+ All Categories
Home > Documents > Dynamical states of low temperature cirrus

Dynamical states of low temperature cirrus

Date post: 25-Feb-2022
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
15
Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/ doi:10.5194/acp-11-3757-2011 © Author(s) 2011. CC Attribution 3.0 License. Atmospheric Chemistry and Physics Dynamical states of low temperature cirrus D. Barahona 1,* and A. Nenes 1,2 1 School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, Atlanta, GA, USA 2 School of Earth and Atmospheric Sciences, Georgia Institute of Technology, Atlanta, GA, USA * now at: NASA Goddard Space Flight Center, Greenbelt, MD, USA Received: 28 November 2010 – Published in Atmos. Chem. Phys. Discuss.: 20 December 2010 Revised: 4 April 2011 – Accepted: 4 April 2011 – Published: 26 April 2011 Abstract. Low ice crystal concentration and sustained in- cloud supersaturation, commonly found in cloud observa- tions at low temperature, challenge our understanding of cir- rus formation. Heterogeneous freezing from effloresced am- monium sulfate, glassy aerosol, dust and black carbon are proposed to cause these phenomena; this requires low up- drafts for cirrus characteristics to agree with observations and is at odds with the gravity wave spectrum in the upper troposphere. Background temperature fluctuations however can establish a “dynamical equilibrium” between ice produc- tion and sedimentation loss (as opposed to ice crystal forma- tion during the first stages of cloud evolution and subsequent slow cloud decay) that explains low temperature cirrus prop- erties. This newly-discovered state is favored at low tempera- tures and does not require heterogeneous nucleation to occur (the presence of ice nuclei can however facilitate its onset). Our understanding of cirrus clouds and their role in anthro- pogenic climate change is reshaped, as the type of dynamical forcing will set these clouds in one of two “preferred” micro- physical regimes with very different susceptibility to aerosol. 1 Introduction Cirrus clouds are composed of ice crystals that form at high altitudes and temperatures typically below 235 K (Prup- pacher and Klett, 1997). They play a key role in climate by modulating the planetary radiative balance (Liou, 1986) and heat transport in the upper troposphere (Ramanathan and Collins, 1991). They strongly impact water vapor transport across the tropopause level (Jensen and Pfister, 2004) and play an important role in lower stratospheric chemistry (Pe- Correspondence to: A. Nenes ([email protected]) ter, 1997). Cirrus may be affected by aircraft emissions (Se- infeld, 1998) and long range transport of pollutants and play an important (but highly uncertain) role in anthropogenic cli- mate change. A key microphysical parameter required for understand- ing the climate impact of cirrus is their concentration of ice crystals, N c . At temperatures between 200 and 235 K cirrus ice crystals form primarily by homogenous freezing of super- cooled deliquesced aerosol (DeMott et al., 2003; Heymsfield and Sabin, 1989), which occurs if the saturation ratio with respect to ice, S , (i.e., the ratio of water vapor partial pres- sure to its equilibrium value over ice) reaches a characteris- tic threshold value, S hom (Koop et al., 2000). Heterogeneous freezing of water upon existing aerosol particles (termed “ice nuclei”, IN) can also occur (at S lower than S hom ) and con- tribute to ice crystal concentrations (DeMott et al., 2003; Froyd et al., 2009), especially in polluted and dust-rich re- gions (Barahona et al., 2010a; Haag et al., 2003). The level of water vapor supersaturation (i.e., S - 1) is the thermody- namic driver for ice formation, and is generated by expansion of air parcels forced by large scale dynamics, gravity waves, and small scale turbulence (Kim et al., 2003). At temperatures below 200 K (typically near the tropical tropopause layer, TTL) the simple conceptual model for cir- rus formation presented above is at odds with observations (Jensen et al., 2010; Kr¨ amer et al., 2009; Peter et al., 2006). Temperature fluctuations from mesoscale gravity waves are common at high altitudes and can produce localized vertical motion with updraft velocity as large as 1 m s -1 (Bacmeis- ter et al., 1999; Herzog and Vial, 2001; Jensen and Pfister, 2004; Sato, 1990). Such motion can increase the rate of ex- pansion cooling at the point of freezing so that a large num- ber of ice crystals is nucleated before the local supersatura- tion is depleted by ice crystal growth. Homogeneous freez- ing driven by gravity wave motion would produce high ice crystal number concentration, N c , between 1 and 10 cm -3 Published by Copernicus Publications on behalf of the European Geosciences Union.
Transcript

Atmos. Chem. Phys., 11, 3757–3771, 2011www.atmos-chem-phys.net/11/3757/2011/doi:10.5194/acp-11-3757-2011© Author(s) 2011. CC Attribution 3.0 License.

AtmosphericChemistry

and Physics

Dynamical states of low temperature cirrus

D. Barahona1,* and A. Nenes1,2

1School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, Atlanta, GA, USA2School of Earth and Atmospheric Sciences, Georgia Institute of Technology, Atlanta, GA, USA* now at: NASA Goddard Space Flight Center, Greenbelt, MD, USA

Received: 28 November 2010 – Published in Atmos. Chem. Phys. Discuss.: 20 December 2010Revised: 4 April 2011 – Accepted: 4 April 2011 – Published: 26 April 2011

Abstract. Low ice crystal concentration and sustained in-cloud supersaturation, commonly found in cloud observa-tions at low temperature, challenge our understanding of cir-rus formation. Heterogeneous freezing from effloresced am-monium sulfate, glassy aerosol, dust and black carbon areproposed to cause these phenomena; this requires low up-drafts for cirrus characteristics to agree with observationsand is at odds with the gravity wave spectrum in the uppertroposphere. Background temperature fluctuations howevercan establish a “dynamical equilibrium” between ice produc-tion and sedimentation loss (as opposed to ice crystal forma-tion during the first stages of cloud evolution and subsequentslow cloud decay) that explains low temperature cirrus prop-erties. This newly-discovered state is favored at low tempera-tures and does not require heterogeneous nucleation to occur(the presence of ice nuclei can however facilitate its onset).Our understanding of cirrus clouds and their role in anthro-pogenic climate change is reshaped, as the type of dynamicalforcing will set these clouds in one of two “preferred” micro-physical regimes with very different susceptibility to aerosol.

1 Introduction

Cirrus clouds are composed of ice crystals that form athigh altitudes and temperatures typically below 235 K (Prup-pacher and Klett, 1997). They play a key role in climateby modulating the planetary radiative balance (Liou, 1986)and heat transport in the upper troposphere (Ramanathan andCollins, 1991). They strongly impact water vapor transportacross the tropopause level (Jensen and Pfister, 2004) andplay an important role in lower stratospheric chemistry (Pe-

Correspondence to:A. Nenes([email protected])

ter, 1997). Cirrus may be affected by aircraft emissions (Se-infeld, 1998) and long range transport of pollutants and playan important (but highly uncertain) role in anthropogenic cli-mate change.

A key microphysical parameter required for understand-ing the climate impact of cirrus is their concentration of icecrystals,Nc. At temperatures between 200 and 235 K cirrusice crystals form primarily by homogenous freezing of super-cooled deliquesced aerosol (DeMott et al., 2003; Heymsfieldand Sabin, 1989), which occurs if the saturation ratio withrespect to ice,S, (i.e., the ratio of water vapor partial pres-sure to its equilibrium value over ice) reaches a characteris-tic threshold value,Shom (Koop et al., 2000). Heterogeneousfreezing of water upon existing aerosol particles (termed “icenuclei”, IN) can also occur (atS lower thanShom) and con-tribute to ice crystal concentrations (DeMott et al., 2003;Froyd et al., 2009), especially in polluted and dust-rich re-gions (Barahona et al., 2010a; Haag et al., 2003). The levelof water vapor supersaturation (i.e.,S −1) is the thermody-namic driver for ice formation, and is generated by expansionof air parcels forced by large scale dynamics, gravity waves,and small scale turbulence (Kim et al., 2003).

At temperatures below 200 K (typically near the tropicaltropopause layer, TTL) the simple conceptual model for cir-rus formation presented above is at odds with observations(Jensen et al., 2010; Kramer et al., 2009; Peter et al., 2006).Temperature fluctuations from mesoscale gravity waves arecommon at high altitudes and can produce localized verticalmotion with updraft velocity as large as 1 m s−1 (Bacmeis-ter et al., 1999; Herzog and Vial, 2001; Jensen and Pfister,2004; Sato, 1990). Such motion can increase the rate of ex-pansion cooling at the point of freezing so that a large num-ber of ice crystals is nucleated before the local supersatura-tion is depleted by ice crystal growth. Homogeneous freez-ing driven by gravity wave motion would produce high icecrystal number concentration,Nc, between 1 and 10 cm−3

Published by Copernicus Publications on behalf of the European Geosciences Union.

