Review
Gas-phase tropospheric chemistry of biogenic volatile organic compounds: a review

https://doi.org/10.1016/S1352-2310(03)00391-1Get rights and content

Abstract

Large quantities of non-methane organic compounds are emitted into the atmosphere from biogenic sources, mainly from vegetation. These organic compounds include isoprene, C10H16 monoterpenes, C15H24 sesquiterpenes, and a number of oxygenated compounds including methanol, hexene derivatives, 2-methyl-3-buten-2-ol, and 6-methyl-5-hepten-2-one. In the troposphere these organic compounds react with hydroxyl (OH) radicals, nitrate (NO3) radicals and ozone (O3), and play an important role in the chemistry of the lower troposphere. In this article the kinetics, products and mechanisms of the tropospheric reactions of biogenic organic compounds are presented and briefly discussed.

Introduction

It is now well recognized that a wide variety of volatile non-methane organic compounds (referred to hereafter as biogenic volatile organic compounds (BVOCs)) are emitted into the atmosphere from vegetation (Guenther et al (1995), Guenther et al (2000); Fall, 1999; Fuentes et al., 2000; Geron et al., 2000). Table 1 lists a subset of the total number of BVOCs observed as plant emissions, chosen to be representative of the organic compound classes involved and including the dominant emissions. Guenther et al. (1995) have estimated that 1150×1012 g carbon (1150 Tg C) year−1 of BVOCs are emitted worldwide. As discussed in detail elsewhere (see, for example, Fall, 1999; Fuentes et al., 2000), the emission rates of BVOCs are, in general, dependent on temperature and light intensity. Although there are large uncertainties in the magnitude of emission rates of individual (and total) BVOCs, a recent estimate for North America (Guenther et al., 2000) suggests that of an estimated 84 Tg C year−1 of BVOC emissions, 30% are isoprene, 25% terpenoid compounds (see Fig. 1, Fig. 2 for structures of selected C10 and C15 BVOCs), and 40% are non-terpenoid compounds including methanol, hexene derivatives, and 2-methyl-3-buten-2-ol.

Emission inventories of BVOCs and of anthropogenic non-methane organic compounds (NMOCs) indicate that on regional and global scales the emissions of BVOCs exceed those of anthropogenic compounds, by a factor of ∼10 worldwide and a factor of ∼1.5 for the USA (Lamb et al (1987), Lamb et al (1993); World Meteorological Organization, 1995). Because of the higher atmospheric reactivity of most BVOCs compared to many anthropogenic NMOCs [calculated lifetimes of BVOCs are typically a few hours or less (see Table 1) compared to a few days for most anthropogenic NMOCs (Atkinson and Arey, 1998; Atkinson, 2000)], BVOCs are calculated to play a dominant role in the chemistry of the lower troposphere and atmospheric boundary layer (Fuentes et al., 2000).

In the presence of NO emitted from combustion sources (mainly anthropogenic and exemplified by vehicle exhaust in an urban area such as Los Angeles, CA) and, to a lesser extent, from soils, atmospheric reactions of BVOCs lead to the formation of O3 and other manifestations of photochemical air pollution (National Research Council, 1991). The only significant formation route of O3 in the troposphere is the photolysis of NO2:NO2+hν→NO+O(3P)O(3P)+O2+MO3+M(M=air)Organic peroxy (RO2) radicals and HO2 radicals formed during the photooxidations of biogenic and anthropogenic NMOCs react with NO to form NO2:RO2+NORO+NO2HO2+NOOH+NO2whose photolysis then leads to net O3 formation through reactions (1) and (2).

Even for such a highly urbanized area as Los Angeles, CA, the estimated summer-day BVOC emissions are 125–140 ton/day (Benjamin et al., 1997), which is ∼15% of the estimated 2000 summertime anthropogenic NMOC emissions of 937 ton/day and ∼33% of the 413 ton/day of NMOC emissions calculated to be the upper value allowable if the 120 ppbv Federal National Ambient Air Quality Standard (NAAQS) for O3 (1-h average) is to be achieved in the Los Angeles air basin (South Coast Air Quality Management District, 2002). BVOCs therefore make the attainment of the NAAQS for O3 more difficult utilizing only control of anthropogenic NMOC emissions, and it may be necessary in many parts of the USA, and indeed in many urban areas worldwide, to also control anthropogenic NOx emissions from, for example, fossil-fueled power plants and gasoline- and diesel-fueled vehicles. Region-specific strategies utilizing VOC and/or NOx controls in the USA must now, of course, be reexamined in the light of the new Federal 80 ppbv, 8-h average ozone standard. For example, while the 120 ppbv, 1-h standard was violated in the Los Angeles air basin on 40 and 36 days in 2000 and 2001, respectively, the corresponding violations of the 80 ppbv, 8-h standard were 111 and 100 days (South Coast Air Quality Management District, 2002).

