Sei sulla pagina 1di 10

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 104, NO.

D9, PAGES 11,633–11,641, MAY 20, 1999

Atmospheric processing of organic aerosols


G. Barney Ellison
Department of Chemistry and Biochemistry, University of Colorado, Boulder

Adrian F. Tuck
NOAA Aeronomy Laboratory, Boulder, Colorado

Veronica Vaida
Department of Chemistry and Biochemistry, University of Colorado, Boulder

Abstract. We suggest a chemical model for the composition, structure, and atmospheric
processing of organic aerosols. This model is stimulated by recent field measurements
showing that organic compounds are a significant component of atmospheric aerosols. The
proposed model organic aerosol is an “inverted micelle” consisting of an aqueous core
that is encapsulated in an inert, hydrophobic organic monolayer. The organic materials
that coat the aerosol particles are surfactants of biological origin. We propose a chemical
mechanism by which the organic surface layer will be processed by reactions with
atmospheric radicals. The net result of an organic aerosol being exposed to an oxidizing
atmosphere is the transformation of an inert hydrophobic film to a reactive, optically
active hydrophilic layer. Consequently, processed organic aerosols can grow by water
accretion and form cloud condensation nuclei, influencing atmospheric radiative transfer.
Radiative transfer may be affected directly by the chromophores left on the surface of the
aerosol after chemical transformation. The chemical model yields certain predictions
which are testable by observations. Among them is a curve of the percent organic material
as a function of particle diameter which predicts that a high fraction of the mass of the
upper tropospheric aerosol will be organic. Atmospheric processing of organic aerosols
will lead to the release of small organic fragments into the troposphere which will play a
subsequent role in homogeneous chemistry. Organic aerosols are likely to act as a
transport vehicle of organics and other water insoluble compounds into the atmosphere.
We speculate that biomass burning will produce a similar coating of surfactants derived
from land sources. Finally, it is pointed out that the radical-induced transformation of the
surface layer of aerosol particles from hydrophobic to hydrophilic offers an additional
means by which the biosphere, through atmospheric chemistry, can affect the radiative
balance.

1. Introduction al., 1991; Odum et al., 1997]. Field measurements documented


that organic aerosols are widespread in continental and marine
Aerosols are ubiquitous in the troposphere and are recog- environments. Ambient cloud condensation nuclei contain 10 –
nized to play important roles in climate and elsewhere through 20% organic carbon (by mass), suggesting the need for a model
their interaction with solar and terrestrial radiation [Andreae, of the physicochemical structure of organic atmospheric aero-
1995; Andreae and Crutzen, 1997; Dickerson et al., 1997]. Re- sols [Li et al., 1998]. Questions about the role of sulfate, sea
flection of radiation by the Earth’s atmosphere is augmented salt, and organic compounds in determining the composition of
by natural and anthropogenic aerosols through their size- aerosols in the marine boundary layer have been the focus of
dependent light-scattering effects and by their ability to act as recent field studies, a notable example being the first Aerosol
condensation nuclei and increase the reflectivity of clouds Characterization Experiment (ACE 1) performed at Cape
[Twomey et al., 1984; Charlson et al., 1992; Pandis et al., 1995; Grim, Tasmania [Murphy et al., 1997; Middlebrook et al., 1998;
Rivera-Caprio et al., 1996; Schwartz, 1996; Tegen et al., 1997; Murphy et al. 1998a, b]. In the context of ACE 1, the compo-
Liao and Seinfeld, 1998; Pan et al., 1998]. Chemical composi- sition of single aerosol particles was directly measured using
tion plays an important role in determining the optical and particle analysis by laser mass spectrometry (PALMS) [Murphy
chemical properties of aerosols [Fitzgerald, 1991; O’Dowd and et al., 1997; Middlebrook et al., 1998; Murphy et al. 1998a, b].
Smith, 1993; Quinn et al., 1996; Saxena et al., 1996; Jennings et The PALMS measurements established (1) that pure sulfate
al., 1997; Novakov et al., 1997]. Organic compounds are pro- aerosols are not prevalent in the marine boundary layer and
duced anthropogenically as well as by combustion, vegetation, (2) that a significant fraction of the mass of the sea-salt aero-
and oceanic phytoplankton [Hildemann et al., 1991; Pandis et
sols is organic material (and not CO2 or carbonate) [Middle-
Copyright 1999 by the American Geophysical Union. brook et al., 1998; Murphy et al., 1998a]. The best estimate from
Paper number 1999JD900073. these PALMS measurements is that ;10% of the marine aero-
0148-0227/99/1999JD900073$09.00 sols is organic matter. Organic compounds in marine boundary
11,633
11,634 ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS

ends of the long-chain surfactant are inward toward the ionic


brine core and the aliphatic, hydrophobic tails are facing out-
ward to the atmosphere. The world’s oceans are teeming with
life. Approximately 50% of the oxygen in our planet’s atmo-
sphere is produced by marine organisms (mostly phytoplank-
ton). All living organisms are made of carbon and all are
enclosed by plasma membranes. These biological membranes
are exceedingly complex structures composed of a mixture of
lipids, hydrophobic proteins, and carbohydrates. The exact
composition of these membranes varies considerably from or-
ganism to organism; however, the components of these mem-
branes are all hydrophobic. In particular, lipid membranes are
universally described as having a bilayer structure with polar
heads facing the aqueous medium and sets of hydrophobic,
alkyl chains folded inward as in Figure 1a [Roberts and Caserio,
1977; Gill et al., 1983; Alberts et al., 1989; Ables et al., 1992]. The
long hydrophobic tails of the lipid membranes are composed of
fatty acids (shown in Figure 1) and those most commonly
found are palmitic acid [CH3(CH2)14COOH] and stearic acid
[CH3(CH2)16COOH] [Bruice, 1995]. Phosphatidylcholines
(lecithins) such as shown in Figure 1b are the typical constit-
Figure 1. (a) Lipid bilayer membrane. (b) Phosphatidylcholine. uents of lipid membranes [Ables et al., 1992; Bruice, 1995].
Disintegration of the lipid bilayers in seawater at pH 8 will
produce phospholipids which slowly hydrolyze to fatty acids.
layer aerosols are expected to significantly affect their direct As cells die, they decompose and the hydrophobic cellular
and indirect radiative properties and, consequently, to play an constituents rise toward the ocean’s surface where they accu-
important, if difficult to assess, role in controlling the Earth’s mulate. The phospholipids and fatty acids (such as stearic acid,
climate [Saxena et al., 1995]. The recent ACE-1 results provide CH3(CH2)16COOH, for example) are common products of
the database needed to develop predictive models of radiative lipid membrane disintegration and will be important surfac-
and chemical properties of aerosols in the marine boundary tants. Because seawater has a pH of ;8, most fatty acids will
layer. The central difficulty with these high levels of observed exist as carboxylates [CH3(CH2)16CO2 2 ] and will be floating on
organic content in atmospheric aerosol is that they are inex- the surface of the ocean at the air/water interface [Quinn et al.,
plicable on the basis of bulk solubility (Henry’s law). 1972; Marty et al., 1979]. The seas cover 71% of the Earth and
We describe a model that proposes a chemical structure for have large regions of water coated with surfactants. Photosyn-
marine aerosols, accounts for the organic content, and de- thesis by marine organisms produces 30 – 60 3 1015 g of or-
scribes the chemical evolution or “atmospheric processing” of ganic carbon yr21 [Duarte and Cebrián, 1996]. Most of the
the organic material. Our model of an organic aerosol consists organic matter in the sea originates from phytoplankton pro-
of an aqueous core encapsulated by an inert, hydrophobic duction [Biddanda and Benner, 1997]. If one estimates that
organic monolayer, in much the same manner suggested 15 20% of the net primary production of carbon is exported from
years earlier [Gill et al., 1983]. However, Gill et al. concluded the ecosystem, this amounts to ;9 3 1015 g C yr21. If 5% of
that the inertness of this outer monolayer would minimize any the organic carbon is lipid and 1/2 of the exported lipids survive
participation in atmospheric chemistry or cloud microphysics. to rise to the surface of the ocean, one conjectures that the
We point out that this surface organic grease layer will react “biological oil slick” is about 1/2 3 5% 3 9 3 1015 g C yr21 or
with atmospheric hydroxyl radicals. The resulting surface film 2 3 1014 g C yr21 [Biddanda and Benner, 1997]. This is 200 Tg
is functionalized and rendered hydrophilic. The optical and C yr21 which will be “accumulated” or concentrated at the
chemical properties of organic aerosols as well as their ability ocean/atmosphere interface. More than 70% of the fatty acids
to nucleate will be dramatically enhanced as a result of the in seawater and the surface microlayer are present in esterified
chemistry we describe. Our proposed organic aerosol model form as triglycerides, stearyl esters, and phospholipids [Marty et
displays a complex set of properties that are consistent with al., 1979]. These estimates roughly corroborate our contention
recent field measurements [Middlebrook et al., 1998; Murphy et that a monolayer of hydrophobic organics will cover the sur-
al., 1998a]. Furthermore, this model makes a number of pre- face of the ocean and is available for incorporation on to the
dictions that may be amenable to test by field observations. surface of marine aerosols [Seinfeld et al., 1998].
Finally, there is an intriguing suggestion of a novel atmospheric Marine aerosol particles are formed by mechanical ejection
transport mechanism for organics and other water insoluble from the ocean’s surface, and as they form, they acquire a
molecules from the ocean into the atmosphere. coating of hydrocarbon surfactants [Pockels, 1891; Tseng et al.,
1992; O’Dowd and Smith, 1993; Russel et al., 1994]. There will
be water molecules present in these nascent aerosols, but as
2. Formation, Structure, and Composition of H2O evaporation continues and the inner core shrinks, the
Model Organic Marine Aerosols NaCl/H2O sphere will be covered with an organic layer with
Our picture of a marine aerosol is that of a brine core polar heads anchored to the “brine,” as sketched in Figure 2.
surrounded by an organic surfactant layer which is called an The binding energy of carboxylates (RCO2 2 ) to H2O clusters is
inverted micelle [Roberts et al., 1977; Gill et al., 1983; Adamson ;20 kcal mol21.
et al., 1997]. By “inverted” we mean that the polar, hydrophilic Our model for a marine aerosol is sketched in Figure 3 and
ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS 11,635

is a film of hydrophobic hydrocarbons covering a salt particle.


The model we propose emphasizes that all of the hydrophobic
organics will be on the surface of the aerosol and not dissolved
within the droplet. Aerosol size distributions have been mea-
sured during ACE 1 to be in the range of 0.02 and 0.8 mm with
concentrations of 500 cm23 for aerosols of 0.2 mm diameter
[Brechtel et al., 1998]. PALMS can characterize particles of
diameter $0.16 mm, which led us to choose 0.2 mm as the
diameter of the model aerosol described. The solubility of
individual alkanoic and alkenoic acids in water is miserably Figure 3. An inverted micelle model for a marine aerosol;
small [Streitweiser et al., 1992], and we estimate that each 0.2 the numbers are worked out for a representative 0.2 mm particle.
mm aerosol particle will contain no more than 25 dissolved
fatty acid molecules within the brine core (in a total of ;108
water molecules). In light of the negligible water solubility nascent aerosol is composed of “inert,” aliphatic hydrocarbon
(even lower in brine) of the functionalized long-chain hydro- chains that are exposed to the oxidizing troposphere.
carbons, we propose that the only structural picture of organic For illustrative purposes, we choose the overall aerosol di-
aerosols consistent with the large measured organic content is ameter to be ;0.2 mm as we explicitly wish to test the prop-
that the hydrophobic organics must reside at the exterior of the erties of our model against the ACE-1 field observations [Mur-
aerosol. Measurements of the temperature- and concentra- phy et al., 1997; Middlebrook et al., 1998; Murphy et al. 1998a, b].
tion-dependent surface tension of aqueous solutions of organic The core of the marine aerosol is a salt/water brine that will be
compounds have recently been used to determine standard- composed of about 1 3 108 molecules in a sphere 2000 Å in
free energies of adsorption from the vapor phase [Donaldson et diameter coated with a 25 Å deep film of about 6 3 105
al., 1998]. The results show that while the partitioning from surfactant molecules; recall that 1 Å [ 1028 cm 5 10210 m.
bulk to the surface is minor for soluble organics, water- The hydrophobic material (fatty acids and phospholipids) will
insoluble species stay on the surface. This result and the hy- be arranged on the surface of the aerosol (Figure 3) with the
drophobic nature of the surface will affect the growth and polar groups binding to the NaCl/H2O brine.
reactivity of droplets in the micrometer size range [Gill et al., If the nascent marine aerosol particles are properly de-
1983; Shulman et al., 1996]. scribed by Figure 3, then they will be chemically inert and will
The PALMS measurement that ;10% of marine aerosol not react with most atmospheric species (acid, base, H2O, O2,
mass is organic material [Middlebrook et al., 1998] is incompat- O3, SO2, NO, NO2, etc.). Because the marine aerosol is coated
ible with the notion that the organics are dissolved within the with hydrophobic molecules, water will not stick to it, and these
droplet. We suggest that the working model of a nascent ma- nascent marine aerosol particles will not nucleate water drop-
rine aerosol is that pictured in Figure 3; in it the exterior of the lets. The effect is large and is several orders of magnitude
[Rideal, 1925; Langmuir and Schaefer, 1943; Archer and La Mer,
1955]. Laboratory studies show that formation of organic sur-
factant layers on deliquescent salt particles effectively limit
water uptake by salt aerosols in atmospheric conditions, al-
though in much larger particles than we consider here [An-
drews and Larson, 1993; Wagner et al., 1996]. Measurements of
the loss rate of H2SO4 on to NaCl have shown that aerosols
coated with stearic acid resulted in significantly lower mass
accommodation coefficients, corroborating the predictions
presented here [Jefferson et al., 1997].