3758 D. Barahona and A. Nenes: Dynamical states of low temperature cirrus

near the TTL (Barahona et al., 2010a). Such high concentra-tions however are not observed;Nc remains low, sometimeseven lower (0.005–0.2 cm−3) than concentrations observedin weak updraft zones at cold temperatures (Kramer et al.,2009; Lawson et al., 2008). This “lowNc” paradox is ac-companied by other phenomena, such as low supersaturationrelaxation times (Kramer et al., 2009), which in turn leads tosustained supersaturation levels inside clouds (i.e., “the su-persaturation puzzle”) (Gao et al., 2004; Kramer et al., 2009;Peter et al., 2006), high clear-sky supersaturation (Jensen etal., 2005), and broad ice crystal size distributions (i.e., largecrystal sizes, Jensen et al., 2008). These phenomena occurdespite the strong dynamical forcing and the ample amountsof deliquesced aerosol available for homogeneous freezing.Suppressed freezing by organics (Murray, 2008), slow wa-ter vapor transfer to the ice phase (Gao et al., 2004; Mageeet al., 2006), and freezing to cubic instead of hexagonal ice(Murray et al., 2005), have been proposed to explain thesefeatures. These mechanisms are however not capable of ex-plaining the lowNc in low temperature cirrus clouds (Peteret al., 2006). Lacking the predictive understanding of suchphenomena hinders the ability of climate models to capturethe climate effects of cirrus clouds and their response to an-thropogenic perturbations.

Heterogeneous freezing of IN as the main path of cirrusformation has been proposed to explain the features of cir-rus clouds at low temperature (Abbatt et al., 2006; Jensenet al., 2010; Murray et al., 2010). Owing to their ability tofreeze ice at much lower supersaturation than homogeneousfreezing requires, IN can deplete water vapor, reduce super-saturation and inhibit homogeneous freezing; this can drasti-cally reduce the number of ice crystals that forms in the cirrus(Barahona and Nenes, 2009b; DeMott et al., 1994; Karcheret al., 2006). Much of the anthropogenic impact on cirrusclouds and climate is thought to occur through this IN-Ncfeedback mechanism (Lohmann and Feichter, 2005). Dust(Khvorostyanov et al., 2006), effloresced ammonium sulfate(Abbatt et al., 2006; Jensen et al., 2010; Wise et al., 2010),and glassy aerosol (Murray et al., 2010) have been identifiedas potential heterogeneous IN at the TTL .

The evolution of cirrus clouds at lowT has been addressedin several studies. Using a one dimensional (1-D) cloudmodel, Jensen and Pfister (2004) found that the superposi-tion of temperature fluctuations along Lagrangian trajecto-ries near the TTL resulted in rapid cooling cycles that in-creased the rate of crystal production by homogeneous nu-cleation leading toNc above 1 cm−3. Khvorostyanov etal. (2006) used a 1-D cirrus model to investigate the evo-lution of a cirrus layer initialized at 200 K and found thatregardless of the predominant nucleation mechanism, tem-perature fluctuations increased the maximumNc in the cloud(up to 0.6 cm−3). Nc however rapidly decreased after the ini-tial freezing pulse due to the vertical advection of ice crys-tals precluding new nucleation events and dilutingNc downto about 0.05 cm−3. Gensch et al. (2008) used box model

simulations along Lagrangian trajectories to test homoge-neous and heterogeneous nucleation scenarios in the forma-tion of cirrus at lowT . It was found that only heterogeneousnucleation scenarios (with prescribed IN number concentra-tion around 0.1 cm−3) would result inNc close to observa-tions. This conclusion was echoed by Froyd et al. (2009) andJensen et al. (2010) using 1-D models along prescribedT

trajectories.The formation of cirrus clouds however exhibits complex

non-linear behavior that may not be captured by box and 1-D models. Ice falling through active freezing zones (typi-cally located at the top of the cirrus layer, Spichtinger andGierens, 2009b) in clouds consume water vapor and can in-hibit homogeneous freezing much like IN do (Kay et al.,2007; Spichtinger and Gierens, 2009b). Their effectivenessdepends on their residence time in freezing zones and hencedepends on their size. Large ice crystals tend to quickly fallout of freezing zones and have limited effect on new ice for-mation events; small crystals (typically those with terminalvelocity, uterm, less or equal to the mean updraftu of thecirrus layer) fall slowly and can remain long enough in theupper part of the cloud to affect new freezing events. Thissuggests that at low temperatures, preexisting (and typicallysmall, Kramer et al., 2009) ice crystals may locally dehydratethe freezing zone sufficiently to inhibit the formation of newice. The rate of crystal production is not uniform through thefreezing zone, as the “local” saturation ratio,S, and updraftvelocity,u (defined at the scale of individual cloud “parcels”∼100–102 m, Pruppacher and Klett, 1997) may be affectedby fluctuations in wind speed and temperature induced bygravity waves (Jensen et al., 2010; Karcher and Haag, 2004;Kim et al., 2003). These internalS variations are usually ne-glected in cirrus cloud studies on the basis that the long-termevolution of the cloud is determined by the mean values ofS

andu.In this work we analyze the range of conditions for which

heterogeneous freezing may explain the features of cirrusclouds at low temperature, and propose an alternative view(based on a statistical description of cirrus formation andevolution) in which the interplay of temperature fluctuations,and ice crystal production and sedimentation leads to pre-viously unidentified natural cirrus states of low ice crystalconcentration and sustained high supersaturation.

2 Heterogeneous freezing at low temperature

The impact of IN onNc depends on their concentration,NIN . If too low (NIN < 1× 10−4 cm−3), a negligible im-pact is seen onNc, as too few (heterogeneously-frozen)ice crystals form to quench supersaturation below the ho-mogeneous/heterogeneous freezing threshold (Barahona andNenes, 2009b). LowNc favors large crystal size and there-fore heterogeneously frozen ice crystals may sediment out ofthe cloud layer before significantly modifyingS (Spichtinger

Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/

D. Barahona and A. Nenes: Dynamical states of low temperature cirrus 3759

and Gierens, 2009a). WhenNIN approaches a characteristic“limiting” concentration (which depends on updraft velocity,the IN freezing threshold and size),Nlim , supersaturation isquenched, homogeneous freezing is suppressed, andNc de-creases steeply (Barahona and Nenes, 2009b). ForNIN ≥

Nlim , homogeneous nucleation is inhibited andNc = NIN .Thus, Nlim is the minimumNc that can form in an activenucleation zone in a freshly-formed cirrus cloud (Barahonaand Nenes, 2009b) and presents the maximum reduction inNc possible from IN.

Figure 1 showsNc accounting for the competition betweenhomogeneous and heterogeneous freezing during cloud for-mation (Barahona and Nenes, 2009b) (details of the calcu-lation are described Sect. 3.3). AsNc is strongly influencedby the vertical velocity at the point of freezing and much lesssensitive to smallT variations, the curves in Fig. 1 representthe peakNc that would be obtained in box models after a sin-gle freezing event (Barahona and Nenes, 2008; Hoyle et al.,2005; Karcher and Lohmann, 2002; Spichtinger and Gierens,2009a). IfNIN is always very close toNlim , competition be-tween homogeneous and heterogeneous freezing could yieldNc close to observations. This requiresNIN ∼ 0.1 cm−3,which is 20-fold higher than measured dust concentrations(∼.005 cm−3) at the tropopause level (Froyd et al., 2009).Ammonium sulfate aerosol is present at much higher concen-trations than dust, and can serve as IN (Abbatt et al., 2006;Wise et al., 2010) if a fraction of them is effloresced (whichis possible, given that it deliquesces at∼90% relative humid-ity) (Fountoukis and Nenes, 2007; Shilling et al., 2006).

To inhibit homogeneous freezing and reproduce obser-vations ofNc, the concentration of ammonium sulfate INneeds to be within 10% ofNlim ; if concentrations fall below0.9Nlim , homogeneous freezing is triggered and predictedNcis significantly above observations (Fig. 1). If higher concen-tration thanNlim is present, homogeneous freezing is com-pletely suppressed, but too many crystals still form (Bara-hona and Nenes, 2008). In fact, if all ammonium sulfate isavailable as IN,Nc from heterogeneous freezing and pure ho-mogeneous freezing are always comparable (Fig. 2), becausecrystals formed from ammonium sulfate IN are very small(with size close to the dry aerosol; 0.02–0.05 µm, Froyd etal., 2009) and grow too slowly to quench supersaturation be-fore a large fraction of the aerosol freezes heterogeneously.Nc is within observed values only if the average size ofcrystals at the point of freezing is 2 µm or larger (Fig. 2),which is too large for upper tropospheric aerosol (Froyd etal., 2009). Experimental studies suggest that heterogeneousfreezing of ammonium sulfate IN atT ∼ 240 K can be veryselective (about 1 in 105 particles nucleate ice, Shilling etal., 2006). If the same selectivity maintains at lowerT , toofew IN would be available to prevent homogeneous freezing(therefore resulting in highNc). Higher nucleation selectivity(e.g., about 1 in 102 particles actively nucleating ice) wouldresult in complete inhibition of homogeneous freezing andstill maintainNc close to observations (not shown). A pure

