kinetic limitations on cloud droplet formation and impact

14
1 Appeared in Tellus B , 53 , 133–149, 2001. Kinetic Limitations on Cloud Droplet Formation and Impact on Cloud Albedo Athanasios Nenes Department of Chemical Engineering,Mail Code 210-41, California Institute of Technology, Pasadena, California, 91125, USA Steven Ghan Pacific Northwest National Laboratory, Richland, Washington, 99352, USA Hayder Abdul-Razzak Texas A&M University-Kingsville, Kingsville, Texas, 78363, USA Patrick Y.Chuang NCAR, Boulder, Colorado, 80307, USA John H. Seinfeld Department of Chemical Engineering,Mail Code 210-41, California Institute of Technology, Pasadena, California, 91125, USA Abstract. Under certain conditions mass transfer limitations on the growth of cloud condensation nuclei (CCN) may have a significant impact on the number of droplets that can form in a cloud. The assumption that particles remain in equilibrium until activated may therefore not always be appropriate for aerosol populations existing in the atmosphere. This work identifies three mechanisms that lead to kinetic limitations, the effect of which on activated cloud droplet number and cloud albedo is assessed using a one-dimensional cloud parcel model with detailed microphysics for a variety of aerosol size distributions and updraft velocities. In assessing the effect of kinetic limitations, we have assumed as cloud droplets not only those that are strictly activated (as dictated by classical K¨ohler theory), but also unactivated drops large enough to have an impact on cloud optical properties. Aerosol number concentration is found to be the key parameter that controls the significance of kinetic effects. Simulations indicate that the equilibrium assumption leads to an overprediction of droplet number by less than 10% for marine aerosol; this overprediction can exceed 40% for urban type aerosol. Overall, the effect of kinetic limitations on cloud albedo can be considered important when equilibrium activation theory consistently overpredicts droplet number by more than 10%. The maximum change in cloud albedo as a result of kinetic limitations is less than 0.005 for cases such as marine aerosol; however albedo differences can exceed 0.1 under more polluted conditions. Kinetic limitations are thus not expected to be climatically significant on a global scale, but can regionally have a large impact on cloud albedo.

Upload: others

Post on 03-May-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Kinetic Limitations on Cloud Droplet Formation and Impact

1

Appeared in Tellus B , 53 , 133–149, 2001.

Kinetic Limitations on Cloud Droplet Formation andImpact on Cloud Albedo

Athanasios NenesDepartment of Chemical Engineering,Mail Code 210-41, California Institute of Technology,Pasadena, California, 91125, USA

Steven Ghan

Pacific Northwest National Laboratory, Richland, Washington, 99352, USA

Hayder Abdul-Razzak

Texas A&M University-Kingsville, Kingsville, Texas, 78363, USA

Patrick Y.Chuang

NCAR, Boulder, Colorado, 80307, USA

John H. SeinfeldDepartment of Chemical Engineering,Mail Code 210-41, California Institute of Technology,Pasadena, California, 91125, USA

Abstract. Under certain conditions mass transfer limitations on the growthof cloud condensation nuclei (CCN) may have a significant impact on thenumber of droplets that can form in a cloud. The assumption that particlesremain in equilibrium until activated may therefore not always be appropriatefor aerosol populations existing in the atmosphere. This work identifies threemechanisms that lead to kinetic limitations, the effect of which on activatedcloud droplet number and cloud albedo is assessed using a one-dimensionalcloud parcel model with detailed microphysics for a variety of aerosolsize distributions and updraft velocities. In assessing the effect of kineticlimitations, we have assumed as cloud droplets not only those that are strictlyactivated (as dictated by classical Kohler theory), but also unactivated dropslarge enough to have an impact on cloud optical properties. Aerosol numberconcentration is found to be the key parameter that controls the significanceof kinetic effects. Simulations indicate that the equilibrium assumption leadsto an overprediction of droplet number by less than 10% for marine aerosol;this overprediction can exceed 40% for urban type aerosol. Overall, the effectof kinetic limitations on cloud albedo can be considered important whenequilibrium activation theory consistently overpredicts droplet number bymore than 10%. The maximum change in cloud albedo as a result of kineticlimitations is less than 0.005 for cases such as marine aerosol; however albedodifferences can exceed 0.1 under more polluted conditions. Kinetic limitationsare thus not expected to be climatically significant on a global scale, but canregionally have a large impact on cloud albedo.

Page 2: Kinetic Limitations on Cloud Droplet Formation and Impact

1. Introduction

Much of the uncertainty associated with quantify-ing the indirect climatic effect of aerosols originates inthe complex relationship between aerosols and clouddroplets. Approximate analytical expressions that pre-dict those aerosol particles that activate to form clouddroplets have been proposed for implementation in gen-eral circulation models (GCMs) (Ghan et al. [1993]; Liuand Wang [1996]; Abdul-Razzak et al. [1998]). These pa-rameterizations generally rely on the assumption thatparticles are at equilibrium with the ambient (super-saturated) water vapor concentration until activatedas cloud condensation nuclei (CCN). The number ofdroplets formed in a cloud can therefore be estimatedfrom the number of CCN active at the maximum su-persaturation in the cloud updraft. The problem ofdroplet nucleation parameterization is then reduced tothe problem of determining the maximum supersatura-tion in the cloud parcel.

The assumption that all particles respond instanta-neously to any changes in supersaturation leads to aproblem: the amount of water absorbed by the largestaerosol particles when they activate can be larger thanthe amount of water vapor available in the cloud par-cel. This problem, known for a long time, is not se-rious when predicting the number of droplets, becausealthough not activated, these particles have an equilib-rium saturation ratio very close to unity, as activatedparticles do. As a consequence, their growth can be pa-rameterized as if they were activated. In addition, thenumber of large particles that give rise to this problemare usually negligible compared to the concentration ofsmaller particles of the distribution, so the errors incloud droplet number overall are expected to be small.

The assumption of equilibrium, however, can lead toa discrepancy in droplet number as a result of masstransfer limitations. Chuang et al. [1997] have shownthat under certain circumstances growth kinetics mayretard the growth of CCN sufficiently to limit the num-ber of activated droplets formed. By comparing thetime scale for particle growth at equilibrium with thatfor actual condensational growth, Chuang et al. [1997]conclude that particles with critical supersaturation lessthan a threshold value do not have time to grow largerthan their critical size, and thus do not activate. Thissuggests that equilibrium models that diagnose dropletformation from maximum supersaturation may system-atically overestimate the number of activated dropletsformed. Such a systematic bias could have implica-tions for estimates of indirect aerosol radiative forcing

of climate. It is difficult, however, to draw firm conclu-sions simply from a comparison of timescales becausegrowth kinetics depend on the full time history of su-persaturation and particle growth. Furthermore, thetimescales controlling the growth of the droplets changedrastically as the populations grows. A more conclusivemethod of evaluating the importance of growth kineticsis obtained by explicitly simulating the kinetic growthprocess and then comparing the simulated number ofdroplets formed with that predicted by equilibrium the-ory for the same CCN concentration and maximum su-persaturation.