3. Chemical Processing of Organic Aerosols


Previous writers who arrived at an inverted micelle model of
organic aerosols pointed out that the organic coating will be
inert and optically inactive and concluded that their effect on
radiative transfer will be negligible [Gill et al., 1983]. We ex-
tend the inverted micelle idea by pointing out that the atmo-
sphere will react with the aerosol particle in Figure 3 and
transform it. Atmospheric hydroxyl radicals (and, to a much
lesser extent, Cl atoms) trigger a cascade of organic free radical
reactions on the aerosol which lead to processing of the inert
and hydrophobic organic coating. The chain reactions initiated
by OH radicals will oxidize the marine aerosol particles so that
they become coated with alcohol, aldehyde, ketone, and car-
Figure 2. Formulae for some common hydrophobic mole- boxylic acid functional groups. The thermochemistry of OH
cules that reside at the ocean’s surface and on the surface of reactions with hydrocarbons is exothermic. Because the bond
sea spray. Typical organic molecules such as the anions of energy for water is so high, DH 298 (HO™H) 5 119.30 6 0.05
palmitoleic acid, a phosphatidylcholine, and stearic acid are kcal mol21, hydroxyl radical can abstract H atoms from all
shown. hydrocarbons but alkynes (C§C™H) [Berkowitz et al., 1994]. For
11,636 ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS

but because DH 298 (HO2™H) is 87.1 6 0.6 kcal mol21, the


hydroperoxyl radical is a poor H atom abstracting reagent
[Clifford et al. 1998; Seinfeld and Pandis, 1998]. The important
reactions of O3 and NO3 with alkenes will have a major effect
on these organic films to the extent that unsaturated double
bonds (™CH¢CH™) exist; although, as seen above, OH reacts at
a considerable fraction of the collision rate at these double
bonds [Atkinson, 1994; Finlayson-Pitts and Pitts, 1997; de Gouw
Figure 4. Atmospheric processing of a 0.2 mm organic ma- and Lovejoy, 1998]. Commonly 50% of the mammalian phos-
rine aerosol.
pholipids contain an unsaturated ester (as shown in Figure 1b),
but the alkenes are generally buried along the alkyl chain. The
fraction of unsaturated phospholipids in phytoplankton is
example, H atom abstraction from the alkyl tail of a carboxylic likely higher than mammalian cells. This is because the fluidity
acid is exothermic by ;20 kcal mol21. of a plasma membrane is controlled by the ratio of saturated
CH3CH2(CH2)nCOOH 1 OH 3 H2O fatty esters to unsaturated esters [Alberts et al., 1989; Ables et al.
1992]. Mammalian membranes function at 378C, but phyto-
1 CH3CH(CH2)nCOOH (1) plankton live in the colder (188C) ocean; consequently, the
phytoplankton lipid bilayers will probably contain a higher
Similarly, it is exothermic by about 60 kcal mol21 for hydroxyl
fraction of alkenes to maintain the membrane fluidity at ocean
radicals to add to double bonds. For a long flexible molecule
temperatures. Note that while the product of ozonolysis of
such as an alkenoic acid, OH probably does not even need a
alkenes will be aldehydes and ketones, the product of reaction
third body to add.
with NO3 will be organic nitrates, mainly produced at night.
CH3CH¢CH(CH2)nCH2COOH Some oxidation of hydrophobic surfactants will occur right
at the ocean’s surface. Solar radiation between 280 and 320 nm
OH is reported to produce OH in the ocean by photolysis of dis-
u solved organic matter [Mopper and Zhou, 1990]. The concen-
1 OH 3 CH3CH™CH(CH2)nCH2COOH (2) tration of hydroxyl radical in the sea was indirectly measured
The kinetics of [OH 1 alkane] reactions has been studied and found to be r(OH) > 8 3 105 cm23 at the ocean’s surface.
and these processes are fast. As a model for hydroxyl 1 alkyl However, any reaction of hydrophobic surfactants with aque-
CH abstraction, CH3(CH2)n COOH 1 OH 3 carboxyalkyl ous OH radicals will steadily convert these hydrophobic hydro-
radical 1 H2O, consider [Atkinson, 1994; DeMore et al., 1997] carbons to water-soluble organics. So oxidation of hydrophobic
the following: hydrocarbons by aqueous OH simply removes them from the
ocean’s surface, and they are not available for incorporation
Reaction k(298 K)/cm3 s21 into the aerosol.
OH 1 CH3CH3 3 CH3CH2 1 H2O 3 3 10213 The alkyl radicals resulting from OH abstractions/addition
OH 1 CH3CH2CH3 3 CH3CHCH3 1 H2O 1 3 10212 at the surface of the aerosol R react rapidly with oxygen to
OH 1 CH3CH2CH2CH2CH3 3 4 3 10212 produce alklyperoxyl radicals; the high-pressure-limiting rate
CH3CH2CH2CHCH3 1 H2O
OH 1 CH3(CH2)5CH¢CH2 3 7 3 10211 constants will be k(R 1 O2) $ 10212 cm3 s21 [Lightfoot et al.,
CH3(CH2)5CH™CH2OH 1992].

We conjecture that tropospheric hydroxyl radicals will oxidize R 1 O2 3 RO2 (3)


the organic layer of marine aerosol particles with a rate con-
The surface-bound alkyl radicals will all react with atmo-
stant, k(OH 1 aerosol) $ 1 3 10212 cm3 s21.
spheric oxygen at an encounter-controlled rate to produce
The steady state rate of impingement of OH radicals onto an
surface-bound peroxyl radicals, aerosol/RO2. These surface-
aerosol surface can be estimated. The relationship between the
bound peroxyl radicals will rapidly be transformed into a vari-
flux of incident molecules striking a surface and the ambient
ety of oxygenated functional groups. As marine aerosol parti-
pressure is given by F/molecules cm22 s21 5 3.52 3 1022
cles rise from the ocean, the organic material is oxidized by
P(torr)(MT)21/2 [Somorjai, 1981]. For OH radicals (M 5 17
atmospheric reactions to produce alcohols (™OH), alkenes
amu) at 300 K, F 5 4.9 3 10 20 P(torr) cm22 s21, and the
(™C¢C™), aldehydes (™CHO), ketones (™CO™), and carboxylic
steady state concentration of hydroxyl radical in the tropo-
acids (™COOH). Consequently, the surface coating of the ma-
sphere is r(OH) > 106 cm23 or 2.8 3 10211 torr [Seinfeld and
rine particle in Figure 3 is “functionalized” and changed, as
Pandis, 1998]. Consequently, the steady state flux is F(OH, 300
suggested by Figure 4.
K) 5 1.4 3 1010 cm22 s21. An aerosol “sphere” that is 0.2 mm
in diameter has a surface area of p D 2 or 1.3 3 1029 cm2, while
the volume is p D 3 /6 or 4.2 3 10215 cm3. Thus the aerosol
particle consists of a core of 1.3 3 108 molecules covered by 4. Discussion and Conclusions
;6.3 3 105 molecules. In a single second a 0.2 mm aerosol is The model organic aerosol proposed and its atmospheric
struck by ;15 hydroxyl radicals, which implies that in 3.5 3 104 processing (see Figure 4) lead to some definite predictions
s (10 hours), every molecule on the aerosol’s surface will suffer concerning its size distribution, aging process, and radiative
a collision with hydroxyl. The rate constants used in this cal- and chemical properties, all of which can be tested by atmo-
culation are approximate and do not justify a more sophisti- spheric measurement.
cated kinetic analysis suggested by Fuchs and Sutugin [1970]. Primary sea-salt aerosols are produced by the mechanical
There is 100 times more HO2 in the troposphere than OH, disruption of the ocean surface. The intermediate size particles
ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS 11,637