Fig. 1. Ice crystal concentration,Nc, as a function of updraftvelocity, u. The cloud was assumed to form atT = 185 K andp = 100 hPa (details provided in Sect. 3.3.1). Low values ofu cor-respond to cloud formation driven primarily by large scale dynam-ics, whereasu > 50 cm s−1 is characteristic of cirrus developing inthe vicinity of convective systems with intense gravity wave break-ing (Kim et al., 2003). Solid line indicateNc calculated for purehomogeneous freezing, dashed line forNIN = Nlim , and dotted forNIN = 0.75Nlim . ForNIN = Nlim , Nc lies close to the observed val-ues foru < 50 cm s−1 (Kramer et al., 2009) but is very sensitive tosmall fluctuations inNIN .

heterogeneous scenario of ice nucleation on ammonium sul-fate however implies a maximum supersaturation below 20%(Fig. 3b), i.e.,S greater than 1.2 would be rarely observedas newly formed crystals would rapidly remove supersatura-tion. This is at odds with observations of relative humiditythat suggest that clear-sky supersaturation in the vicinity ofcirrus up to 70% (and occasionally above) is very common atlow temperature (Kramer et al., 2009). Hence the lowNc andhigh S observed at high-level cirrus can be reconciled withbox-model results only if the concentration of ammoniumsulfate IN is remarkable constant (0.1± 0.01 cm−3), the con-centration of dust is exceptionally large, or, the fluctuationsin vertical velocity from gravity wave motion are neglected.

The freezing fraction of organic glassy aerosol can bemuch lower than that of ammonium sulfate and maintainNIN close toNlim (hence yield lowNc, Fig. 3a) if the ver-tical velocity is below 15 cm s−1 (Murray et al., 2010). Atlarger updrafts however, homogeneous nucleation is trig-gered, producing highNc (Fig. 3a). When integrated overa normal distribution of updrafts with standard deviationσu(Fig. 3c), Nc remains within observed values forσu up to40 cm s−1 at T = 195 K. The onset of homogeneous nucle-ation occurs at even loweru for colder temperatures andNc deviates from observations forσu as low as 10 cm s−1

at T = 185 K. Predominance of heterogeneous nucleationfrom glassy IN would also imply maximum supersaturationaround 30% if the new formed ice crystals efficiently remove

www.atmos-chem-phys.net/11/3757/2011/ Atmos. Chem. Phys., 11, 3757–3771, 2011

3760 D. Barahona and A. Nenes: Dynamical states of low temperature cirrus

Fig. 2. Simulations of ice crystal concentration by pure heteroge-neous freezing.Nc is presented as a function of the initial size of theice nuclei. Conditions (Lawson et al., 2008) used wereT = 185 K,p = 100 hPa, and vapor-to-ice deposition coefficient,αd, of 0.07(dashed line), and 1.0 (solid line). The IN population was assumedto be monodisperse with total number concentration of 100 cm−3

(Lawson et al., 2008). Freezing of solid ammonium sulfate wasassumed to occur in a “burst” around the heterogeneous freezingthreshold described by sigmoidal freezing spectrum with inflectionpoint Shet= 15% (Abbatt et al., 2006), where 99% of the aerosolfreeze within a 2% supersaturation interval aboutShet (inset plot).

supersaturation (Barahona et al., 2010a; Murray et al., 2010).However as the freezing fraction of glassy IN is small,S canincrease even after heterogeneous nucleation has occurred.Still, S would have to remain belowShom for Nc to remainlow unless homogeneous freezing is suppressed. The laterscenario is however not supported by observations. Both, in-cloud and clear-sky RH are generally limited by the homoge-neous freezing threshold indicating efficient supersaturationremoval by homogeneously-frozen ice crystals (Kramer etal., 2009; Selkirk et al., 2010). Although uncertainty in RHcan be typically up to 20% (Kramer et al., 2009), it is stillsmaller than the difference between the homogeneous andheterogeneous freezing thresholds, typically between 30%and 40%, giving this support to the idea that homogeneousfreezing occurs at low temperatures. All together, this im-plies that in the presence of (ubiquitous)T fluctuations, thepresence of glassy IN may contribute, but not fully accountfor the observed characteristics of lowT cirrus.

3 Parcel statistical ensemble model

The main processes affecting the evolution ofNc and meansaturation ratio,So, within a cirrus layer are the freezing ofnew ice, the sedimentation of existing ice crystals, the lift-ing of air masses (which generates supersaturation), and therelaxation (i.e., mass transfer) of water vapor to/from the ice

Fig. 3. Comparison between heterogeneous effects from solid am-monium sulfate (Abbatt et al., 2006) and glassy citric acid aerosol(Murray et al., 2010), using the analytical model of Barahona andNenes (2009a) for homogeneous and heterogeneous freezing.(a)Maximum ice crystal concentration as a function of updraft velocityfor a single freezing event.(b) Maximum supersaturation achievedfor a single freezing event.(c) Ice crystal concentration averagedover a normal distribution of updraft velocities with zero mean andstandard deviationσu. The gray lines represent the range ofNctypically observed (Kramer et al., 2009).

phase. The magnitude of each process can be expressed interms of a characteristic timescale, i.e.,τfr, τsed, τlift , andτrelfor freezing, sedimentation, lifting, and relaxation, respec-tively. Fluctuations inS andu can have a strong impact onall cloud processes; we therefore represent them in terms of a

Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/

D. Barahona and A. Nenes: Dynamical states of low temperature cirrus 3761

probability distribution centered about the cirrus-average sat-uration ratio,So, and vertical velocity,u. The width of theseprobability distributions is largely determined by the meanamplitude of temperature fluctuations,δT (Bacmeister et al.,1999; Hoyle et al., 2005; Karcher and Burkhardt, 2008). Therate of ice production is given by the frequency with whichS exceeds the homogeneous freezing threshold (Karcher andBurkhardt, 2008) times the length and intensity of each freez-ing event (henceτfr) (Barahona and Nenes, 2008; Pruppacherand Klett, 1997). The same fluctuations also affect the localmass transfer rate between the ice and vapor phases, so thatwhen averaged over the cloud, water deposition/sublimationoccurs at an “effective” saturation ratio,Seff, that may differfrom So.

Supersaturation and crystal number in the cirrus cloudare determined using a “Lagrangian statistical ensemble” ap-proach. This involves determining the time-dependant stateof i homogeneous Lagrangian “parcels” that move with a(time-dependant) vertical velocity,ui ; ensemble averagingof the parcel solutions (outlined below) weighted by the ap-propriate probability distribution give approximate equationsthat describe the time-dependant properties for the whole cir-rus. From these considerations, simple equations can be de-rived that represent the evolution ofNc andSo in the cirrus(Sects. 3.1, 3.2).

3.1 Evolution of saturation ratio

In the absence of ice nucleation, the rate of change of satu-ration ratio,S, within the ith Lagrangian parcel is given by(Barahona and Nenes, 2009a; Seinfeld and Pandis, 1998)

dSi

dt= αuiSi −γ

∞∫Dmin

D2c,i

dDc,i

dtnc,i(Dc)dDc (1)

whereα =g1HsMwcpRT 2 −

gMaRT

and γ =ρi

ρa

π2

Map

Mwpoi, 1Hs is the

latent heat of sublimation of water,g is the acceleration ofgravity, cp is the heat capacity of air,po

i is the ice satura-tion vapor pressure atT (Murphy and Koop, 2005),p is theambient pressure,Mw andMa are the molar masses of wa-ter and air, respectively,R is the universal gas constant,ρi

andρa are the ice and air densities, respectively, andDc isthe volume-equivalent diameter of an ice particle (assumingspherical shape).nc,i(Dc) is the ice crystal size distributionin theith parcel, and

dDc,i

dt=

G(Si −1)

Dc,i(2)

where G ≈

[ρiRT

4poi D

vMw+

1Hsρi

4kaT

(1HsMw

RT−1

)]−1

, ka is the

thermal conductivity of air,D′

v = D′

v(T ,p,αd) is the watervapor diffusion coefficient from the gas to ice phase correctedfor non-continuum effects, andαd is the water vapor depo-

sition coefficient. Substituting Eq. (2) into Eq. (1) providesafter evaluation of the integral,

dSi

dt= αuiSi −

(Si −1)

τrel,i(3)

whereτrel,i =(βNc,iDc,i

)−1is the relaxation time scale in

the ith parcel,β = γG, and,Nc,i , Dc,i are the concentrationand mean size of ice crystals in theith parcel, respectively.