Furthermore, it is incorrect to consider cloud dropsas only those that are strictly activated (as definedby Kohler theory); the presence of large unactivateddroplets cannot be neglected. This issue becomes evenmore important when slightly soluble substances arepresent in the aerosol (Laaksonen et al. [1998]). Thecondensational growth of an aerosol population pre-scribes that particles with very large dry diameters, al-though not activated, lead to roughly the same size ofactivated droplets and thus belong to the cloud dropletpopulation. With this in mind, one can define as CCNall particles that produce droplets that are larger thanthe smallest particle that is strictly activated. Thiscloud droplet definition differs slightly from that givenby Pandis et al. [1984]; we do not consider as droplets allthe particles that exceed their critical diameter or havea critical supersaturation lower than ambient supersat-uration, but only those comparable in size to those thatare strictly activated.

Based on the previous discussion, the inertial mecha-nism of Chuang et al. [1997] is believed not to contributeto any bias in the predicted droplet number when theequilibrium assumption is invoked. However, there areother kinetic limitation mechanisms that produce parti-cles that are much smaller than activated drops; in thissense, these mechanisms act to decrease the number ofcloud droplets from that predicted strictly on the basisof equilibrium activation. This study will focus mainlyon these mechanisms.

In the sections that follow, the mechanisms that leadto kinetic limitations in cloud droplet formation are pre-sented. The models used for evaluating kinetic effectsare then described, together with the relevant criteriaused for assessing the potential climatic importance ofthese effects. Finally, simulations are presented, for avariety of updraft conditions and aerosol types, fromwhich conclusions can be derived regarding the effect ofkinetic limitations on cloud droplet number concentra-tion and albedo.

2

Page 3: Kinetic Limitations on Cloud Droplet Formation and Impact

3

Figure 1. Illustration of the kinetic limitation mecha-nisms.

2. Kinetic Limitation Mechanisms

There are three kinetic limitation processes that caninhibit the formation of cloud droplets. These mecha-nisms will be explained with the help of Figure 1, whichillustrates typical cloud parcel and droplet equilibriumsupersaturation profiles (S and Seq, respectively) as afunction of cloud depth, as predicted by adiabatic par-cel theory. Each aerosol equilibrium curve correspondsto particles containing a different amount of solute. Theequilibrium curves vary with respect to cloud depth as aconsequence of droplets changing size as they traversethrough the cloud column. According to equilibriumactivation theory, any particle with a critical supersat-uration, Sc, less than the maximum supersaturation,Smax, encountered in the parcel will activate. However,the time which particles are exposed to a supersatu-ration level is a crucial parameter; that time must besufficiently long to allow the particle to reach its criti-cal diameter. The droplet can be considered activatedonly when the wet diameter exceeds its critical value.Furthermore, to ensure constant growth of the droplet,the ambient supersaturation should be high enough foractivated droplets to continuously grow throughout theduration of cloud formation.

The yellow curve of Figure 1 represents an aerosolparticle that activates and remains so throughout theentire time of cloud formation. The critical supersatu-ration of the particle, Sc3, is less than Smax; the time

needed for activation is also less than the time duringwhich S ≥ Sc3. When this particle activates, its equi-librium curve will be at a maximum (with Seq = Sc);subsequently, the particle is activated and Seq drops.As can be seen, the particle Seq is always less than theparcel supersaturation, so the driving force for growth,S − Seq, is always positive. This guarantees that theparticle will remain activated throughout the cloud.

The same cannot be said for all the aerosol typesdepicted in Figure 1. The first mechanism that limitsthe formation of activated droplets is the inertial mech-anism described by Chuang et al. [1997]. This mecha-nism is illustrated in Figure 1 for the particles with acritical supersaturation Sc4 (blue curve). These parti-cles have a large dry diameter and a very low criticalsupersaturation. The timescale of cloud formation isnot sufficient for these particles to reach their criticaldiameter. Nonetheless, the driving force for growth isalways positive, and these particles continuously grow,attaining a wet diameter similar (and actually larger) tothose of the activated droplets. Thus, even though theseparticles do not activate, they cannot be distinguishedfrom activated droplets, and so should be treated assuch.

The red curve corresponds to a particle with a rela-tively high critical supersaturation, Sc1. The time dur-ing which S > Sc1 is not sufficient for activation. Asa result, the particle initially grows, but subsequentlyevaporates to become an interstitial aerosol particle.This kinetic effect is the second of the three mecha-nisms identified, and is termed the “evaporation mech-anism”. Although this is an inertial mechanism (in thesense that small particles do not respond fast enough tochanges in ambient supersaturation), a different nameis assigned because the particles affected behave muchdifferently than those subject to the inertial mechanismof Chuang et al. [1997].

Finally, some particles can initially activate but be-come interstitial aerosol through the third mechanism,the so-called “deactivation mechanism”. This is illus-trated in Figure 1 for particles of critical supersatura-tion Sc2 (green curve). These particles are exposed toa supersaturation that exceeds their critical values suf-ficiently long to activate and do so initially. However,after a while, the parcel supersaturation drops belowthe droplet Seq. When this happens, the growth drivingforce, S−Seq, becomes negative, and these droplets be-gins evaporating. The rate of evaporation can be quitefast, and the droplet may deactivate and become partof the interstitial aerosol, thus decreasing the number ofcloud droplets. This mechanism is not a result of mass

Page 4: Kinetic Limitations on Cloud Droplet Formation and Impact

4

transfer kinetic limitation, but rather a dynamic effectarising from the limited available water vapor; watertransfers from activated drops to other sizes that canstill grow.

The deactivation and evaporation mechanisms ren-der the affected aerosol much smaller than the activateddroplet size, leading to considerably smaller contribu-tion to cloud optical properties. On the other hand,whereas the inertial mechanism prevents droplets fromactivating, it does not produce droplets that are dif-ferentiated from other activated droplets as the othertwo mechanisms do. When using the equilibrium as-sumption, it is expected that the inertial mechanismdoes not lead to any bias in predicted droplet number.The same cannot be said for the evaporation and deac-tivation mechanisms, which tend to decrease the num-ber of cloud droplets that are formed. Therefore usingthe equilibrium assumption in the presence of these twomechanisms will tend to overestimate the droplet num-ber.