Section snippets

Tropospheric loss processes for BVOCS

As with other volatile organic compounds (Atkinson, 2000), the potential removal and transformation processes for BVOCs are wet and dry deposition, photolysis, reaction with the hydroxyl (OH) radical, reaction with the nitrate (NO3) radical, and reaction with ozone (O3). Reaction with chlorine (Cl) atoms may also be important in, for example, coastal areas (Oum et al., 1998). For most BVOCs, dry and wet deposition is probably of minor importance, though these physical removal processes could be

Lifetimes of biogenic organic compounds in the troposphere

Rate constants for the gas-phase reactions of many of the BVOCs emitted from vegetation with OH radicals, NO3 radicals and O3 have been measured. These rate constants can be combined with assumed ambient tropospheric concentrations of OH radicals, NO3 radicals and O3 to calculate the BVOC lifetime (time for decay of the BVOC to 1/e of its initial concentration) with respect to each of these loss processes (as shown in Table 1 for selected BVOCs). The data in Table 1 indicate that many of the

Reaction mechanisms and products

The initial reactions of OH radicals, NO3 radicals and O3 with NMOCs (including BVOCs) have been elucidated over the past two decades (see, for example, Atkinson (1997a), Atkinson (2000); Calvert et al., 2000) and the reactions of BVOCs have been previously reviewed by Atkinson and Arey (1998) and Calogirou et al. (1999a). For the BVOCs listed in Table 1 there are two general reaction mechanisms: (1) addition to CC bonds by OH radicals, NO3 radicals and O3 and (2) H-atom abstraction from C–H

Conclusions

Emissions of non-methane organic compounds from vegetation are to a large extent composed of compounds containing reactive CC bonds. All BVOCs react with OH radicals and many also react rapidly with NO3 radicals and O3 and have calculated lifetimes in the troposphere of a few hours or less. While the kinetics of the gas-phase reactions of biogenic NMOCs with OH radicals, NO3 radicals and O3 appear to be reasonably well understood, the products formed from these reactions and the detailed

Acknowledgements

The authors gratefully thank the National Science Foundation (Grant No. ATM-9909852) and the University of California Agricultural Experiment Station for support.

References (154)

  • T.S. Christoffersen et al.

    Cis-pinic acid, a possible precursor for organic aerosol formation from ozonolysis of α-pinene

    Atmospheric Environment

    (1998)
  • R. Fall

    Biogenic emissions of volatile organic compounds from higher plants

  • G. Fantechi et al.

    Mechanistic studies of the atmospheric oxidation of methyl butenol by OH radicals, ozone and NO3 radicals

    Atmospheric Environment

    (1998)
  • C. Geron et al.

    A review and synthesis of monoterpene speciation from forests in the United States

    Atmospheric Environment

    (2000)
  • T. Gierczak et al.

    Atmospheric fate of methyl vinyl ketone and methacrolein

    Journal of Photochemistry and Photobiology A: Chemistry

    (1997)
  • A. Guenther et al.

    Natural emissions of non-methane volatile organic compounds, carbon monoxide, and oxides of nitrogen from North America

    Atmospheric Environment

    (2000)
  • R. Gutbrod et al.

    Formation of OH radicals from the gas phase ozonolysis of alkenesthe unexpected role of carbonyl oxides

    Chemical Physics Letters

    (1996)
  • M. Jang et al.

    Newly characterized products and composition of secondary aerosols from the reaction of α-pinene with ozone

    Atmospheric Environment

    (1999)
  • S. Koch et al.

    Formation of new particles in the gas-phase ozonolysis of monoterpenes

    Atmospheric Environment

    (2000)
  • B. Lamb et al.

    A national inventory of biogenic hydrocarbon emissions

    Atmospheric Environment

    (1987)
  • B. Lamb et al.

    A biogenic hydrocarbon emission inventory for the U.S.A. using a simple forest canopy model

    Atmospheric Environment

    (1993)
  • A. Alvarado et al.

    Products of the gas-phase reactions of O(3P) atoms and O3 with α-pinene and 1,2-dimethyl-1-cyclohexene

    Journal of Geophysical Research

    (1998)
  • A. Alvarado et al.

    Kinetics of the gas-phase reactions of OH and NO3 radicals and O3 with the monoterpene reaction products pinonaldehyde, caronaldehyde, and sabinaketone

    Journal of Atmospheric Chemistry

    (1998)
  • J. Arey et al.

    Product study of the gas-phase reactions of monoterpenes with the OH radical in the presence of NOx

    Journal of Geophysical Research

    (1990)
  • S.M. Aschmann et al.

    Formation yields of methyl vinyl ketone and methacrolein from the gas-phase reaction of O3 with isoprene

    Environmental Science and Technology

    (1994)
  • S.M. Aschmann et al.