Equation (3) provides the supersaturation “state” for everyLagrangian parcel considered in the ensemble. Knowledgeof the distribution ofui (from the spectrum of gravity wavesin the cirrus) can then be used to “drive” the parcels in the en-semble to find the resulting distribution ofSi . Averaging iscarried out first over all parcels reaching a given cloud levelwith vertical velocityuj (referred to as the “j th cloud veloc-ity state”), and then averaging over all cloud states. Based onthis, the average saturation ratio,So, of the cloud over a timeinterval1t is

So(t) =

+∞∫−∞

∫X(t)

1∫0

Si(µ,x,τ )P (µ,x,τ )dτdxdµ (4)

whereµ =uu, u andu are the instantaneous and average verti-

cal velocity, respectively,x denotes the position in the cloud,τ =

t ′

1t, wheret ′ is the averaging time, andX(t) is the do-

main of x. P(µ,x,τ ) is the normalized probability at timet ′ of finding a parcel between positionx andx +dx (wheredx =

dxdydzVcloud

), with vertical velocity withinu andu+du.Equation (1) can be simplified, by considering that fluctu-

ations generated by gravity waves are random in nature (i.e.,follow a Gaussian distribution, Fig. 4d). Thus, under the as-sumption thatP(µ,x,τ ) does not vary with space and timeover1t , P(µ,x,τ ) ' P(µ) and Eq. (4) simplifies to

So(t) =

+∞∫−∞

∫X(t)

1∫0

Si(µ,x,τ )P (µ)dτdxdµ (5)

Equation (5) assumes thatSo is affected by processes that actthroughout the volume of the cirrus cloud. Other processes,like entrainment and radiative cooling, are neglected. Al-though this will not affect the conclusions of our study, theycould be included in future studies e.g. indirectly throughappropriate modification of the vertical velocity distribution(Barahona and Nenes, 2007).

DefiningSj =∫

X(t)

1∫0

Si(µ,x,τ )dτdx as the average super-

saturation of parcels in the “j ” velocity state over the timeinterval1t , Eq. (5) can be rewritten as

So(t) =

+∞∫−∞

Sj (µj )P (µj )dµj (6)

the time derivative of which gives,

www.atmos-chem-phys.net/11/3757/2011/ Atmos. Chem. Phys., 11, 3757–3771, 2011

3762 D. Barahona and A. Nenes: Dynamical states of low temperature cirrus

Fig. 4. Evolution of a cirrus cloud under pure homogeneous freez-ing, lifting at 1 cm s−1 with initial T = 195 K and cloud thickness,H = 500 m, andαd = 1. Shown are(a) the ice crystal number con-centration,(b) mean supersaturation,(c) characteristic timescalesof freezing (gray dots), relaxation (solid lines), and sedimentation(dotted lines), and,(d) frequency distribution of vertical velocity,for different values of the mean amplitude temperature fluctuations,δT .

dSo

dt=

+∞∫−∞

dSj (µj )

dtP (µj )dµj+

+∞∫−∞

Sj (µj )dP (µj )

dtdµj (7)

the second integral on the right hand side of Eq. (7) de-pends on the source of vertical velocity fluctuations. Distantsources of gravity waves result in stationaryP(µj ), hencedP (µj )

dt→ 0. HoweverP(µj ) can be perturbed by near con-

vective and orographic sources; in such casesP(µj ) is notcompletely Gaussian and exhibits a tail towards high veloc-ities (Bacmeister et al., 1999). For the purpose of this study

it is assumed thatdP (µj )

dt= 0, which implies that the char-

acteristic amplitude of temperature fluctuations,δT , remainsconstant during the entire period of simulation. Equation (7)then becomes

dSo

dt≈

+∞∫−∞

dSj (µj )

dtP (µj )dµj (8)

Using the definition ofSj ,

dSj

dt=

∫X(t)

1∫0

dSi(µ,x,τ )

dtdτdx (9)

Substitution of Eq. (3) into above provides

dSj

dt=

∫X(t)

1∫0

{αujSi−

(Si −1)

τrel,i

}dτdx=αuj

∫X(t)

1∫0

Sidτdx

∫X(t)

1∫0

(Si −1

τrel,i

)dτdx (10)

which can be rewritten as,

dSj

dt= αuj Sj −

∫X(t)

1∫0

(Si −1

τrel,i

)dτdx (11)

IntroducingSeff,j so that,

∫X(t)

1∫0

(Si −1

τrel,i

)dτdx =

(Seff,j −1

) ∫X(t)

1∫0

1

τrel,idτdx (12)

From Eq. (3),

1

τrel,j=

∫X(t)

1∫0

1

τrel,idτdx =

[βNcDc

]µ=µj

(13)

Combining Eqs. (12) and (13), Eq. (11) can be written as

dSj

dt= αuj Sj −

Seff,j −1

τrel,j(14)

whereτrel,j is the relaxation time scale associated with thej th state. Seff,j is an “effective” saturation ratio for de-position/sublimation processes, defined below. IntroducingEq. (14) into Eq. (8),

dSo

dt=

+∞∫−∞

(αuj Sj −

Seff,j −1

τrel,j

)P(µj )dµj (15)

or,

dSo

dt=

+∞∫−∞

αuj SjP(µj )dµj−

+∞∫−∞

(Seff,j−1

τrel,j

)P(µj )dµj (16)

The first term in the right hand side of Eq. (16) must be equalto uSo, as in the absence of deposition/sublimation,So inthe layer increases exponentially with time (Pruppacher andKlett, 1997). With this, Eq. (16) becomes,

dSo

dt=

So

τlift−

+∞∫−∞

(Seff,j −1

τrel,j

)P(µj )dµj (17)

whereτlift = (αu)−1. Equation (17) must be solved for eachtime step specifyingP(µj ) and then evaluatingSeff,j and theintegral on the right hand side. SinceP(µj ) is determined

Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/

D. Barahona and A. Nenes: Dynamical states of low temperature cirrus 3763

by the random overlapping of gravity waves of different fre-quency and amplitude (e.g.,uj is given by a Fourier seriesin time, Sect. 3.3), then for a time step of integration muchsmaller thanτlift (∼102 s) Eq. (17) can be approximated by

dSo

dt= αuSo−

Seff,j −1

τrel,j

∣∣∣∣∣µ=µj

(18)

whereSeff,j andτrel,j are calculated at the instantaneous ver-tical velocity.

Equation (18) gives the evolution of theSo in the cirruscloud; its solution however requires the knowledge ofSeff,j .This is accomplished by considering the properties of the dif-ferent parcels reaching the cloud layer att . For example, ifSi in theith parcel is a pseudo-steady state,dSi

dt∼ 0 (Korolev

and Mazin, 2003) and from Eq. (3),

Si,ss=τlift ,i

τlift ,i −τrel,i(19)

whereτlift ,i = (αui)−1 and Si,ss is the steady state satura-

tion ratio in theith parcel. If τlift ,i < 0 thenSi,ss< 1, andvice-versa. Thus, ifuj < 0, the layer would likely be sub-saturated over1t (e.g., Eq. 6), and vice-versa whenuj >

0. Thus, depending on the sign ofuj there is net deposi-tion/sublimation of water vapor in the cloud layer. Not allparcels however reach steady state; therefore the degree ofsaturation/subsaturation associated with thej th state dependson the probability distribution of saturation within the cloudylayer,Ps(S,So,δT ), which is a function ofSo and the aver-age amplitude of temperature fluctuations,δT . Thus,Seff foruj < 0 is found by averaging over all states that would leadto subsaturation, i.e.,Ps(S,δT ,So) for which S < 1. Sim-ilarly, when uj > 0, the supersaturated (S > 1) region ofPs(S,δT ,So) is used,

Seff,j =

b∫aS

dPs(S,δT ,So)dS

dS

b∫a

dPs(S,δT ,So)dS

dS

(20)

where a=

{1 uj > 00 uj ≤ 0

andb =

{Shom uj > 01 uj ≤ 0

The homogeneous freezing threshold,Shom, is set as the up-per limit of Ps(δT ,So) as ice crystal production quickly re-moves supersaturation aboveShom (Karcher and Burkhardt,2008; Karcher and Haag, 2004).

3.2 Evolution of ice crystal number concentration

The evolution of the number concentration within a cloudylayer is given by

dNc

dt=

dNc

dt

∣∣∣∣fr

+dNc

dt

∣∣∣∣sed

(21)

where dNcdt

∣∣∣fr

is the rate production of ice crystals within

the layer, anddNcdt

∣∣∣sed

is their sedimentation rate. If homo-

geneous and heterogeneous nucleation are active,dNcdt

∣∣∣fr

is

given by,

dNc

dt

∣∣∣∣fr

=dNc

dt

∣∣∣∣fr,hom

+dNc

dt

∣∣∣∣fr,het

where dNcdt

∣∣∣fr,hom

and dNcdt

∣∣∣fr,het

are the ice crystal produc-

tion rates from homogeneous and heterogeneous nucleation,respectively.