In summary, if kinetic limitations affect mainly largerparticles (through the inertial mechanism), the equilib-rium assumption should not induce a large error in pre-dicted droplet number. However, if kinetic effects ap-ply mainly to the smaller particles of a distribution, notonly can the equilibrium assumption yield large error inpredicted droplet number, but the droplet number willbe sensitive to fluctuations in parcel supersaturation.Any factor that can influence supersaturation history(such as mode radius, number concentration, and up-draft velocity) will, in turn, affect all the relevant timescales and thus the extent and type of kinetic limita-tions.

3. Measures of Kinetic Limitations

Assessing the effect of kinetic limitations on clouddroplet formation requires first the calculation of twoquantities, Neq and Nkn, the number concentrations ofdroplets based on equilibrium and kinetic approaches,respectively. Neq is equal to the concentration of parti-cles with critical supersaturation, Sc, less than or equalto the maximum supersaturation, Smax, achieved in theascending air parcel (as calculated by the parcel model).Neq is based on the assumption that the particles thatcan activate do so instantaneously, and is the upperlimit to the number of droplets that can be formed.Nkn is the actual droplet concentration, and is equal tothe number of particles that are larger than the acti-vated particle with the smallest dry diameter (i.e. thathas a wet radius larger than its critical value). This

number contains droplets that have a critical supersat-uration less than the parcel maximum supersaturationbut which are not larger than their critical size. Criticalparameters are calculated from classical Kohler theory(e.g., Seinfeld and Pandis [1998]).

The importance of kinetic growth limitations ondroplet formation will be measured in terms of both theactivated droplet number and the cloud albedo. Basedon the variation of Nkn and Neq with cloud depth, z,one can define the total droplet ratio at any height z,α (z),

α (z) = Nkn (z)/Neq (z) (1)

which expresses the ratio of actual droplet number tothe maximum droplets attainable at a certain distanceabove cloud base. The droplet ratio includes bothstrictly activated and unactivated droplets. It is alsouseful, when assessing kinetic effects, to examine theportion of the droplet population that is strictly acti-vated. For this purpose, the unactivation ratio, φ (z),defined as the fraction of droplets that are not activated,is used:

φ (z) = Nunact (z)/Nkn (z) (2)

where Nunact (z) is the number of unactivated dropletsin the distribution. For example, a φ (z) = 0.2 meansthat 20% of the droplets are smaller than their activa-tion diameter and thus are not strictly activated.

Profiles of α (z) and φ (z) can provide insight regard-ing the kinetic limitation mechanisms present. Figure 2presents a qualitative sketch of the three types of α (z)profiles seen in the simulations. Each type represents acase where different kinetic limitation mechanisms areactive. The “type 1” profile is observed when the in-ertial mechanism is the only type of kinetic limitationactive. In this case, φ (z) initially attains large val-ues that decreases further up in the cloud column (notshown), and α (z) approaches unity with increasing z.If the evaporation mechanism is active, α (z) initiallyincreases and approaches an asymptotic value less thanunity; the fraction that never activates are the smallparticles that evaporate. This is the “type 2” profileof Figure 2. Finally, if the deactivation mechanism ispresent, then the droplet ratio would initially increase,reach a maximum, and then begin to decrease as parti-cles evaporate and deactivate.

It is difficult to determine a priori when a discrep-ancy in cloud droplet number is important. Placedin the context of the effect on albedo, this issue be-comes more straightforward; the discrepancy in droplet

Page 5: Kinetic Limitations on Cloud Droplet Formation and Impact

5

Figure 2. Illustration of the three types of dropletratio profiles seen in the simulations.

number can be considered significant when the albedois biased by an amount comparable to the change in-duced by anthropogenic effects. Furthermore, GCMscurrently implement a cloud drop number calculationfor determining cloud albedo, so it is directly relevantto examine the potential error from the equilibrium ac-tivation assumption. In calculating cloud albedo, thecloud liquid water content and effective radius of thedroplet distribution are used, with the assumption thatthe droplet distribution is narrow. Furthermore, the ef-fect of interstitial aerosol on the liquid water contentand optical properties are neglected.

4. Cloud Parcel and Albedo Models

A cloud parcel model is the simplest tool that canbe used to simulate the evolution of droplet distri-butions throughout a non-precipitating cloud column.These models predict a number concentration profilethat starts from zero at cloud base and reaches anasymptotic value further up. In reality, droplet num-ber and size are also affected by turbulent mixing anddowndrafts, which cannot be correctly accounted for ina single parcel model. Near cloud base, where kineticeffects are strongest, droplets from upper levels tendto dry out and do not participate in the droplet dis-tribution. Therefore, one would still expect a dropletnumber concentration that is zero at cloud base andquickly reaches an asymptotic value. Such distributionshave been measured and predicted by more compre-

hensive models (Considine and Curry [1998]), so thisvariation in droplet number with cloud height must beconsidered when calculating optical properties. Finally,existing theoretical aerosol-cloud parameterizations inGCMs are based on adiabatic cloud parcel model equa-tions, and so estimating cloud albedo sensitivity to ki-netic effects is appropriately carried out using adiabaticparcel model calculations. In order to assess the dif-ferences arising from the kinetic and thermodynamicassumptions, a droplet growth model has been incor-porated within the framework of an adiabatic parcelmodel.

4.1. Cloud Parcel Model

The adiabatic cloud parcel model is based upon theparcel model described by Pruppacher and Klett [1997]and Seinfeld and Pandis [1998]. Conservation of heatand moisture for a rising air parcel can be expressed as

dT

dt= −gV

cp− L

cp

dwv

dt(3)

dwv

dt= −dwc

dt(4)

where T is the temperature of the air, V is the updraftvelocity, and wv and wc are the mixing ratios of watervapor, and liquid water in the parcel, respectively.