(0.08 –1 mm diameter) are especially important in influencing


climate since they are involved in radiative transfer, directly by
their ability to absorb and scatter solar radiation and indirectly
by their role in producing condensation nuclei and making
clouds more reflective [Twomey et al., 1984; Charlson et al.,
1992; Novakov and Penner, 1993; Quinn et al., 1996; Rivera-
Caprio et al., 1996; Schwartz, 1996; Tegen et al., 1997; Liao and
Seinfeld, 1998; Pan et al., 1998]. Individual marine particles of
this intermediate size (diameter $ 0.16 mm) have been studied
in situ only recently via particle analysis by PALMS and have
been shown to contain a significant organic component (5–
50% by mass) [Murphy et al., 1997; Middlebrook et al., 1998;
Murphy et al., 1998a, b].
We can make a crude estimate of “organic mass transfer”
from the surface of the ocean into the troposphere for the
marine aerosol pictured in Figure 3. Fatty acids are the main
hydrophobic components of lipid membranes, and those most
commonly found are palmitic acid [CH3(CH2)14COOH] and
stearic acid [CH3(CH2)16COOH]. Consider the case of an
aerosol coated with stearate. If the aerosol particle is 0.2 mm in
diameter, then it has an area of 1.3 3 1029 cm2 and a volume Figure 5. A graph of the variation of the (carbon film)/(total
of about 4.2 3 10215 cm3. Such a particle contains ;1.3 3 108 aerosol) mass ratio as a function of the diameter of a spherical
aerosol [Middlebrook et al., 1998; Gill et al., 1983].
molecules of water/salt that make up the brine core. Because
stearate [CH3(CH2)16CO2 2 ] has only a single polar end, it can
bind only a single monolayer to the brine core. The areal size
(“footprint”) of a fatty acid molecule bound on a water surface, aerosol particles, which are the survivors of the drops that are
the “Pockels limit,” is ;20 6 2 Å2 [Pockels, 1891; Adamson formed by mechanical ejection from the ocean’s surface. These
and Gast, 1997]. The molecular weight of CH3(CH2)16CO2 2 is drops shed water molecules until hydrocarbon encapsulation
283 amu; consequently, the “load” of stearate carried away intervenes. Our model clearly suggests that the organic species
from the ocean by a 0.2 mm aerosol particle into the tropo- will preferentially reside on the surface of intermediate size
sphere will be (1.3 3 1029/[20 Å2])(283/6.02 3 1023) or about (0.1–1 mm) and large ($1 mm) aerosols.
3 3 10216 g. If we assume that most of the “brine core” is Processing of the inert, organic surface layer can only be
water, then the entire aerosol has an approximate weight of initiated by reactions with atmospheric OH (and, to a lesser
about 5 3 10215 g. Consequently, the mass ratio of the (or- extent, Cl atoms). Reaction rates at the interface are thought
ganic layer)/(entire aerosol) of a 0.2 mm marine aerosol par- to be fast but are not known quantitatively. We estimated
ticle is ;5%. This mass ratio of the (organic layer)/(entire earlier that aerosol processing in the atmosphere will take
aerosol) will change with the size of the aerosol particle with several hours/days. Convective clouds can transport insoluble
diameter D and will vary as D 21 . Organics will be about 50% materials from low altitude to the upper troposphere and lower
of the weight of small particles (0.01 mm diameter), while stratosphere in a few hours during deep convection in the
larger ones (1 mm diameter) will have only 1% organic mass. tropics [Gidel, 1983; Chatfield and Crutzen, 1984; Pickering et
Our estimated mass ratio of the (organic layer)/(entire aero- al., 1992; Danielsen, 1993; Kritz et al., 1993]. Our model pre-
sol) for a 0.2 mm aerosol of 5% is roughly that reported by the diction, superimposed on the timescale for vertical transport in
PALMS field measurements [Middlebrook et al., 1998]. Figure the tropics, suggests that at least some of this population of
5 is a sketch of the fraction of the aerosol’s mass belonging to organic aerosols could be transported up into the stratosphere.
the hydrocarbon film. We have plotted the percent carbon We emphasize here that the timescales are approximate, be-
versus the diameter of the aerosol. The solid experimental cause of the paucity of quantitative data for the relevant rate
circles are from Gill et al. [1983], and the open circle is the coefficients. However, a very interesting competition is likely
PALMS measurement [Middlebrook et al., 1998]. to exist between the chemical processing time and the tropo-
The ocean surface is covered with a “biological oil slick,” spheric residence time (estimated to be 5–15 days due to rain
and water evaporation will be suppressed by the low vapor scavenging) [Balkanski et al., 1993] of the organic aerosols.
pressure organic film [Rideal, 1925; Gershy, 1983]. Conse- Our model organic aerosol sketched in Figure 3 provides a
quently, sea-spray-generated aerosols will shrink in size by possible vehicle for transport into the atmosphere of water-
continuously ejecting H2O molecules up to the point where a soluble species trapped in the “brine core” as well as water-
complete hydrocarbon surface monolayer encapsulates the insoluble substances “dissolved” in the encapsulating organic
aerosol. At this point, water evaporation ceases, and the aero- film. This “bus” would also deliver smaller organics (such as
sol no longer shrinks in size. For the marine aerosols charac- acetone, OCS, methyl halides, simple alcohols, small alde-
terized by PALMS (diameter . 0.16 mm) the measured 10% hydes) into the upper troposphere [Singh et al., 1995; Kotch-
organic mass corresponds to a fatty acid surface monolayer enruther et al., 1999]. In particular, the “bus” might provide a
(Figure 5) [Middlebrook et al., 1998]. One prediction of our natural means to transport CH3I into the atmosphere from the
model is that marine aerosols will exhibit a multimode size ocean, and the PALMS field measurements commonly find
distribution. The simplest view would be that there will be iodide associated with organic fragments [Murphy et al., 1997;
small sulfate aerosol particles, which originate by the clustering Middlebrook et al., 1998]. Of course, the oxidative destruction
of water about a nucleation point, and intermediate to big of the hydrophobic film that encapsulates the brine core by
11,638 ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS

OH, O2, NO will shred the large alkyl chains and will produce of the nascent aerosol has no accessible electronic states (elec-
small carbon fragments which will be ejected into the tropo- tronic states of such compounds can only be reached in the
sphere. vacuum ultraviolet) [Ellison, 1974; Robin, 1985]. However, the
Much of the OH, O2, NO surface chemistry can be antici- processed aerosol organic coating of alcohols, aldehydes, and
pated by analogy to well-known gas phase homogeneous chem- ketones have near-ultraviolet states, making these compounds
istry, but there surely are chemical surprises awaiting us [At- effective atmospheric chromophores. In the near IR, very weak
kinson, 1994]. As a possible example, notice that RO2 radicals vibrational C™H overtones (from the organic coating) and
readily oxidize NO to NO2 with an exothermicity of 212 kcal O™H overtones (from the water core) can be excited. The key
mol21 (250 kJ mol21); the rate constant for peroxyl radical proposed intermediates in processing organic aerosols are sur-
oxidation of NO is ;8 3 10212 cm3 s21 [Eberhard and Howard, face-bound RO2 radicals which have been shown to have al-
1997]. For ozone and sulfur dioxide, peroxide oxidation reac- lowed electronic transitions in the near IR (7000 –9000 cm21)
tions are even more exothermic. [Clifford et al., 1998]. While the cross sections of the surface-
bound species are not known, they are expected, as befitting
(R4a) CH3O2(2A0) 1 NO 3 CH3O(2E) 1 NO2 allowed electronic states, to be 104 to 106 times larger than that
D rxnH 298 5 212 kcal mol21 of vibrational overtones. Scaling for the number of molecules
in the core versus those on the surface and explicitly using the
(R4b) CH3O2(2A0) 1 O3(1A1) 3 CH3O(2E) 1 2O2 differences in cross sections, we expect the aerosol with surface
RO2 to be 2– 4 orders of magnitude more absorbent in the near
D rxnH 298 5 232 kcal mol21 IR than comparably sized “uncoated” or “unprocessed” ma-
rine aerosols.
(R4c) CH3O2(2A0) 1 SO2 3 CH3O(2E) 1 SO3
Most importantly, processing of organic aerosols provides
D rxnH 298 5 222 kcal mol21 “defect sites” on the aerosol surface where nucleation rates
can be efficient [Thomas et al., 1998]. Consequently, processed
Yet the gas phase (R4b) and (R4c) do not go at room tem- aerosols can form condensation nuclei, scatter light more ef-
perature [Tyndall et al., 1998]. However, the kinetics on an fectively, and increase the reflectivity of clouds. The size of
aerosol surface of (R4b) peroxyl radical oxidation of O3 to these processed organic aerosols will be a function of the
generate alkoxy radicals or (R4c) with SO2 to produce organic aerosol age. Nucleation in this case occurs entirely indepen-
sulfates might be another story. We note that current field dently of the sulfate content of the aerosol. The model result in
measurements of SO2 oxidation to sulfate could not distinguish Figure 5 can be applied to the recent ACE-1 field observations
between this route and the other oxidation paths in the bulk that pure sulfate particles larger than 0.13 mm at Cape Grim
liquid phase. were extremely scarce and that few pure sulfate particles were
The temperature drop from the tropical ocean surface to the acting as cloud condensation nuclei in that region [Murphy et
tropopause (17 km above) is 1158C. Any hydrocarbons that al., 1998a]. The model presented here for processing of organic
boil in the range 2858C , T , 308C will form inverted micelles aerosols provides an alternative path to aerosol nucleation and
on any aerosol en route. The vertical transport will probably formation of cloud condensation nuclei.
occur mainly in the Intertropical Convergence Zone (ITCZ). We consider some potential objections to our picture in this
The D 21 dependence of the organic content implied by the paragraph. To the suggestion that gas phase coagulation of
inverted micelle model suggests that the aerosols in the upper smaller particles could account for the organic content of the
tropical troposphere will contain a high fraction of organic actinically active aerosol sizes, we respond that if it did, our
material (Figure 5). This prediction is testable by observation. model would still constitute a viable description. We do not
Once in the stratosphere, the aerosol particle will warm, and accept, however, that surfactant molecules could accrete at the
some of the lighter organics will evaporate. This is another way surface of a particle by collisions from the gas phase; the vapor
that aerosols could act as a “bus” to transport organics to the pressure of such molecules is far too low. For example, the
stratosphere via the ITCZ. We expect that the inverted micelle boiling point of stearic acid [CH3(CH2)16COOH] is listed as
could “protect” the marine core in Figure 3 and allow it to also 1838–1848C at 1 torr vapor pressure; the boiling point at atmo-
reach the stratosphere. The overwhelming balance of evidence spheric pressure is listed as 3608C, accompanied by decompo-
suggests that there is little marine Cl in the middle and upper sition. Because of the low vapor pressure, nor do we consider
stratosphere. Russell et al. [1996] show that nearly all the Cl oxidation of surfactants in the gas phase by the chemistry we
and F in the upper stratosphere comes from CFCs. However, describe.
Rosenlof et al. [1997] find that most of the air crossing the inner There are some obvious questions and complexities associ-
tropical tropopause (100 mbar) goes sideways instead of up- ated with our model. For example, in the real atmosphere the
ward: the flux up through 50 mbar is only about 10% of the surface film is likely to be made up of several to many different
mass that goes up through 100 mbar. So, the lowermost strato- molecular species, raising the question of how ordered
sphere, where the greatest ozone losses are, could be affected (aligned) the individual molecules are. It has long been known
by a flux of “protected” brine. There are very few measure- that a single double bond, particularly for the cis isomer, can
ments in the lower stratosphere that could be used in this affect surfactant layer thickness [Langmuir, 1917]. The degree
context, nevertheless, one example of high chloride (0.4 ppbv of functionalization as radical processing proceeds could also
of HCl vapor) is available [Tuck et al., 1995]. have major effects in this respect. It has been asserted that
The direct and indirect aerosol effect on radiative transfer removal of a single molecule could provide a “hole” through
can be predicted as a consequence of this chemical model for which water could exchange between the core and the atmo-
aerosol processing. The processed, “aged” aerosol contains sphere, to the extent that 90 –99% of the inhibition by the film
chromophores which will be significantly more absorbing than could be lost [Barger and Garrett, 1970]. Indeed, in an ordered
their aliphatic precursors. The saturated hydrocarbon coating spherical film, the “hairy ball theorem” necessitates the exis-
ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS 11,639