3.2.1 Homogeneous nucleation

Ice crystal production by homogeneous nucleation is drivenby local motions and occurs within single parcels whenSi > Shom. The maximum ice crystal concentration producedby homogeneous freezing within theith parcel is given by(Barahona and Nenes, 2008; Pruppacher and Klett, 1997)

Nc,i = No

1−exp

tmax,i∫0

voJ (Si)dt

(22)

wheretmax,i is the time at which ice crystal nucleation stops,J is the homogeneous nucleation rate coefficient andNo, voare the deliquesced aerosol number concentration and av-erage volume, respectively. Taking the time derivative ofEq. (22) gives,

dNc,i

dt

∣∣∣∣fr,hom

= NovoJ (Si)exp

tmax,i∫0

voJ (Si)dt

(23)

which can be approximated by (Barahona and Nenes 2008)

dNc,i

dt

∣∣∣∣fr,hom

≈ NovoJmax,i exp

−vo

αui

Smax∫0

J (Si)dSi

(24)

whereJmax,i = J (Smax,i). Smax,i is the maximum saturationratio reached in theith parcel, calculated by settingdSi

dt= 0

in Eq. (1),

Smax,i =γ

αui

∞∫Do

D2c,i

dDc,i

dtni,nuc(Dc)dDc (25)

whereni,nuc(Dc) is the size distribution of the recently nucle-ated ice crystals, andDo is the mean size of the deliquescedaerosol. Equation (25) assumes that only recently nucleatedice crystals are contained within the parcel. In reality, a frac-tion of preexisting crystals remain in nucleation zones (typ-ically located near the cloud top, Spichtinger and Gierens,2009b) inhibiting the homogeneous freezing of ice. Ice crys-tals experience gravitational settling, hence only those crys-tals with terminal velocity,uterm, belowu would be found at

www.atmos-chem-phys.net/11/3757/2011/ Atmos. Chem. Phys., 11, 3757–3771, 2011

3764 D. Barahona and A. Nenes: Dynamical states of low temperature cirrus

the cloud top. Adding the consumption of water vapor frompreexisting crystals to the right hand side of Eq. (25) gives

Smax,i =γ

αui

∞∫Do

D2c,i

dDc,i

dtni,nuc(Dc)dDc

+

Dterm∫Dmin

D2cdDc

dtnc(Dc)dDc

(26)

wherenc(Dc) is the cloud ice crystal size distribution,Dtermis the size of the crystal for whichuterm = u, andDmin isthe minimum size of the preexisting crystals in the cloud.Equation (26) can be combined with Eq. (2) to obtain

Smax,i =γ

αui

∞∫Do

D2c,i

dDc,i

dtni,nuc(Dc,Smax,i)dDc

+GNcDcfps(Smax,i −1)]

(27)

where,fps=1

NcDc

Dterm∫Dmin

Dcnc(Dc)dDc, is the fraction of pre-

existing ice crystals remaining in nucleation zones. As icecrystals remaining in the cloud layer were produced by pre-existing freezing events, Eq. (27) provides a link between thehistory of different parcels and the nucleation of new crystals.The analytical solution of Eq. (27) is presented elsewhere(Barahona and Nenes, 2009b; Barahona et al., 2010b).

The rate of ice crystal production by homogeneous nucle-ation in thej th cloud velocity state is given by the concentra-tion of nucleated crystals over the freezing timescale,

dNc,j

dt

∣∣∣∣fr,hom

=Ps(S >Shom)No

τhom,j

Hv(uj ) (28)

where τ−1hom,j = voJmax,j exp

(−

voαuj

Smax∫0

J (Sj )dSj

).

Hv(uj ) is the Heaviside function and is introduced becausehomogeneous nucleation is very unlikely in parcels withnegative vertical velocity (i.e., updraft must be maintainedfor some time beforeShom is reached after which it is quicklydepleted by crystal nucleation and growth, Barahona andNenes, 2008; Karcher and Lohmann, 2002) .Ps(S > Shom)

represents the fraction of parcels for whichS > Shom. Usingthe same averaging procedure as for the supersaturationequation, we obtain

dNc

dt

∣∣∣∣fr,hom

= No Ps(S >Shom)Hv(µj )

τhom,j

∣∣∣∣µ=µj

(29)

3.2.2 Heterogeneous nucleation

The formulation of the ice crystal production rate by het-erogeneous freezing is simplified by using the ice nucle-ation spectrum,NIN(S,T ) (Barahona and Nenes, 2009a) .

Assuming thatNIN(S,T ) is weakly dependent onT andNIN(S,T ) = No,hetfhet(S), then following Eq. (29) we write,

dNc

dt

∣∣∣∣fr,het

= No,hetdfhet

dS

dS

dtPs(S >Shet)Hv(µj )

∣∣µ=µj

(30)

where,No,het is the total number concentration of IN species,andfhet, Shet its freezing fraction and heterogeneous nucle-ation threshold, respectively. AsNIN is usually small,S isnot immediately depleted by ice crystal growth anddS

dtcan

be approximated by the instantaneous rate of increase of su-persaturation (this is further justified in the case of glassy INas dfhet(S)

dSis constant , Murray et al., 2010). Therefore,

dNc

dt

∣∣∣∣fr,het

=No,hetdfhet

dSαujS Ps(S>Shet)Hv(µj )

∣∣µ=µj

(31)

which can be written as

dNc

dt

∣∣∣∣fr,het

= No,hetPs(S >Shet)Hv(µj )

τhet,j

∣∣∣∣µ=µj

(32)

whereτhet,j =

(dfhetdS

αujS)−1

.

3.2.3 Sedimentation of ice crystals

Sedimentation processes out of the cloud layer depend pri-marily on the bulk properties of the cloud, i.e., the mean icecrystal size distribution and number concentration (interac-tion of individual parcels with falling crystals within the layeris accounted for in Eq. 27). The ice crystal loss rate by sedi-mentation is then given by,

dNc

dt

∣∣∣∣sed

=1

H

∞∫Dmin

uterm(Dc)n(Dc)dDc (33)

whereH is the cloud layer thickness. Asuterm∼ Dc (Heyms-field and Iaquinta, 2000), Eq. (33) can be further simplifiedto

dNc

dt

∣∣∣∣sed

=Ncuterm

H=

Nc

τsed(34)

whereuterm= uterm(Dc).

3.3 Numerical solution

3.3.1 Competition between homogeneous andheterogeneous freezing

Calculation of ice crystal number concentration,Nc in in-situ cirrus from combined homogeneous and heterogeneousnucleation in Figs. 1 and 3 is done using an analytical pa-rameterization developed for in situ formed cirrus clouds andfreezing fractions below 0.6 (Barahona and Nenes, 2009b).When the calculated freezing fraction exceeds 0.6, a sig-moidal increase inNc is assumed (Barahona et al., 2010a),

Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/

D. Barahona and A. Nenes: Dynamical states of low temperature cirrus 3765

in agreement with parcel model simulations and field obser-vations (Barahona and Nenes, 2008; DeMott et al., 2003).For combined homogeneous and heterogeneous nucleation,it was assumed that the IN freeze instantaneously at a super-saturation freezing threshold,shet, of 15%, typical of deposi-tion mode IN (Abbatt et al., 2006), with a 0.1 µm diameter atfreezing (Froyd et al., 2009). Glassy aerosol was assumed tohave a total concentration of 50 cm−3 and a freezing fractiongiven by the nucleation spectrum of Murray et al. (2010).

3.3.2 Vertical velocity spectrum

Observations suggest that the spectrum of gravity wave per-turbations near the tropopause is pseudorandom in nature, aresult of the superposition of waves from different sources,and therefore varies temporally and spatially (Bacmeister etal., 1999; Jensen and Pfister, 2004; Kim et al., 2003; Sato,1990). On average, the associated vertical velocity spec-trum can be approximated by a Gaussian function, althoughthis may underestimate the frequency of high amplitude per-turbations (Bacmeister et al., 1999; Herzog and Vial, 2001;Kim et al., 2003). A representative spectrum of vertical ve-locity fluctuations can be generated using a Fourier series(Bacmeister et al., 1999; Jensen and Pfister, 2004) of theform u = u+

∑j

A($j )cos($j t +mH +ϕ) wherem is the

vertical wave number,H is the cloud thickness, and$j ,A($j ), andϕ, are the wave frequency, phase, and amplitude,respectively. We have adopted this representation as follows.For each cirrus simulation, a time series ofu was generatedover the frequency interval$ = [3.35×10−7,9.44×10−4

]

Hz (Jensen and Pfister, 2004), using randomly generatedϕ

andm. A($j ) was calculated using a power spectrum scal-ing law of −1.85 for $j > 1× 10−5 Hz and of−0.25 for$j ≤ 1×10−5 Hz (Herzog and Vial, 2001; Jensen and Pfis-ter, 2004). This procedure resulted in a normal distributionof u (Fig. 4d) centered aroundu. The maximum amplitudewas assumed to occur at$j = 1×10−3 Hz (Jensen and Pfis-ter, 2004) as it reproduces the results of Gayet et al. (2004)(Fig. 4 green line) which give positiveu around 0.23 m s−1

for δT = 1 K (i.e., A(1× 10−3) ≈ 2.1δT ). Representativetime series foru(t) are presented in Fig. 5a.