In (4), the condensation rate for a population of wa-ter droplets consisting of Ni droplets of radius ri, i =1...n can be expressed as

dwc

dt=

4πρw

ρa

n∑

i=1

Nir2i

dri

dt(5)

where the particle growth rate is determined from,

dr

dt=

G

r(S − Seq) (6)

with G given by

G =1

ρwRTp∗vD′

vMw+ Lρw

k′aT(LMwRT −1)

(7)

The supersaturation S is given by wv/w∗v − 1. Byintegration of the supersaturation balance equation,

dS

dt=

1w∗v

[dwv

dt− (S + 1)

(∂w∗v∂T

dT

dt− ∂w∗v

∂paρagV

)]

(8)

Page 6: Kinetic Limitations on Cloud Droplet Formation and Impact

6

and the equilibrium supersaturation Seq is given byKohler theory,

Seq = exp

2Mwσw

RTρwri− 3nsMw

4πρw

(r3i − r3

dry,i

)− 1 (9)

Here we have used the hydrostatic relation to relatechanges in atmospheric pressure to vertical velocity in(8). D′

v in (9) is the diffusivity of water vapor in air,modified for non-continuum effects,

D′v =

Dv

1 + Dv

acr

√2πMw

RT

(10)

where ac = 1.0 is the condensation coefficient. k′a in(7) is the thermal conductivity of air modified for non-continuum effects,

k′a =ka

1 + ka

aT rρcp

√2πMa

RT

(11)

where aT = 0.96 is the thermal accommodation coeffi-cient, and ns in (9) is the number of moles of solute perparticle,

ns =4ρpενπd3

i

3Ms(12)

where di is the dry particle diameter, and

ρp = [(1− ε) /ρu + ε/ρs]−1 (13)

is the mean particle density. Surface tension σw is ex-pressed (in J m−2) as a function of the parcel temper-ature T (Seinfeld and Pandis [1998]), σw = 0.0761 −1.55×10−4 (T − 273). Other symbols are defined in theAppendix. Equations (3)-(8) constitute a closed systemof ordinary differential equations that are solved numer-ically using the LSODE solver of Hindmarsh [1983].

4.2. Cloud Albedo

Cloud albedo, Rc, is calculated based on the two-stream approximation of a non-absorbing, horizontallyhomogeneous cloud (Lacis and Hansen [1974]),

Rc =τ

7.7 + τ(14)

where τ is the cloud optical depth,

τ =

H∫

0

3ρawL (z)2ρwreff (z)

dz (15)

where wL (z) is the liquid water mixing ratio profilealong the cloud column, calculated from the parcelmodel simulations (after transforming the Lagrangiansolution into Eulerian form by setting wL (z) = wL (t)at t = z/V ).ρw is the water density, ρa is the air densityand reff (z) is the cloud droplet distribution effectiveradius,

reff =

∞∫0

r3n(r)dr

∞∫0

r2n(r)dr

(16)

where n(r) is the droplet size distribution. These ex-pressions yield values for cloud albedo that are of rea-sonable accuracy for relatively thick clouds composed ofnarrow distributions of large droplets (Hatzianastassiouet al. [1997]).

Assuming that the interstitial aerosol has a negligibleamount of liquid water, and that the droplet populationis effectively monodisperse, the expression for reff (z)is given by

reff (z) =3wL (z)

4πNi (z) ρw(17)

where Ni (z) can be either Nkn (z) or Neq (z). In thefirst case, the albedo calculated will be the “kineticalbedo”, while the second will be the “thermodynamicalbedo”. The thermodynamic albedo tends to be higherthan the kinetic albedo. This is because for the sameamount of cloud liquid water, the number of dropletspredicted is larger in the thermodynamic case, hencethe effective radius according to (17) would be smallerthan in the kinetic case.

The quantity used for assessing the importance of ki-netic effects, in terms of cloud albedo, is the differencebetween thermodynamic and kinetic albedo. Becausethe thermodynamic albedo is larger than the kinetic(as explained before), this difference will be positive.Furthermore, since this difference depends on the clouddepth, we select the cloud depth for which this differ-ence is maximum.

5. Simulation Parameters

The importance of kinetic limitations depends oncloud thickness, updraft velocity, and aerosol character-

Page 7: Kinetic Limitations on Cloud Droplet Formation and Impact

7

istics. To explore the dependence on these parameters,we shall consider a range of values spanning observa-tions of boundary layer clouds. This study does notexamine all possible phenomena that influence aerosolactivation behavior, such as changes in surface tension,and the presence of slightly soluble substances. For ex-ample, a decrease in surface tension (from surfactantspecies in the aerosol) should enhance kinetic effects,because a) the critical radius for activation becomeslarger, so particles need more time to activate, and b) alarger fraction of the aerosol population can activate, somore particles compete for cloud water. Solution non-idealities are not considered; they do not have a sig-nificant impact since droplets dilute considerably undersupersaturated conditions (Young and Warren [1992]).Finally, the effect of uncertainty in the accommodationcoefficient is not examined.

5.1. Key Parameters

Cloud thickness influences the transit time of airparcels rising through a cloud, and hence the time avail-able for particle growth. Boundary layer clouds are typ-ically 100-500 m thick, with most in the range 200-400m (Nichols [1984]; Duynkerke et al. [1995]; Frisch et al.[1995a]; White et al. [1995]). To explore the dependenceof kinetic effects on cloud thickness, we consider valuesranging from 10 to 1000 m.

Updraft velocity influences both the transit time andthe maximum supersaturation in a cloud updraft. Themaximum supersaturation achieved is lower in weakerupdrafts, so only particles with relatively low criticalsupersaturations can be activated. Observed updraftvelocities in boundary layer clouds vary widely, but val-ues derived from measured vertical velocity variance aretypically 30-50 cm s−1 (Nichols [1984]; Duynkerke et al.[1995]; Frisch et al. [1995b]). We explore the depen-dence on updraft by considering updraft velocities of10, 30, 100, and 300 cm s−1.

5.2. Aerosol characteristics

The dependence of kinetic effects on aerosol numberconcentration and mean radius will be explored by con-sidering a variety of aerosol size distributions. We con-sider idealized size distributions in which the numberconcentration and mean radius are prescribed in orderto clearly sort out the physics involved. We then con-sider size distributions that more closely approximateambient distributions.

Size distributions are of the single or multiple log-normal form,

dn (r)d ln r

=nm∑

i=1

Ni√2π ln σi

exp[− ln2 (r/rg,i)

2 ln2 σi

](18)

where Ni is the aerosol number concentration, rg,i thenumber mode radius, σi is the geometric standard devi-ation for mode i, and nm is the number of modes in thedistribution. For single modes we consider a single valuefor σ = 2, but a wide range in number concentration(100-3000 cm−3) and mode radius (0.02-0.1 µm). Therange in mode radius and number concentration is ap-propriate for accumulation mode particles, which com-prise most CCN. For multiple modes we have selectedfour of Whitby [1978] trimodal representations, namelythe marine, clean continental, average background, andurban aerosol. The parameters of these four distribu-tions are listed in Table 1. The size distributions refer todry size, while the chemical composition of the aerosolis assumed pure ammonium sulfate.

