. S., Kloster, S., Mahowald, N. M., et al. (2012). The changing radiative forcing of fires: Global model estimates for past, present and future. Atmospheric Chemistry and Physics, 12, 10857– 10886.CrossRefGoogle Scholar
Wiedinmyer, C., Akagi, S. K., Yokelson, R. J., et al. (2011). The Fire INventory from NCAR (FINN): A high resolution global model to estimate the emissions from open burning. Geoscientific Model Development, 4, 625–641.CrossRefGoogle Scholar
Wolfe, G. M., and Thornton, J. A. (2011). The Chemistry of Atmosphere-Forest Exchange (CAFE) Model – Part 1: Model description and characterization. Atmospheric Chemistry and Physics, 11, 77–101.CrossRefGoogle Scholar
Wu, S., Mickley, L. J., Kaplan, J. O., and Jacob, D. J. (2012). Impacts of changes in land use and land cover on atmospheric chemistry and air quality over the 21st century. Atmospheric Chemistry and Physics, 12, 1597–1609.CrossRefGoogle Scholar
Yoshioka, M., Mahowald, N. M., Conley, A. J., et al. (2007). Impact of desert dust radiative forcing on Sahel precipitation: Relative importance of dust compared to sea surface temperature variations, vegetation changes, and greenhouse gas warming. Journal of Climate, 20, 1445–1467.CrossRefGoogle Scholar
Zhang, Q., Jimenez, J. L., Canagaratna, M. R., et al. (2007). Ubiquity and dominance of oxygenated species in organic aerosols in anthropogenically-influenced Northern Hemisphere midlatitudes. Geophysical Research Letters, 34, L13801, doi:10.1029/2007GL029979.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.

  • Aerosols, Chemistry, and Climate
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Ecological Climatology
  • Online publication: 05 November 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781107339200.032
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.

  • Aerosols, Chemistry, and Climate
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Ecological Climatology
  • Online publication: 05 November 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781107339200.032
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.

  • Aerosols, Chemistry, and Climate
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Ecological Climatology
  • Online publication: 05 November 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781107339200.032
Available formats
×