3.3.3 Ice crystal production

The homogeneous freezing timescale,τhom,j , was calculatedusing the parameterization of Barahona and Nenes (2008,2009b, a). Precursor aerosol was assumed to be composed ofdeliquesced ammonium sulfate, lognormally distributed withdry mean geometric diameter of 40 nm, geometric dispersionof 2.3, and number concentration of 100 cm−3 (Lawson etal., 2008). To account for possible compositional impactson crystal growth kinetics, the water-vapor deposition coef-ficient was varied between 0.006 (Magee et al., 2006) and1.0. Homogeneous freezing is described using the param-eterization of Koop et al. (2000). The termPs(S > Shom)

Fig. 5. Time series of(a) updraft velocity,u, (b) total water con-tent,qtot, andDc (c) mean ice crystal diameter,Dc, for the condi-tions presented in Fig. 4;(d) time series ofDc for initial S andNcof −0.2 and 0, respectively, and homogeneous and heterogeneousnucleation active (conditions similar to Fig. 8c).

in Eq. (28) is the probability of findingS aboveShom, andis introduced to account for the threshold behavior of ho-mogeneous freezing (Karcher and Burkhardt, 2008; Koop etal., 2000). The effect of preexisting ice crystals on freez-ing was accounted for by allowing a fraction ofNc to de-plete water vapor and increaseτhom,j (Barahona and Nenes,2009b; Barahona et al., 2010b). The fraction of preexist-ing crystals remaining in freezing zones was calculated as

fps=1

NcDc

Dterm∫Dmin

Dcnc(Dc)dDc, wheren(Dc) is the ice crys-

tal size distribution,Dmin is the minimum pre-existing crys-tal size, andDterm is the crystal size for which its termi-nal velocity, uterm, is equal to the uplift velocity of thecirrus later,u. uterm was calculated assuming ice crystalshave columnar shape with maximum dimension equal toDc(Heymsfield and Iaquinta, 2000). Following Heymsfield andPlatt (1984) it was assumedn(Dc) = AD−3.15

c ; the parame-tersA andDmin were calculated from the moments ofn(Dc):Nc =

∫∞

Dminn(Dc)dDc andDc =

1Nc

∫∞

DminDcn(Dc)dDc. The

calculation ofDc is described below. Integration of equationsof Eqs. (18) and (21) was accomplished using a fixed timestep of 2 s. Initial values forNc,0 = 0.01 cm−3 andSo = 1.0were set. Sensitivity to using different initial values affectedonly the time required to establish dynamic equilibrium (bya few hours) and is assessed in Fig. 8.

The heterogeneous freezing timescale,τhet,j , (e.g., Eq. 32)was calculated at each time step using the instantaneous ver-tical velocity,uj , and the effective supersaturationSeff,j . Forthe simulations presented in Sect. 4, the aerosol freezing frac-tion, fhet, was calculated from the spectrum of Murray etal. (2010). No,het andShet were set to 1 cm−3 and 1.35, re-spectively (Murray et al., 2010).

www.atmos-chem-phys.net/11/3757/2011/ Atmos. Chem. Phys., 11, 3757–3771, 2011

3766 D. Barahona and A. Nenes: Dynamical states of low temperature cirrus

3.3.4 Ice crystal sedimentation

The rate of ice crystal sedimentation over the cloud scale,H , was assumed proportional to the terminal velocity of themean crystal sizeDc (Eq. 33). Other removal processes (icecrystal sublimation and detrainment) are neglected;H how-ever was varied over a wide interval (100 to 5000 m) to ac-count for the uncertainty associated with neglecting theseprocesses.Dc was calculated so that the total water vaporin the layer was partitioned between ice and vapor phases,

i.e., Dc =

(6qice

πρiNc

)1/3whereqice = qtot −

poSoMwRT

, ρi is the

ice density (Pruppacher and Klett, 1997),R is the univer-sal gas constant,Mw is the molecular mass of water, andpo is the saturation water vapor pressure over ice (Murphyand Koop, 2005); the minimum ice crystal size was set to4 µm in agreement with theoretical studies and experimen-tal observations (Barahona and Nenes, 2008; Durran et al.,2009; Kramer et al., 2009). Loss of total water content,qtot, from the cloudy layer is also accounted for by solution

of dqtotdt

= −π6 ρiD

3c

dNcdt

∣∣∣sed

. Representative time profiles of

Dc andqtot are presented in Fig. 5. The timescale of relax-ation atu = uj , τrel,j , was calculated usingNc andDc of thecloud layer (e.q., Eq. 13). Mass transfer limitation from non-continuum effects are taken into account in the calculation ofτhom,j but neglected in the calculation ofτrel,j . The latter isjustified asDc > 10 µm.

4 Cirrus in dynamical equilibrium

The model developed in Sect. 3 is used to describe the evolu-tion of a cirrus layer at low temperature taking into accountthe effect of internalS variations onNc. Forward integrationof Eqs. (18) and (29) and (34) is carried out using the proce-dure described in Sect. 3.3. A wide range of initial conditionsand model parameters are selected to describe the cirrus evo-lution under different scenarios.

Figure 4 shows the evolution of a cirrus layer subject togravity-wave fluctuations with an initial average temperatureof 195 K and lifting atu = 1 cm s−1. Only homogeneousfreezing is considered (heterogeneous freezing is “switchedoff”, i.e., No,het= 0). For values ofδT > 1 K, the cloud ini-tially experiences a strong homogeneous nucleation pulse,so thatNc initially increases steeply (Fig. 4a); the consump-tion of water vapor by crystal growth decreasesSo (Fig. 4b)which prevents any new freezing events.Nc slowly decreasesfrom sedimentation loss; only after enough ice crystals sed-iment out of the cloud layer,So increases (e.g.,δT = 1 K,green lines) and new freezing events occur. ForδT > 1.4 K(purple lines) this is possible even if the layer remains onaverage slightly subsaturated (So ∼ 1) because the probabil-ity distribution of S is broad enough for a non-negligibleprobability withS > Shom. However it is likely that recentlyformed crystals will sublimate within a few hours in the sub-saturated environment returning the moisture to the layer (ice

crystal removal by sublimation is not considered). The cir-rus is then maintained by new, independent freezing events.This “pulse-decay” behavior is characterized byτsed� τrelso ice crystals reside long enough in the cloud to relax super-saturation (Fig. 4c); this behavior is also consistent with theparcel model concept of cirrus (where highNc and lowSocoexist within the parcel). The subsaturation levels (Fig. 4b)achieved in this state are in agreement with in situ observa-tions of relative humidity in dissipating clouds (Gao et al.,2004; Kramer et al., 2009).

The cirrus evolution is however quite different whenδT issmall; the distribution ofS is narrow, and substantial ice pro-duction is only possible after enough supersaturation (i.e.,So) builds up in the cloudy layer to allow a non-negligibleprobability whereS > Shom. Thus, freezing events produc-ing largeNc (associated with largeu fluctuations; Fig. 4d) areless frequent. LowNc allows the formation of large ice crys-tals (Fig. 5c, d) which sediment out of the layer before sub-stantially depleting supersaturation, leading to new freezingevents. This “dynamic equilibrium” between ice productionand loss is a previously unidentified microphysical regime ofcirrus, characterized byτsed∼ τrel,j (Fig. 4c); it maintainslow Nc and highSo in the cloudy layer (Fig. 4a, b) andis consistent with observations of low-temperature cirrus.Clouds in “dynamic equilibrium” also exhibit broad crys-tal size distribution, because large ice crystals coexist withfreshly-formed (small) crystals in the cloud. The averageice crystal size in this case converges to values around 15 to20 µm, in agreement with observations (Kramer et al., 2009).Before equilibrium is reached,Dc exhibits larger values dueto the model’s sensitivity to initial conditions (Fig. 5c, d).

When simulations (such as those of Fig. 4) are placed on a“state diagram” ofNc vs.So, the two microphysical regimesdescribed above clearly emerge. Examples are presented inFigs. 6 and 7 for a range model parameters and a varietyof δT (lines of distinct color). For example, decreasing thecloud thickness to 100 m (Fig. 4b) increases the sedimenta-tion rate (Eq. 34) allowingS to replenish quickly and facil-itating the onset of equilibrium. Similarly, the rate of icecrystal growth increases withT (Fig. 4d) increasing the av-erage ice crystal size (hence decreasingτrel andτsed) and fa-cilitating the onset of equilibrium. Dynamical equilibriumwould however lead to lowerNc than observed atT ∼ 225 K(Gayet et al., 2004; Kramer et al., 2009) and it is likely thatclouds at these conditions would evolve following a pulse-decay behavior. Figure 7 shows state diagrams for differentvalues ofu, T andH for the similar initial conditions. In gen-eral, progression towards a “dynamic equilibrium” is favoredwhen supersaturation replenishes quickly (i.e., at highu) asice crystal growth and sedimentation are favored (leading tolow τsed), and vice-versa for “pulse-decay” behavior. Rapidconsumption of supersaturation by the growing ice crystalsalso decreases the time between freezing pulses and replen-ishment ofS. Thus the period of theS andNc oscillations inthe dynamic equilibrium state is mainly controlled byτrel.

Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/

D. Barahona and A. Nenes: Dynamical states of low temperature cirrus 3767

Fig. 6. Sensitivity ofNc andSo evolution to cloud formation con-ditions for different values ofδT (color scheme same as in Fig. 4);(a) same conditions as in Fig. 4,(b) cloud thickness,H = 100 m(increased ice crystal removal rate),(c) deposition coefficient equalto 0.006 (Magee et al., 2006) (slow water vapor transfer), and(d)initial temperature 225 K and cloud lifting at 5 cm s−1. The yellowstar in each panel indicates initial conditions. The arrows indicatethe temporal progression along each trajectory. The integration timewas 40 h cases, except in(d) were it was 15 h.

Clouds in the “dynamic equilibrium” regime are also muchless sensitive to slow water vapor deposition than predictedby box-model simulations. Figure 6c shows that dynamicequilibrium is possible even forαd as low as 0.006 (com-pared toαd = 1 used in Figs. 4 and 6a for the same sim-ulation conditions). Still, the high rate of production of icecrystals forαd = 0.006 increasesNc and decreasesDc, there-fore increasingτsedand slowing the replenishment of super-saturation. Thus, dynamic equilibrium is only possible forδT < 0.8 as opposed toδT < 1.0 for αd = 1. Figure 6c how-ever shows that the existence of strong kinetic limitations tothe diffusional growth of ice crystals cannot be ruled out.

4.1 Effect of heterogeneous IN and initial conditions

It is important to study the sensitivity of the dynamical statesof cirrus to initial conditions used in the simulation and to thepresence of heterogeneous IN. Figure 8a showsS andNc forthe conditions of Figs. 4 and 6a but starting atSo,ini = −0.4andNc,o = 0. For this set of initial conditions, the onset ofoscillating behavior is delayed by a few hours before super-saturation is reached. This implies that the temperature of thefirst freezing pulse is lower than in the case withSo,ini = 1,slightly increasing its strength; the system however dampensout these variations and eventually follows a similar trajec-tory as in Fig. 6a.

a b

c d

e f

Fig. 7. Similar to Fig. 6, but varying cloud mean vertical velocity,u,initial layer temperature,To, cloud thickness,H , and mean ice crys-tal terminal velocity,uterm. The yellow star in each plot indicatesinitial conditions. The integration time was 40 h in foru = 1 cm s−1

and 15 h foru = 5 cm s−1.

Figure 8b and c present simulations where both homoge-neous and heterogeneous nucleation are active for differentinitial conditions. IN are assumed to originate from glassyaerosol withNo,het= 1 cm−3. By allowing ice crystal for-mation at lowS, heterogeneous IN decrease the ice crystalproduction rate by homogeneous nucleation hence lower themaximumNc. This makes the system more stable to highamplitude vertical velocity perturbations, facilitating the on-set of dynamic equilibrium, which can be maintained up toδT ∼ 1.2. The “stabilizing” effect of IN is exemplified inFig. 8c where homogeneous freezing was suppressed, i.e.,only heterogeneous freezing is active. This leads to a damp-ened response to vertical velocity fluctuations so that dynam-ical equilibrium is possible even forδT as high as 1.8 K.Thus, in a cirrus cloud where heterogeneous nucleation isdominant, dynamical equilibrium is very robust against ex-ternal perturbations. The values ofNc and S at the equi-librium state are however not significantly influenced by theinitial conditions nor the ice nucleation mechanism, e.g., af-ter sometime the system oscillates about the same values asin Figs. 6 and 7. This means that the presence of IN can“help” the cirrus to achieve dynamical equilibrium withoutmodifying the equilibrium values.

www.atmos-chem-phys.net/11/3757/2011/ Atmos. Chem. Phys., 11, 3757–3771, 2011

3768 D. Barahona and A. Nenes: Dynamical states of low temperature cirrus

Fig. 8. Similar to Fig. 6, varying initialS andNc and the activefreezing mechanism.

Figures 6 to 8 also show that the “dynamic equilibrium”state occurs spontaneously whenδT goes below a charac-teristic transition value (which depends onu, T and the pre-dominant freezing mechanism). It can also be reached after acloud initially resides in a “pulse-decay” state, ifδT is closeto the characteristic value (δT ∼ 1 K in Fig. 6). When max-imum Nc and time-averagedSo are presented on the statediagram for all simulations considered where homogeneousfreezing is active, the conditions ofδT that separate “pulse-decay” and “dynamic equilibrium” regimes seem to be uni-versal (Fig. 9). If only heterogeneous nucleation is active,the cloud resides mostly in the dynamic equilibrium state.

5 Conclusions and implications

From the discussion above, cold cirrus clouds will reside inthe “dynamic equilibrium” regime ifδT is below a charac-teristic threshold. High-amplitude, orographically-generatedgravity waves are ubiquitous (Kim et al., 2003) but oftenlose intensity with altitude, weakening their contribution tothe background spectrum of temperature fluctuations.δT

can thus decrease enough at high altitude for cirrus to transi-tion from a “pulse-decay” to a “dynamic equilibrium” state(Fig. 9). This would explain why lowNc and highSo are ob-served at low temperatures near the tropopause. Dynamicalequilibrium is also possible at warmer conditions (particu-larly for high u; Fig. 6d) but require smallδT ; given thathigh amplitude fluctuations are widespread at lower altitudes(Hoyle et al., 2005), cirrus clouds are likely forced to alwaysfollow a pulse-decaying behavior. Heterogeneous IN never-theless may help to stabilize the system so that the dynamicequilibrium manifests at higherδT than for clouds with purehomogeneous freezing.

Fig. 9. Maximum ice crystal concentration obtained during thecloud evolution simulations against the time-averaged mean satu-ration ratio. Results presented for all simulations carried out in thisstudy. Integration time varied between 15 and 40 h. Symbols arecolored by the value ofδT used. Regions where the cloud sponta-neously transitions to a “pulse-decay” and “dynamic equilibrium”state are noted; the “transitional” region marks where the cloud gen-erally initially exhibited “pulse-decay” behavior over few hours andthen transitioned to a “dynamic equilibrium” regime. (©) Homo-geneous freezing is active; (1) Homogeneous freezing suppressed.

In summary, cirrus clouds at low temperature exhibit char-acteristics (e.g., lowNc and sustained high saturation ratios)that cannot be explained with the simple “conventional” pic-ture of homogeneous freezing driven by expansion coolingin the presence of ubiquitous temperature fluctuations. Evenif heterogeneous nucleation is dominant, conventional mod-els of cirrus could explain the characteristics of low-T cir-rus only for weak updrafts, and require neglecting the higheramplitude components of the vertical velocity spectrum. In-stead, we show that small-scale fluctuations from the actionof gravity waves can switch a cloud into a previously un-known “dynamic equilibrium” regime, with sustained levelsof low Nc and high saturation ratios consistent with “puz-zling” characteristics observed in low temperature cirrus.

With this study, a new understanding for cirrus cloudsemerges, where the “unperturbed” microphysical state is oneof dynamical equilibrium with low crystal number and highsupersaturation. Only when the mean amplitude of tem-perature fluctuations exceeds a threshold value (δT ∼ 1 Kwhen homogeneous freezing is active) cirrus exhibit the well-known “pulse-decay” microphysical state. Throughout muchof the atmosphere, the latter state dominates, simply becauseδT is larger than the characteristic threshold value. In theTTL, δT is still remarkably large (0.6–0.8 K) (Bacmeister etal., 1999; Jensen and Pfister, 2004; Sato, 1990), but doesnot systematically exceed the threshold for “pulse-decay”behavior, so cirrus regress to their “unperturbed” dynamic-equilibrium state. The presence of heterogeneous IN can

Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/

D. Barahona and A. Nenes: Dynamical states of low temperature cirrus 3769

also dampen the effect of vertical velocity perturbations onice crystal production facilitating the transition to dynamicequilibrium. However, the existence of IN is not a neces-sary condition to explain the characteristics of cirrus cloudsat low temperature. Figure 6c shows that dynamic equilib-rium states can also exist even under conditions of strong ki-netic limitations (very lowαd) and therefore their existencecannot be ruled out based on simplified models of cirrus for-mation.

The structure and responses of cirrus to dynamical and mi-crophysical forcings can also be portrayed. For example,cirrus formed in the region of convective anvils might ex-hibit “pulse-decay” state until gravity-wave fluctuations de-cay to below theδT threshold and transition to a dynamic-equilibrium state. For the same reasons, IN impacts on cirrusproperties can be strong for clouds in pulse-decay state, butnot for clouds in dynamic equilibrium; e.g., IN can force thecloud to fall towards equilibrium however will not modifythe equilibrium state. In conclusion, the discovery of dy-namic equilibrium states reshapes our understanding of cir-rus clouds and their role in anthropogenic climate change, asthe type of dynamical forcing and the presence of IN will setthese clouds in one of two “preferred” microphysical regimeswith very different susceptibility to anthropogenic aerosol.