In all kinetic simulations, particles are assumed ini-tially to be in equilibrium with a relative humidity of98%. For the idealized size distributions, we consider200 size bins spaced equally in log radius. Using asize range between about Dp,g/10σ and 10σDp,g cov-ers total particle number to within 10−7%. The sim-ulations exhibit little sensitivity with respect to initialrelative humidity concentrations and denser discretiza-tion schemes.

6. Effect of Kinetic Limitations onCloud Droplet Number

We first explore the parametric dependence of ki-netic limitations using the single log-normal size dis-tributions. We then consider various multimode log-normal distributions that resemble ambient aerosol.

6.1. Single log-normal size distributions

From the simulations we calculate the α (z) and φ (z)profiles. The results of these simulations are summa-rized in Table 2, which gives the characteristics of theseratios for all the mode radii, number concentrations,and updrafts examined. As expected, the droplet ratio(for constant mode radius and number concentration)in most cases approaches an asymptotic value for largecloud depths. Kinetic effects become more apparent asupdraft velocity decreases and aerosol concentrationsincrease. For a mode radius of 0.03 µm, the dropletnumber is reduced by 40% for an updraft velocity of 10cm s−1 and an aerosol number concentration of 3000

Page 8: Kinetic Limitations on Cloud Droplet Formation and Impact

8

Table 1. Aerosol distribution parameters (rg,i in µm, Ni in cm−3) [Whitby , 1978]

Aerosol type Nuclei Mode Accumulation Mode Coarse Moderg,1 σ1 N1 rg,2 σ2 N2 rg,3 σ3 N3

Marine 0.005 1.6 340 0.035 2.0 60 0.31 2.7 3.1Clean Continental 0.008 1.6 1000 0.034 2.1 800 0.46 2.2 0.72Average Background 0.008 1.7 6400 0.038 2.0 2300 0.51 2.16 3.2Urban 0.007 1.8 106000 0.027 2.16 32000 0.43 2.21 5.4

cm−3. In this particular example, the unactivation ra-tio φ (z) is close to zero. This means that most (morethan 95%) of the cloud drops are activated. In this par-ticular case, the droplet ratio profile is “type 3”, so bothevaporation and deactivation mechanisms are present.

The dependence of kinetic effects on Ni is furtherpronounced as mode radius increases. For example,the droplet ratio for an aerosol number concentrationof 3000 cm−3 and a mode radius of 0.1 µm can be closeto zero for a large portion of the cloud, because thetime required for growth is very large for weak updrafts(this is also seen in φ (z), which ranges between 1.0and 0.083). In this particular case, the α (z) profile is“type 2”, so the evaporation mechanism is present. Thefact that α (z) is maximum at 700 m above cloud baseindicates that kinetic effects are very strong. This isnot surprising, given the mode radius and concentra-tion of particles. As updraft velocity increases, super-saturation, being the driving force for particle growth,also increases and activates particles lower in the cloud.The dependence on cloud thickness is also stronger fora mode radius of 0.1 µm than for 0.03 µm because a)maximum supersaturation is achieved further up fromcloud base, and, b) the larger particles respond moreslowly to variations in supersaturation. The deactiva-tion mechanism is responsible for decreasing the dropletratio for low updrafts and high number concentrations(> 300 cm−3). Another striking feature, as evidencedby φ (z), is that the inertial mechanism is negligible forparticles of smaller modal diameter; in the larger sizedistributions, roughly half of the particles are not acti-vated during the first 50-100 m of the cloud. Despitethe difference in modal size, the fraction of small par-ticles that fail to activate is roughly the same at largecloud depths.

The simulations indicate that a large mode radiustends to accentuate kinetic effects. By examining α (z)and φ (z) as a function of mode radius for a number

concentration of 100 cm−3, we can see that evapora-tion and deactivation effects are negligible for all moderadii; most droplet ratio profiles are of type 1. Inertialeffects (as shown by the value of φ (z) at 20 m) be-come increasingly important as mode radius increases.On the other hand, α (z) never drops below 0.95. Thisis an important point: for low concentration of parti-cles, kinetic effects that influence droplet number arenegligible regardless of particle diameter. The inertialmechanism is the only type of limitation experienced bythese particles.

Kinetic effects are more prominent when the numberconcentration increases to 1000 cm−3. When the con-centration of particles is this high, α (z) profile typesindicate that the deactivation and evaporation mech-anisms are much more prominent than for a particleconcentration of 100 cm−3. This suggests that thesekinetic limitation mechanisms are present primarily athigh number concentrations.

6.2. Trimodal log-normal size distributions

From the previous section it is clear that kinetic lim-itations on droplet formation are important for highnumber concentrations, and large mean particle size canfurther enhance these effects. Although one can imaginecombinations of number concentration and mean parti-cle size that would yield significant kinetic limitations,such combinations may not be realistic. High parti-cle concentrations are normally associated with smallparticle sizes, and vice versa. An assessment of the im-portance of kinetic limitations on ambient droplet for-mation should therefore consider size distributions thatrepresent ambient aerosol. In doing so, we examine thetrimodal log-normal fits to measured size distributionsfor a variety of aerosol types in Table 1.

Although the number concentration for the nucleimode can be quite large, the particles are so smallthat the supersaturation necessary for activating them

Page 9: Kinetic Limitations on Cloud Droplet Formation and Impact

9

Table 2. Characteristics of α(z) and φ(z) profiles for single lognormal aerosol distribution runs. αmax is themaximum value of α(z), encountered at zαmax above cloud base.

rg,i (µm) Ni (cm−3) V (m s−1) Profile type αmax zαmax (m) α (1000m) φ (20m) φ (1000m)