    Products of the gas-phase reactions of the OH radical with α- and β-pinene in the presence of NO

    Journal of Geophysical Research

    (1998)
  • Aschmann, S.M., Atkinson, R., Arey, J., 2002a. Products of reaction of OH radicals with α-pinene. Journal of...
  • R. Atkinson

    Kinetics and mechanisms of the gas-phase reactions of the NO3 radical with organic compounds

    Journal of Physical and Chemical Reference Data

    (1991)
  • R. Atkinson

    Gas-phase tropospheric chemistry of organic compounds

    Journal of Physical and Chemical Reference Data, Monograph

    (1994)
  • R. Atkinson

    Gas-phase tropospheric chemistry of volatile organic compounds1. alkanes and alkenes

    Journal of Physical and Chemical Reference Data

    (1997)
  • R. Atkinson

    Atmospheric reactions of alkoxy and β-hydroxyalkoxy radicals

    International Journal of Chemical Kinetics

    (1997)
  • R. Atkinson et al.

    Atmospheric chemistry of biogenic organic compounds

    Accounts of Chemical Research

    (1998)
  • R. Atkinson et al.

    Atmospheric chemistry of the monoterpene reaction products nopinone, camphenilone, and 4-acetyl-1-methylcyclohexene

    Journal of Atmospheric Chemistry

    (1993)
  • R. Atkinson et al.

    Kinetics of the gas-phase reactions of OH radicals with a series of α, β-unsaturated carbonyls at 299±2 K

    International Journal of Chemical Kinetics

    (1983)
  • R. Atkinson et al.

    Formation of 3-methylfuran from the gas-phase reaction of OH radicals with isoprene and the rate constant for its reaction with the OH radical

    International Journal of Chemical Kinetics

    (1989)
  • R. Atkinson et al.

    Rate constants for the gas-phase reactions of O3 with a series of monoterpenes and related compounds at 296±2 K

    International Journal of Chemical Kinetics

    (1990)
  • R. Atkinson et al.

    Formation of OH radicals in the gas-phase reactions of O3 with a series of terpenes

    Journal of Geophysical Research

    (1992)
  • R. Atkinson et al.

    Formation of O(3P) atoms and epoxides from the gas-phase reaction of O3 with isoprene

    Research in Chemical Intermediates

    (1994)
  • R. Atkinson et al.

    Rate constants for the gas-phase reactions of cis-3-hexen-1-ol, cis-3-hexenylacetate, trans-2-hexenal, and linalool with OH and NO3 radicals and O3 at 296±2 K, and OH radical formation yields from the O3 reactions

    International Journal of Chemical Kinetics

    (1995)
  • R. Atkinson et al.

    Evaluated kinetic and photochemical data for atmospheric chemistry, organic species

    Journal of Physical and Chemical Reference Data

    (1999)
  • J. Baker et al.

    Reactions of stabilized Criegee intermediates from the gas-phase reactions of O3 with selected alkenes

    International Journal of Chemical Kinetics

    (2002)
  • I. Barnes et al.

    Kinetics and products of the reactions of NO3 with monoalkenes, dialkenes, and monoterpenes

    Journal of Physical Chemistry

    (1990)
  • K.H. Becker et al.

    Production of hydrogen peroxide in forest air by reaction of ozone with terpenes

    Nature

    (1990)
  • H.-J. Benkelberg et al.

    Product distributions from the OH radical-induced oxidation of but-1-ene, methyl-substituted but-1-enes and isoprene in NOX-free air

    Physical Chemistry Chemical Physics

    (2000)
  • T. Berndt et al.

    Gas-phase reaction of NO3 radicals with isoprenea kinetic and mechanistic study

    International Journal of Chemical Kinetics

    (1997)
  • T. Berndt et al.

    Products and mechanism of the gas-phase reaction of NO3 radicals with α-pinene

    Journal of the Chemical Society—Faraday Transactions

    (1997)
  • T. Berndt et al.

    Reaction of NO3 radicals with 1,3-cyclohexadiene, α-terpinene, and α-phellandrenekinetics and products

    Berichte der Bunsen-Gesellschaft für Physikalische Chemie

    (1996)
  • A. Calogirou et al.

    Atmospheric oxidation of linalool

    Naturwissenschaften

    (1995)
  • A. Calogirou et al.

    Gas-phase reactions of nopinone, 3-isopropenyl-6-oxo-heptanal, and 5-methyl-5-vinyltetrahydrofuran-2-ol with OH, NO3, and ozone

    Environmental Science and Technology

    (1999)
  • Calvert, J.G., Atkinson, R., Kerr, J.A., Madronich, S., Moortgat, G.K., Wallington, T.J., Yarwood, G., 2000. The...
  • Cited by (1016)

    View all citing articles on Scopus
    1

    Also Department of Chemistry, University of California, Riverside, CA 92521, USA. Tel.: +1-909-787-4191; fax: +1-909-787-5004.

    View full text