tence of a whorl point in mapping rods on a sphere. However, Acknowledgments. GBE is supported by a grant from the Chemi-
real atmospheric films are more likely to resemble a bramble cal Physics Program, U.S. Department of Energy (DE-FG02-
87ER13695). VV would like to acknowledge a CRCW Faculty Fellow-
patch than a wheat field, and such “holes” could be rare. It is ship from the University of Colorado and the National Science
natural to ask if such films would exist on terrestrial as well as Foundation for continued support. We are indebted to our NOAA
marine aerosol. One mechanism which could volatilize the colleagues Dan Murphy and Anne Middlebrook for sharing their data
waxes and lipids from plants would be biomass burning; there with us and for many fine pointers about their PALMS technique. We
must be a zone of temperatures around flames which will acknowledge many of our colleagues for extensive discussions:
Meredithe L. Applebury (Harvard Medical School), Roberto Bianco
correspond to their boiling points. Terrestrial plant lipids have (University of Colorado), Jamie Donaldson (University of Toronto),
been detected in atmospheric aerosol samples. We believe that Ray Fall (University of Colorado), Greg Frost (NOAA), Vicki Gras-
the radical processing of the exterior surfactant film from a sian (University of Iowa), Anne Jefferson (NOAA), Bruce E. Koel
hydrophobic to a hydrophilic state offers a further means by (USC), Carl A. Koval (University of Colorado), Ron Naaman (Weiz-
mann Institute), Suzanne E. Paulson (UCLA), Stuart Penkett (Uni-
which atmospheric chemistry can affect the radiative balance. versity of East Anglia), Yinon Rudich (Weizmann Institute), Sally A.
It can do so directly via the effect of the aerosol on the radi- Sullivan (Greenlee, Winner, & Sullivan, PC), Pieter Tans (NOAA),
ation field and indirectly via the aerosol’s cloud nucleation and Mark Young (University of Iowa).
properties. Because each hydrogen atom on, say, stearic acid,
CH3(CH2)16COOH, is potentially transformable to an OH References
free radical, the effect could be nonlinear. It will certainly be a Ables, R. H., P. A. Frey, and W. P. Jencks, Biological membranes,
function of latitude, since the greatest concentrations of atmo- chap. 29, in Biochemistry, Jones and Bartlett, Boston, Mass., 1992.
spheric OH occur near the surface in the tropics. The process- Adamson, A. W., and A. P. Gast, Physical Chemistry of Surfaces, John
ing will be much slower in the southern oceans, for example, Wiley, New York, 1997.
Alberts, B., D. Bray, J. Lewis, M. Raff, K. Roberts, and J. D. Watson,
where aerosol-producing high wind speeds are accompanied by The plasma membrane, chap. 6, in Molecular Biology of the Cell,
much lower actinic light levels. As air moves up from the Garland, New York, 1989.
surface toward the tropical tropopause in the ITCZ, cooling Andreae, M. O., Future climates of the world, in World Survey of
from 308 to 2858C, new aerosol films will form as the air cools Climatology, edited by A. Henderson-Sellers, pp. 341–392, Elsevier,
New York, 1995.
through the range of boiling points of the airborne organic Andreae, M. O., and P. J. Crutzen, Atmospheric aerosols: Biogeo-
molecules. Combining this fact with the D21 size dependence, chemical sources and role in atmospheric chemistry, Science, 276,
we predict that aerosols in the upper tropical troposphere will 1052–1058, 1997.
contain a high fraction of organic molecules. Finally, we note Andrews, E., and S. M. Larson, Effect of surfactant layers on the size
that the actual effects in the atmosphere, both radiative and changes of aerosol particles as a function of relative humidity, En-
viron. Sci. Technol., 27, 857– 865, 1993.
chemical, will depend on the actual fluxes of the organic mol- Archer, R. J., and V. K. La Mer, The rate of evaporation of water
ecules, both from the planetary surfaces and to and from the through fatty acid monolayers, J. Phys. Chem., 59, 200 –208, 1955.
aerosols as they move upward in the ITCZ. The data do not Atkinson, R., Gas-Phase Tropospheric Chemistry of Organic Com-
currently exist to quantify these fluxes. pounds, Am. Inst. of Phys., Woodbury, N. Y., 1994.
Balkanski, Y. J., D. J. Jacob, G. M. Gardner, W. C. Graustein, and
We would like to point out that the model presented is K. K. Turekian, Transport and residence times of tropospheric aero-
applicable not only to marine boundary layer aerosols but, sols inferred from a global three-dimensional simulation of 210Pb, J.
more generally, is appropriate for all atmospheric aerosols Geophys. Res., 98, 20,573–20,586, 1993.
containing a significant organic component (especially those Barger, W. R., and W. D. Garrett, Surface active organic material in
with a polar core). The application of this model hinges on the the marine atmosphere, J. Geophys. Res., 75, 4561– 4566, 1970.
Berkowitz, J., G. B. Ellison, and D. Gutman, Three methods to mea-
availability of atmospheric measurements. Such measurements sure RH bond energies, J. Phys. Chem., 98, 2744 –2765, 1994.
have recently been made and have shown organic compounds Biddanda, B., and R. Benner, Carbon, nitrogen, and carbohydrates
to be significant in aerosols in continental sites. As an example, fluxes during the production of particulate and dissolved organic
the nature of the organic coating is important, as shown by the matter by marine phytoplankton, Limnol. Oceanogr., 42, 506 –518,
1997.
hygroscopic behavior of organic aerosols [Saxena et al., 1995]. Brechtel, F. J., S. M. Kreidenweis, and H. B. Swan, Air mass charac-
For nonurban locations (Grand Canyon), organic coatings en- teristics, aerosol particle number concentrations, and number size
hance water absorption by inorganics; organics account for distributions at Macquarie Island during the First Aerosol Charac-
25– 40% of the total water uptake (on average). In urban terization Experiment (ACE 1), J. Geophys. Res., 103, 16,351–16,367,
locations (Los Angeles) the net effect of organics is to diminish 1998.
Bruice, P. Y., Organic Chemistry, chap. 26, Prentice-Hall, Engelwood
water absorption of the inorganics by 25–35% [Rogge et al., Cliffs, N. J., 1995.
1993]. This suggests that some organic layers are hydrophilic Charlson, R. J., J. E. Lovelock, M. O. Andreae, and S. G. Warren,
(Grand Canyon) and others are hydrophobic (Los Angeles). Ocean phytoplankton, atmospheric sulphur, cloud albedo and cli-
Organic compounds in the atmosphere are generated by the mate, Nature, 326, 655– 661, 1987.
Charlson, R. J., S. E. Schwartz, J. M. Hales, R. D. Cess, J. A. Coakley
biosphere and, to some extent, anthropogenically. Charlson Jr., J. E. Hansen, and D. J. Hofmann, Climate forcing by anthropo-
et al. [1987] have presented a case for dimethyl sulfide genic aerosols, Science, 255, 423– 430, 1992.
(CH3SCH3), which is generated in the ocean to form sulfate Chatfield, R. B., and P. J. Crutzen, Sulfur dioxide in remote oceanic
aerosols, and that organic sulfates establish a feedback in the air: Cloud transport to reactive precursors, J. Geophys. Res., 89,
climate system. The model proposed here gives a mechanism 7111–7132, 1984.
Clifford, E. P., P. G. Wenthold, R. Gareyev, W. C. Lineberger, C. H.
for organic aerosols to be processed and to affect aerosol DePuy, V. M. Bierbaum, and G. B. Ellison, Photoelectron spectros-
nucleation independently of their sulfur content. The conse- copy, gas phase acidity, and thermochemistry of tert-butyl hydroper-
quence of atmospheric processing of organic aerosols is to oxide: Mechanism for the rearrangement of peroxyl radicals,
provide a direct and general connection between the biosphere J. Chem. Phys., 109, 10,293–10,310, 1998.
Danielsen, E. F., In situ evidence of rapid vertical, irreversible trans-
and the atmosphere and establishes a mechanism by which the port to lower tropospheric air into the lower stratosphere by con-
former can affect atmospheric radiative transfer and the vective cloud turrets and by large-scale upwelling in tropical cy-
Earth’s climate. clones, J. Geophys. Res., 98, 8665– 8681, 1993.
11,640 ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS

de Gouw, J. A., and E. R. Lovejoy, Reactive uptake of ozone by liquid and its potential impact on marine processes, Science, 250, 661– 664,
organic compounds, Geophys. Res. Lett., 25, 931–934, 1998. 1990.
DeMore, W. B., S. P. Sander, D. M. Golden, M. J. Molina, R. F. Murphy, D. M., D. S. Thomson, and A. M. Middlebrook, Bromide,
Hampson, M. J. Kurylo, C. J. Howard, A. R. Ravishankara, C. E. iodine, and chlorine in single aerosol particles at Cape Grim, Geo-
Kolb, and M. J. Molina, Chemical Kinetics and Photochemical Data phys. Res. Lett., 24, 3197–3200, 1997.
for Use in Stratospheric Modeling, in Evaluation 12, NASA, 1997. Murphy, D. M., J. R. Anderson, P. K. Quinn, L. M. McInnes, R. J.
Dickerson, R. R., S. Kondragunta, G. Stenchikov, K. L. Civerolo, B. G. Brechtel, S. M. Kreidenweis, A. M. Middlebrook, M. Pósfai, D. S.
Doddridge, and B. N. Holben, The impact of aerosols on solar Thomson, and P. R. Buseck, Influence of sea-salt on aerosol radia-
ultraviolet radiation and photochemical smog, Science, 278, 827– tive properties in the Southern Ocean marine boundary layer, Na-
830, 1997. ture, 392, 62– 65, 1998a.
Donaldson, D. J., and D. Anderson, Adsorption of atmospheric gases Murphy, D. M., D. S. Thomson, A. M. Middlebrook, and M. E. Schein,
at the air-water interface, I, C1–C4 alcohols, acids, and acetone, J. In situ single particle characterization at Cape Grim, J. Geophys.
Phys. Chem. A, in press, 1998. Res., 103, 16,485–16,491, 1998b.
Duarte, C. M., and J. Cebrián, The fate of marine autotrophic pro- Novakov, T., and J. E. Penner, Large contribution of organic aerosols
duction, Limnol. Oceanogr., 41, 1758 –1766, 1996. to cloud-condensation-nuclei concentrations, Nature, 365, 823– 826,
Eberhard, J., and C. J. Howard, Rate coefficients for the reactions of 1993.
some C3 to C5 hydrocarbon peroxy radicals with NO, J. Phys. Chem., Novakov, T., C. E. Corrigan, J. E. Penner, C. C. Chuang, O. Rosario,
101, 3360, 1997. and O. L. M. Bracero, Organic aerosols in the Caribbean trade
Ellison, G. B., Electronic states of olefins, Ph.D., thesis, Yale Univ., winds: A natural source?, J. Geophys. Res., 102, 21,307–21,313, 1997.
New Haven, Conn., 1974. O’Dowd, C. D., and M. H. Smith, Physicochemical properties of aero-
Finlayson-Pitts, B. J., and J. N. Pitts Jr., Tropospheric air pollution: sols over the Northeast Atlantic: Evidence for wind-speed-related
Ozone, airborne toxics, polycyclic aromatic hydrocarbons, and par- submicron sea-salt aerosol production, J. Geophys. Res., 98, 1137–
ticles, Science, 276, 1045–1051, 1997. 1149, 1993.
Fitzgerald, J. W., Marine aerosols: A review, Atmos. Environ., 25A, Odum, J. R., T. P. W. Jungkamp, R. J. Griffin, R. C. Flagan, and J. H.
533–545, 1991. Seinfeld, The atmospheric aerosol-forming potential of whole gas-
Fuchs, N. A., and A. G. Sutugin, Highly Dispersed Aerosols, Butter- oline vapor, Science, 276, 96 –99, 1997.
worth-Heinemann, Newton, Mass., 1970. Pan, W., M. A. Tatang, G. J. McRae, and R. G. Prinn, Uncertainty
Gershy, R. M., Characterization of seawater organic matter carried by analysis of indirect radiative forcing by anthropogenic sulfate aero-
bubble-generated aerosols, Limnol. Oceanogr., 28, 309 –319, 1983. sols, J. Geophys. Res., 103, 3815–3823, 1998.
Gidel, L. T., Cumulus cloud transport of transient tracers, J. Geophys. Pandis, S. N., S. E. Paulson, J. H. Seinfeld, and R. C. Flagan, Aerosol
Res., 88, 6587– 6599, 1983. formation in the photooxidation of isoprene and b-pinene, Atmos.
Gill, P. S., T. E. Graedel, and C. J. Weschler, Organic films on atmo- Environ., 25A, 997–1008, 1991.
spheric aerosol particles, fog droplets, cloud droplets, raindrops, and Pandis, S. N., A. S. Wexler, and J. H. Seinfeld, Dynamics of tropo-
snowflakes, Rev. Geophys., 21, 903–920, 1983. spheric aerosols, J. Phys. Chem., 99, 9646 –9659, 1995.
Hildemann, L. M., M. A. Mazurek, G. R. Cass, and B. R. T. Simoneit,
Pickering, K., A. M. Thompson, J. R. Scala, W. K. Tao, and J. Simpson,
Quantitative characterization of urban sources of organic aerosol by
Ozone production potential following convective redistributions of
high-resolution gas chromatography, Environ. Sci. Technol., 25,
biomass burning emissions, J. Atmos. Chem., 14, 297–313, 1992.
1311–1325, 1991.
Pockels, A., Surface tension, Nature, 43, 437– 439, 1891.
Jefferson, A., F. L. Eisele, P. J. Ziemann, R. J. Weber, J. J. Marti, and
Quinn, J. G., and T. L. Wade, Lipid measurements in the marine
P. H. McMurry, Measurements of the H2SO4 mass accommodation
atmosphere and sea surface microlayer, in Baseline Studies of Pol-
coefficient on to polydisperse aerosol, J. Geophys. Res., 102, 19,021–
lutants in the Marine Environment, edited by E. D. Goldberg, Natl.
19,028, 1997.
Sci. Found., Washington, D. C., 1972.
Jennings, S. G., M. Geever, F. M. McGovern, J. Francis, T. G. Spain,
Quinn, P. K., V. N. Kapustin, T. S. Bates, and D. S. Covert, Chemical
and T. Donaghy, Microphysical and physico-chemical characteriza-
tion of atmospheric marine and continental aerosol at Mace Head, and optical properties of marine boundary layer aerosol particles of
Atmos. Environ., 31, 2795–2808, 1997. the mid-Pacific in relation to sources and meteorological transport,
Kotchenruther, R. A., P. V. Hobbs, and D. A. Hegg, Humidification J. Geophys. Res., 101, 6931– 6951, 1996.
factors for atmospheric aerosols off the mid-Atlantic coast of the Rideal, E. K., On the influence of thin surface films on the evaporation
United States, J. Geophys. Res., 104, 2239 –2251, 1999. of water, J. Phys. Chem., 29, 1585–1588, 1925.
Kritz, M. A., S. W. Posner, K. K. Kelly, M. Loewenstein, and K. R. Rivera-Caprio, C. A., C. E. Corrigan, T. Novakov, J. E. Penner, C. F.
Chan, Radon measurements in the lower tropical stratosphere: Ev- Rogers, and J. C. Chow, Derivation of contributions of sulfate and
idence of rapid vertical transport and dehydration of tropospheric carbonaceous aerosols to cloud condensation nuclei from mass size
air, J. Geophys. Res., 97, 8725–9736, 1993. distributions, J. Geophys. Res., 101, 19,483–19,493, 1996.
Langmuir, I., The constitution and fundamental properties of solids Roberts, J. D., and M. C. Caserio, Basic Principles of Organic Chem-
and liquids, J. Am. Chem. Soc., 39, 1848 –1905, 1917. istry, section 18.2, Benjamin, White Plains, N. Y., 1977.
Langmuir, I., and V. J. Schaefer, Rates of evaporation of water Robin, M. B., Higher Excited States of Polyatomic Molecules, Academic,
through compressed monolayers on water, J. Franklin Inst., 235, San Diego, Calif., 1985.
4561– 4566, 1943. Rogge, W. F., M. A. Mazurek, L. M. Hildemann, G. R. Cass, and
Li, Z., A. L. Williams, and M. J. Rood, Influence of soluble surfactant B. R. T. Simoneit, Quantification of urban organic aerosols at a
properties on the activation of aerosol particles containing inorganic molecular level: Identification, abundance, and seasonal variation,
solute, J. Atmos. Sci., 55, 1859 –1866, 1998. Atmos. Environ., 27A, 1309 –1330, 1993.
Liao, H., and J. H. Seinfeld, Effect of clouds on direct aerosol radiative Rosenlof, K. H., A. F. Tuck, K. K. Kelly, J. M. Russell III, and M. P.
forcing of climate, J. Geophys. Res., 103, 3781–3788, 1998. McCormick, Hemispheric asymmetries in water vapor and infer-
Lightfoot, P. D., R. A. Cox, J. N. Crowley, M. Destriau, G. D. Hayman, ences about transport in the lower stratosphere, J. Geophys. Res.,
M. E. Jenkin, G. K. Moortgat, and F. Zabel, Organic peroxy radi- 102, 13,213–13,234, 1997.
cals: Kinetics, spectroscopy and tropospheric chemistry, Atmos. En- Russel, L. M., S. N. Pandis, and J. H. Seinfeld, Aerosol production and
viron., 26A, 1805–1961, 1992. growth in the marine boundary layer, J. Geophys. Res., 99, 20,989 –
Marty, J. C., A. Saliot, P. Buat-Ménard, R. Chesselet, and K. A. 21,003, 1994.
Hunter, Relationship between the lipid compositions of marine Russell, J. M., III, M. Luo, R. J. Cicerone, and L. E. Deaver, Satellite
aerosols, and the sea surface microlayer, and subsurface water, J. confirmation of the dominance of chlorofluorocarbons in the global
Geophys. Res., 84, 5707–5716, 1979. stratospheric chlorine budget, Nature, 379, 526 –529, 1996.
Middlebrook, A. M., D. M. Murphy, and D. S. Thomson, Observations Saxena, P., and L. M. Hildemann, Water soluble organics in atmo-
of organic material in individual marine particles at Cape Grim spheric particles: A critical review of the literature and application
during the First Aerosol Characterization Experiment (ACE 1), J. of thermodynamics to identify candidate compounds, J. Atmos.
Geophys. Res., 103, 16,475–16,483, 1998. Chem., 24, 57–109, 1996.
Mopper, K., and X. Zhou, Hydroxyl radical photoproduction in the sea Saxena, P., L. M. Hildemann, P. H. McMurry, and J. H. Seinfeld,
ELLISON ET AL.: ATMOSPHERIC PROCESSING OF ORGANIC AEROSOLS 11,641