0.03 100 0.1 1 1.000 10 1.000 0.002 0.0000.03 100 0.3 1 1.000 20 1.000 0.006 0.0420.03 100 1.0 1 1.000 20 1.000 0.024 0.0260.03 100 3.0 1 1.000 20 1.000 0.095 0.0120.03 300 0.1 2 0.931 10 0.931 0.005 0.0000.03 300 0.3 3 1.000 10 0.948 0.008 0.0000.03 300 1.0 1 1.000 20 1.000 0.028 0.0000.03 300 3.0 1 1.000 20 1.000 0.095 0.0200.03 1000 0.1 3 0.918 10 0.768 0.014 0.0000.03 1000 0.3 3 1.000 10 0.936 0.019 0.0000.03 1000 1.0 2 0.956 10 0.956 0.037 0.0000.03 1000 3.0 1 1.000 20 1.000 0.103 0.0290.03 3000 0.1 3 0.815 10 0.589 0.052 0.0000.03 3000 0.3 3 0.922 10 0.713 0.050 0.0000.03 3000 1.0 3 0.944 10 0.889 0.066 0.0000.03 3000 3.0 3 1.000 20 0.961 0.141 0.0010.05 100 0.1 1 1.000 10 1.000 0.019 0.0420.05 100 0.3 2 0.974 10 0.974 0.032 0.0000.05 100 1.0 1 1.000 20 1.000 0.092 0.0130.05 100 3.0 1 1.000 20 1.000 0.231 0.0040.05 1000 0.1 3 0.857 10 0.668 0.159 0.0010.05 1000 0.3 3 0.948 10 0.847 0.154 0.0010.05 1000 1.0 2 0.970 20 0.970 0.177 0.0020.05 1000 3.0 1 0.985 20 0.985 0.323 0.0030.10 100 0.1 3 0.974 10 0.948 0.206 0.0050.10 100 0.3 1 1.000 10 1.000 0.244 0.0060.10 100 1.0 1 1.000 20 1.000 0.386 0.0080.10 100 3.0 1 1.000 30 1.000 0.586 0.0090.10 300 0.1 3 0.908 20 0.774 1.000 0.0110.10 300 0.3 3 0.954 20 0.929 0.414 0.0140.10 300 1.0 1 1.000 20 1.000 0.422 0.0240.10 300 3.0 1 0.998 30 0.998 0.671 0.0180.10 1000 0.1 3 0.925 200 0.785 1.000 0.0340.10 1000 0.3 3 0.950 70 0.804 1.000 0.0330.10 1000 1.0 3 0.979 20 0.933 1.000 0.0360.10 1000 3.0 1 1.000 20 0.994 0.755 0.0400.10 3000 0.1 2 0.902 700 0.902 1.000 0.1330.10 3000 0.3 3 0.925 500 0.853 1.000 0.1020.10 3000 1.0 3 0.952 200 0.857 1.000 0.0860.10 3000 3.0 3 0.979 50 0.933 1.000 0.083

Page 10: Kinetic Limitations on Cloud Droplet Formation and Impact

10

1 0 10 0 1 0 0 0C lo u d D ep th (m )

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

Rat

io

1 0 1 0 0 1 0 0 0C lo ud D ep th (m )

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

Rat

io

1 0 1 0 0 1 0 0 0C lo u d D ep th (m )

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

Rat

io

1 0 1 0 0 10 0 0C lo ud D ep th (m )

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

Rat

io

B ack g ro u n d C lea n C o n tin en ta l

M arin e

U rb a n

U p d ra ft V e lo c ity (m s -1 )

0 .1

0 .3

1 .0

3 .0F ig u re 3

Figure 3. Droplet ratio as a function of updraft ve-locity and cloud thickness for the background, cleancontinental, marine, and urban distributions of Table1. Solid lines represent the droplet ratio α (z), whiledashed lines represent the unactivation ratio φ (z).

is never encountered within a cloud. Furthermore, themean particle size of the coarse mode is large enoughfor significant kinetic effects, but the number concentra-tions are usually too small to have a significant impacton droplet number. Thus, the distribution characteris-tics of the accumulation mode is expected to determinethe kinetic effects of the aerosol types in Table 1.

Figure 3 shows α (z) and φ (z) as a function of up-draft velocity and cloud thickness for the aerosol typesshown in Table 1. Simulations reveal that kinetic ef-fects for the marine aerosol are negligible regardless ofupdraft velocity and cloud thickness. This is not sur-prising, given that the number concentration of the ac-cumulation mode, 60 cm−3, is too low for kinetic effectsto be important.

The accumulation mode number concentration for

the clean continental aerosol is much higher, 800 cm−3,so kinetic effects become evident at weak updraft veloc-ities: α (z) is always below 0.85 for V =10 cm s−1, and0.95 for V =30 cm s−1. Results for the clean continentalaerosol are comparable to those illustrated in Table 2for a number mode radius of 0.03 µm and a number con-centration of 1000 cm−3, but with kinetic effects beingsomewhat weaker because of the lower number concen-tration. The ratio φ (z), however, is low and the dropletratio approaches an asymptotic value, so the aerosol issubject to the evaporation mechanism, as some of theparticles fail to activate in the initial stages of cloudformation.

Kinetic limitations on droplet formation are quite im-portant for the average background aerosol, which hasan accumulation mode number concentration of 2300cm−3. The droplet ratio is a little over 0.8 for a 30cm s−1 updraft, which is consistent with the single log-normal size distribution with the same number concen-tration. Although there are strong inertial effects inthe coarse mode, its contribution to the total dropletnumber concentration is rather small. Because of this,φ (z) is low, less than 0.1 for cloud depths larger than20 m, indicating that most of the droplets are strictlyactivated. For weak updrafts, because of the high con-centration of accumulation mode particles, deactivationplays a significant role and reduces the droplet numberby 20%. Overall, the aerosol seems to be subject toboth deactivation and evaporation mechanisms, the de-gree to which depends on the updraft velocity.

With an accumulation mode number concentrationof 32,000 cm−3, one might expect substantial kinetic ef-fects for the urban aerosol, and indeed this is the case.In fact, kinetic effects are so strong, that α (z) is zerountil 70 m above cloud base. The activation ratio neverexceeds 0.8, even for the strongest updrafts. The latterindicates that many particles fail to activate as the su-persaturation decreases following the maximum super-saturation in all updraft cases examined. Inertial effectsare also significant; even for clouds of 200 m thickness,φ (z) is above 0.2. Finally, deactivation is also verystrong since the decrease in α (z) after the maximum isabout 20%.

7. Effect of Kinetic Limitations onCloud Albedo

For very thin clouds (on the order of 10 m), the op-tical depth is small, so the albedo difference itself issmall. As the cloud depth increases, so does opticaldepth. This initially leads to an increase in the albedo

Page 11: Kinetic Limitations on Cloud Droplet Formation and Impact

11

Figure 4. Maximum difference between thermody-namic and kinetic cloud albedo. The single mode log-normal distribution aerosol is used with (a) numbermode radius 0.03 µm, (b) number mode radius 0.1 µm,(c) number concentration of 100 cm−3, and (d) numberconcentration of 1000 cm−3.

difference; after a certain point, the optical depth isso large, that the albedo difference begins to decrease.The point of maximum difference between thermody-namic and kinetic albedo is used for evaluating kineticeffects. A climatically important kinetic effect wouldbe considered for a difference between kinetic and ther-modynamic albedo larger than about 0.005; this albedochange, if relevant globally, could yield an uncertaintyin radiative forcing comparable to that estimated for an-thropogenic indirect aerosol radiative forcing (Facchiniet al. [1999]). It should also be noted that the maximumalbedo difference in the simulations (both single modeand trimodal distributions) was encountered in the first300 m of the cloud. As expected, increasing the kineticeffects in droplet number tends to shift the locus of themaximum higher up in the cloud.

7.1. Single log-normal size distributions

Since the difference between thermodynamic and ki-netic cloud albedo depends on the difference in clouddroplet number, it is expected that the largest differ-ence would be seen for high number concentrations.

The discrepancy is also expected to be enhanced if thedistribution has a large modal radius. This can beseen in the two top panels of Figure 4, which showsthe maximum albedo difference, as a function of num-ber concentration, for two modal radii. For a numbermode radius of 0.03 µm (top left panel), there is largesensitivity of kinetic effects on albedo to number con-centration. For low concentrations (100 cm−3), kineticlimitations are not strong enough to yield a significantdifference in cloud albedo; this difference becomes verylarge however, close to 0.05, when the concentration isaround 3000 cm−3. For a number mode radius of 0.1µm (top right panel), these effects are enhanced con-siderably and the albedo difference approaches 0.7 forthe lowest updraft and highest number concentrations.The cases where albedo difference is below 0.005 seemto be the same for both mode radius cases, when thenumber concentration is below 300 cm−3.

The above statements are further supported by thetwo bottom panels of Figure 4. These show the max-imum albedo difference as a function of mode radius,for two number concentrations. Regardless of modal ra-dius, when the concentration is 100 cm−3 (bottom leftpanel) the albedo difference is insignificant. However,when the concentration becomes 1000 cm−3, kinetic ef-fects are always significant. When the albedo differenceis around 0.1, the effect of mode radius seems to besmall.

By comparing Figure 4 with Table 2, we observe aconsistent trend; the maximum albedo difference tendsto exceed the 0.005 value whenever the droplet ratioconsistently (that is, for all cloud depths) is below 0.9.This is an important point, since it provides a quan-titative measure for when the bias in droplet numberbecomes important.

7.2. Trimodal log-normal size distributions

Figure 5 shows the maximum difference in cloudalbedo between the thermodynamic and kinetic acti-vation models, for the four log-normal distributions ofTable 1. Albedo differences are shown for both the totalnumber concentration in Table 1 and twice these values.The urban aerosol distribution exhibits the largest dif-ference in albedo, which is expected, given that it is thedistribution with the largest kinetic limitation effectsin droplet number. Of the other aerosol classes, thelargest sensitivity is experienced by the clean continen-tal aerosol, where a doubling in concentration leads to afivefold increase in albedo difference (at the highest up-draft velocities). As in section 7.1, kinetic limitationsbecome important for cloud albedo whenever α (z) is

Page 12: Kinetic Limitations on Cloud Droplet Formation and Impact

12

Figure 5. Maximum difference between thermody-namic and kinetic cloud albedo. The aerosol distribu-tions in Table 1 are used with (a) given number concen-trations, and (d) doubled number concentrations.

consistently below 0.9.

8. Summary and Conclusions

There are three mechanisms that lead to kinetic limi-tations on cloud droplet formation. The first, identifiedby Chuang et al. [1997], and termed the inertial mech-anism, arises when particles with critical supersatura-tion less than a threshold value do not have time togrow larger than their critical size and thus do not acti-vate. In the second deactivation mechanism, activateddroplets evaporate to form interstitial aerosol when theparcel supersaturation drops below the droplet equi-librium saturation ratio. In the third mechanism, theevaporation mechanism, particles that could potentiallyactivate (because their critical supersaturation is lowenough) evaporate to form interstitial aerosol becausethe time they are exposed to high supersaturation is not

sufficient to reach their critical radius. Large particles,which are subject to the inertial mechanism, are alsolarge enough to be considered cloud droplets even whenthey are not strictly activated. Particles, however, thatare subject to the deactivation and evaporation mech-anisms become much smaller than activated drops andhence should not be considered cloud droplets. Thus,if deactivation and evaporation significantly reduce thedroplet number concentration, these processes shouldbe accounted for in parameterizations of the activationprocess. The inertial mechanism, although a type ofkinetic limitation, does not generate a discrepancy incloud droplet number. Most of the discrepancy is a re-sult of the smaller particles that fail to activate, throughthe two other mechanisms identified. This also impliesthat for cases which kinetic effects are significant, thedroplet number is quite sensitive to fluctuations in su-persaturation.

In terms of number of cloud droplets, the conditionsmost likely to yield considerable kinetic effects are thoseof high aerosol number concentrations. Weak updraftsand large mode radius tend to enhance kinetic effects.We have also investigated a variety of aerosol types todetermine those subject to important kinetic effects un-der a variety of cloud formation conditions. We havefound that urban aerosol is strongly affected by ki-netic growth limitations, but relatively pristine marineaerosols are not. We have found that for most aerosolsthe percentage errors from neglecting kinetic effects arelarger for weak updrafts than for strong updrafts, eventhough the residence time of air parcels in clouds isshorter for strong updrafts. However, absolute errorsin droplet number are often largest for moderate up-drafts because aerosol activation is inefficient for weakupdrafts, and kinetic limitations are smaller for strongupdrafts.

The conditions most likely to yield considerable dif-ferences in cloud albedo are similar to those for dropletnumber, i.e., weak updrafts, large mode radius, andhigh aerosol concentrations. Simulations indicate thatthe maximum albedo difference is less than 0.005 formarine aerosol (where kinetic limitations in cloud dropletnumber are also not significant). Albedo differencescan, however, exceed 0.1 for urban aerosol.

By comparing the droplet ratio and albedo differenceplots, it can be seen that kinetic effects become impor-tant whenever the droplet ratio is consistently below0.9. Of the three kinetic limitation mechanisms, theevaporation and deactivation mechanisms can influencedroplet number concentrations. However, the deactiva-tion mechanism does not affect aerosol throughout the

Page 13: Kinetic Limitations on Cloud Droplet Formation and Impact

13

cloud and always appears together with the evapora-tion mechanism. So, it is believed that conditions thatare conducive towards the appearance of the evapora-tion mechanism can lead to substantial kinetic effects.As a consequence, kinetic limitations are not expectedoverall to be climatically significant, but can have a no-ticeable local impact.

Existing parameterizations of aerosol activation inGCMs do not account for kinetic limitations. Becauseclosed-form solutions of the activation process are notavailable, parameterizations can be fit to kinetic simu-lations. Future parameterizations should therefore bematched to kinetic simulations that extend beyond thepoint of maximum supersaturation, classify activatedparticles on the basis of their size rather than their crit-ical supersaturation, and also account for fluctuationsin cloud supersaturation.