Organics alter hygroscopic behavior of atmospheric particles, J. Geo- Tseng, R.-S., J. T. Viechnicki, R. A. Skop, and J. W. Brown, Sea-to-air
phys. Res., 100, 18,755–18,770, 1995. transfer of surface-active organic compounds by bursting bubbles, J.
Schwartz, S. E., The whitehouse effect–shortwave radiative forcing of Geophys. Res., 97, 5201–5206, 1992.
climate by anthropogenic aerosols: An overview, J. Aerosol Sci., 27, Tuck, A. F., C. R. Webster, R. D. May, D. C. Scott, S. J. Hovde, J. W.
359 –382, 1996. Elkins, and K. R. Chan, Time and temperature dependences of
Seinfeld, J. H., and S. N. Pandis, Atmospheric Chemistry and Physics: fractional HCl abundances from airborne data in the southern hemi-
From Air Pollution to Climate Change, section 5, John Wiley, New sphere during 1994, Faraday Discuss., 100, 389 – 410, 1995.
York, 1998. Twomey, S., M. Piepgrass, and T. L. Wolfe, An assessment of the
Shulman, M. L., M. Z. Jacobson, R. J. Charlson, R. E. Synovec, and impact of pollution on global cloud albedo, Tellus, Ser. B, 36, 356 –
T. E. Young, Dissolution behavior and surface tension effects of 366, 1984.
organic compounds in nucleating cloud droplets, Geophys. Res. Lett., Tyndall, G. S., T. J. Wallington, and J. C. Ball, FTIR product study of
23, 277–280, 1996. the reactions CH3O2 1 CH3O2 and CH3O2 1 O3, J. Phys. Chem.,
Singh, H. B., M. Kanakidou, P. J. Crutzen, and D. J. Jacob, High 102, 2547–2554, 1998.
concentrations and photochemical fate of oxygenated hydrocarbons Wagner, J., E. Andrews, and S. M. Larson, Sorption of vapor phase
in the global troposphere, Nature, 378, 50 –54, 1995. octanoic acid onto deliquescent salt particles, J. Geophys. Res., 101,
Somorjai, G. A., Chemistry in Two Dimensions, p. 27, Cornell Univ. 19,533–19,540, 1996.
Press, New York, 1981.
Streitweiser, A., C. H. Heathcock, and E. M. Kosower, Introduction to G. B. Ellison and V. Vaida, Department of Chemistry & Biochem-
Organic Chemistry, Table 18.2, Macmillan, New York, 1992. istry, University of Colorado, Boulder, CO 80309-0215. (barney@jila.
Tegen, I., P. Hollrig, M. Chin, I. Fung, D. Jacob, and J. Penner, colorado.edu; vaida@colorado.edu)
Contribution of different aerosol species to the global aerosol ex- A. F. Tuck, Meteorological Chemistry Program, NOAA Aeronomy
tinction optical thickness: Estimates from model results, J. Geophys. Laboratory, R/E/AL6, 325 Broadway, Boulder, CO 80303.
Res., 102, 23,895–23,915, 1997. (tuck@al.noaa.gov)
Thomas, E., S. Trakhtenberg, R. Ussyshkin, and Y. Rudich, Water
adsorption by hydrophobic organic surfaces: Experimental evidence
and implications on the atmospheric properties of organic aerosols, (Received October 7, 1998; revised January 19, 1999;
J. Geophys. Res., in press, 1998. accepted January 21, 1999.)
11,642

Potrebbero piacerti anche