Acknowledgments. This work was supported at PNNLby the NASA Earth Science Enterprise under contractNAS5-98072 and by the U.S. Department of Energy Atmo-spheric Radiation Measurement Program, which is part ofthe DOE Biological and Environmental Research Program.PNNL is operated for the DOE by Battelle Memorial Insti-tute under contract DE-AC06-76RLO 1830. This work assupported at the California Institute of Technology by theOffice of Naval Research.

Notation

ac condensation coefficientaT thermal accommodation coefficientcp specific heat of dry air at constant pressurerdry particle dry radiusDv water vapor diffusivity in airD′

v water vapor diffusivity in air, corrected for non-continuumeffects

g acceleration of gravityG particle growth parameter, as given by (7)H cloud thicknesska thermal conductivity of airk′a thermal conductivity of air, corrected for non-continuum

effectsL latent heat of condensation of waterMa molecular weight of dry airMs molecular weight of the soluteMw molecular weight of watern (r) number size distributionnm number of modes in idealized log-normal distributionsns number of moles of solute per particleN aerosol number concentrationNeq number concentration of particles with Sc smaller

than or equal to Smax

Nkn number concentration of particles with dry radius

larger than that of the smallest activated particleNunact number concentration of large unactivated parti-

cles (with dry radius larger than that of the smallest acti-vated particle, but with a wet radius less than the criticalsize)

pa air pressurep∗v saturation vapor pressure of waterr particle wet radiusrg,i number mode radius of log-normal size distributionreff size distribution effective radiusR universal gas constantRc cloud albedoS supersaturationSc critical supersaturation for activating according to Kohler

theorySeq supersaturation in equilibrium with particle sizeSmax maximum supersaturation in a cloud parcelt timeT temperature of air parcelV parcel updraft velocitywv water vapor mixing ratio in air parcelwc cloud liquid water mixing ratiow∗v p∗v/pa saturation mixing ratio of waterz cloud depth as measured from cloud base (or where

S=0%)α droplet ratio Nkn/Neq

ε mass fraction of soluble material in the dry particleν number of ions the solute dissociates into in solutionρa density of airρw density of waterρp mean density of particleρs density of soluble component of particleρu density of insoluble component of particleσ geometric standard deviation of the aerosol size distri-

butionσw water surface tensionτ optical depthφ unactivation ratio Nunact/Nkn

References

Abdul-Razzak, H., S. Ghan, and C. Rivera-Carpio, A pa-rameterization of aerosol activation. part i: Single aerosoltype, J. Geophys. Res., 103 , 6123–6132, 1998.

Chuang, P., R. Charlson, and J. Seinfeld, Kinetic limitationon droplet formation in clouds, Nature, 390 , 594–5962,1997.

Considine, G., and J. Curry, Effects of entrainment anddroplet sedimentation on the microphysical structure ofstratus and stratocumulus clouds, Q.J.R.Meteorol.Soc.,124 , 123–150, 1998.

Duynkerke, P., H. Zhang, and P. Jonker, Microphysicaland turbulent structure of nocturnal stratocumulus asobserved during astex, J. Aerosol Sci., 124 , 2763–2777,1995.

Facchini, M., M. Mircea, S. Fuzzi, and R. Charlson, Cloud

Page 14: Kinetic Limitations on Cloud Droplet Formation and Impact

14

albedo enhancement by surface-active organic solutes ingrowing droplets, Nature, 401 , 257–259, 1999.

Frisch, A., C. Fairall, and J. Snider, Measurement of stratuscloud and drizzle parameters in astex with a kα-banddoppler radar and a microwave receiver, J. Aerosol Sci.,52 , 2788–2799, 1995a.

Frisch, A., D. Lenschow, C. Fairall, W. Schubert, andJ.S.Gibson, Doppler radar measurements of turbulencein marine stratiform cloud during astex, J. Aerosol Sci.,52 , 2800–2808, 1995b.

Ghan, S., C. Chuang, and J. Penner, A parameterization ofcloud droplet nucleation. part i: Single aerosol species,Atmos. Res., 30 , 197–222, 1993.

Hatzianastassiou, N., W. Wobrock, and A. Flossman, Therole of droplet spectra for cloud radiative properties,Q.J.R.Meteorol.Soc., 123 , 2215–2230, 1997.

Hindmarsh, A., Odepack: A systemetized collection of odesolvers, in Scientific Computing , edited by R. S. et al.,pp. 55–64, North–Holland, New York, 1983.

Laaksonen, A., P. Korhonen, M. Kulmala, and R. Charl-son, Modification of the kohler equation to include solu-ble trace gases and slightly soluble substances, J. AerosolSci., 155 , 853–862, 1998.

Lacis, A. A., and J. E. Hansen, A parameterization for theabsorption of solar radiation in the earth’s atmosphere,J. Aerosol Sci., 31 , 118–133, 1974.

Liu, X., and M. Wang, A parameterization of the effi-ciency of nucleation scavenging of aerosol particles andsome related physicochemical factors, Atmos. Environ.,30 , 2235–2341, 1996.

Nichols, S., The dynamics of stratocumulus: aircraft ob-servations and comparisons with a mixed layer model,Q.J.R.Meteorol.Soc., 110 , 783–820, 1984.

Pandis, S., J. Seinfeld, and C. Pilinis, The smog-fog-smogcycle and acid deposition, J. Geophys. Res., 95 , 18,489–18,500, 1984.

Pruppacher, H., and J. Klett, Microphysics of Clouds andPrecipitation, 2nd ed., Kluwer, Dordrecht, 1997.

Seinfeld, J., and S. Pandis, Atmospheric chemistry andphysics: from air pollution to climate change, John Wi-ley, New York, 1998.

Whitby, K., The physical characteristics of sulfur aerosols,Atmos. Environ., 12 , 135–159, 1978.

White, A., C. Fairall, and J. Snider, Surface-based remotesensing of marine boundary-layer cloud properties, J.Aerosol Sci., 52 , 2827–2838, 1995.

Young, K., and A. Warren, A reexamination of the deriva-tion of the equilibrium supersaturation curve for solubleparticles, J. Aerosol Sci., 49 , 1138–1143, 1992.

A. Nenes and J. H. Seinfeld, Department of Chem-ical Engineering, California Institute of Technology,Pasadena CA, 91125. (e-mail: [email protected];[email protected])

S. Ghan, Pacific Northwest National Laboratory,Richland, Washington, 99352, USA

H. Abdul-Razzak, Texas A&M University-Kingsville,Kingsville, Texas, 78363, USA

P.Y. Chuang, NCAR, Boulder, Colorado, 80307, USA