Acknowledgements.This study was supported by NASA ACMAPand a NSF CAREER award.

Edited by: J. H. Seinfeld

References

Abbatt, J. P. D., Benz, S., Cziczo, D. J., Kanji, Z., and Mohler, O.:Solid ammonium sulfate as ice nuclei: a pathway for cirrus cloudformation, Science, 313, 1770–1773, 2006.

Bacmeister, J., Eckermann, S. D., Tsias, A., Carslaw, K. S., andPeter, T.: Mesoscale temperature fluctuations induced by a spec-trum of gravity waves: A comparison of parameterizations andtheir impact on stratospheric microphysics, J. Atmos. Sci., 56,1913–1924, 1999.

Barahona, D. and Nenes, A.: Parameterization of clouddroplet formation in large scale models: including ef-fects of entrainment, J. Geophys. Res., 112, D16026;doi:16010.11029/16207JD008473, 2007.

Barahona, D. and Nenes, A.: Parameterization of cirrus formationin large scale models: Homogeneous nucleation, J. Geophys.Res., 113, D11211, doi:11210.11029/12007JD009355, 2008.

Barahona, D. and Nenes, A.: Parameterizing the competition be-tween homogeneous and heterogeneous freezing in ice cloud for-mation polydisperse ice nuclei, Atmos. Chem. Phys., 9, 5933–5948,doi:10.5194/acp-9-5933-2009, 2009a.

Barahona, D. and Nenes, A.: Parameterizing the competition be-tween homogeneous and heterogeneous freezing in cirrus cloudformation - monodisperse ice nuclei, Atmos. Chem. Phys., 9,369–381,doi:10.5194/acp-9-369-2009, 2009b.

Barahona, D., Rodriguez, J. M., and Nenes, A.: Sensitivityof the global distribution of cirrus Ice crystal concentration

to heterogeneous freezing, J. Geophys. Res., 115, D23213,doi:23210.21029/22010JD014273, 2010a.

Barahona, D., West, R. E. L., Stier, P., Romakkaniemi, S., Kokkola,H., and Nenes, A.: Comprehensively accounting for the effect ofgiant CCN in cloud activation parameterizations, Atmos. Chem.Phys., 10, 2467–2473,doi:10.5194/acp-10-2467-2010, 2010b.

DeMott, P. J., Meyers, M. P., and Cotton, R. W.: Parameteriza-tion and impact of ice initiation processes relevant to numericalmodel simulations of cirrus clouds, J. Atmos. Sci., 51, 77–90,1994.

DeMott, P. J., Cziczo, D. J., Prenni, A. J., Murphy, D. M., Krei-denweis, S. M., Thompson, D. S., Borys, R., and Rogers, D. C.:Measurements of the concentration and composition of nuclei forcirrus formation, Proc. Natl. Acad. Sci. USA, 100, 14655–14660,2003.

Durran, D. R., Dinh, T., Ammerman, M., and Ackerman, T.: Themesoscale dynamics of thin tropical tropopause cirrus, J. Atmos.Sci., 66, 2859–2873,doi:10.1175/2009JAS3046.1, 2009.

Fountoukis, C. and Nenes, A.: ISORROPIA II: a com-putationally efficient thermodynamic equilibrium model forK+Ca2+Mg2+NH+

4 Na+SO2−

4 NO−

3 Cl−H2O aerosols, Atmos.Chem. Phys., 7, 4639–4659,doi:10.5194/acp-7-4639-2007,2007.

Froyd, K. D., Murphy, D. M., Sanford, T. J., Thomson, D. S., Wil-son, J. C., Pfister, L., and Lait, L.: Aerosol composition of thetropical upper troposphere, Atmos. Chem. Phys., 9, 4363–4385,doi:10.5194/acp-9-4363-2009, 2009.

Gao, R. S., Popp, P. J., Fahey, D. W., Marcy, T. P., Herman, R. L.,Weinstock, E. M., Baumgardner, D. G., Garrett, T. J., Rosenlof,K. H., Thompson, T. L., Bui, P. T., Ridley, B. A., Wofsy, S. C.,Toon, O. B., Tolbert, M. A., Karcher, B., Peter, T., Hudson, P.K., Weinheimer, A. J., and Heymsfield, A. J.: Evidence that nitricacid increases relative humidity in low-temperature cirrus clouds,Science, 303, 516–520, 2004.

Gayet, J. F., Ovarlez, J., Shcherbakov, V., Strom, J., Schumann,U., Minikin, A., Auriol, F., Petzold, A., and Monier, M.: Cirruscloud microphysical and optical properties at southern and north-ern midlatitudes during the INCA experiment, J. Geophys. Res.,109, D20206, doi:20210.21029/22004JD004803, 2004.

Gensch, I. V., Bunz, H., Baumgardner, D. G., Christensen, L. E.,Fahey, D. W., Herman, R. L., Popp, P. J., Smith, J. B., Troy, R.F., Webster, C. R., Weinstock, E. M., Wilson, J., Peter, T., andKramer, M.: Supersaturations, microphysics and nitric acid par-titioning in a cold cirrus cloud observed during CR-AVE 2006:an observation-modelling intercomparison study, Environ. Res.Lett., 3, 035003,doi:10.1088/1748-9326/3/3/035003, 2008.

Haag, W., Karcher, B., Strom, J., Minikin, A., Lohmann, U., Ovar-lez, J., and Stohl, A.: Freezing thresholds and cirrus cloud forma-tion mechanisms inferred from in situ measurements of relativehumidity, Atmos. Chem. Phys., 3, 1791–1806,doi:10.5194/acp-3-1791-2003, 2003.

Herzog, A. and Vial, F.: A study of the dynamics of the equato-rial lower stratosphere by use of ultra-long-duration balloons, J.Geophys. Res., 106, 22745–22761, 2001.

Heymsfield, A. J. and Platt, C. M.: A parameterization of the parti-cle size spectrum of ice clouds in terms of the ambient tempera-ture and ice water content, J. Atmos. Sci., 41, 846–855, 1984.

Heymsfield, A. J. and Sabin, R. M.: Cirrus crystal nucleation byhomogenous freezing of solution droplets., J. Atmos. Sci., 46,

www.atmos-chem-phys.net/11/3757/2011/ Atmos. Chem. Phys., 11, 3757–3771, 2011

3770 D. Barahona and A. Nenes: Dynamical states of low temperature cirrus

2252–2264, 1989.Heymsfield, A. J. and Iaquinta, J.: Cirrus crystals terminal veloci-

ties, J. Atmos. Sci., 57, 916–938, 2000.Hoyle, C. R., Luo, B. P., and Peter, T.: The origin of high ice crystal

number densities in cirrus clouds, J. Atmos. Sci., 62, 2658–2579,2005.

Jensen, E. J. and Pfister, L.: Transport and freeze-drying inthe tropical tropopause layer, J. Geophys. Res., 109, D02207,doi:02210.01029/02003JD004022, 2004.

Jensen, E. J., Smith, J. B., Pfister, L., Pittman, J. V., Weinstock,E. M., Sayres, D. S., Herman, R. L., Troy, R. F., Rosenlof, K.,Thompson, T. L., Fridlind, A. M., Hudson, P. K., Cziczo, D. J.,Heymsfield, A. J., Schmitt, C., and Wilson, J. C.: Ice supersatu-rations exceeding 100% at the cold tropical tropopause: implica-tions for cirrus formation and dehydration, Atmos. Chem. Phys.,5, 851–862,doi:10.5194/acp-5-851-2005, 2005.

Jensen, E. J., Pfister, L., Bui, T. V., Lawson, P., Baker, B., Mo, Q.,Baumgardner, D., Weinstock, E. M., Smith, J. B., Moyer, E. J.,Hanisco, T. F., Sayres, D. S., Clair, J. M. St., Alexander, M. J.,Toon, O. B., and Smith, J. A.: Formation of large ('100 µm)ice crystals near the tropical tropopause, Atmos. Chem. Phys., 8,1621–1633,doi:10.5194/acp-8-1621-2008, 2008.

Jensen, E. J., Pfister, L., Bui, T.-P., Lawson, P., and Baumgardner,D.: Ice nucleation and cloud microphysical properties in tropicaltropopause layer cirrus, Atmos. Chem. Phys., 10, 1369–1384,doi:10.5194/acp-10-1369-2010, 2010.

Karcher, B. and Burkhardt, U.: A cirrus scheme for global cir-culation models, Q. J. Roy. Meteorol. Soc., 134, 1439–1461;doi:1410.1002/Qj.1301, 2008.

Karcher, B. and Lohmann, U.: A parameterization of cirrus cloudformation: homogeneous freezing of supercooled aerosols, J.Geophys. Res., 107, 4010, doi:4010.1029/2001JD000470, 2002.

Karcher, B. and Haag, W.: Factors controlling upper tro-pospheric relative humidity, Ann. Geophys., 22, 705–715,doi:10.5194/angeo-22-705-2004, 2004.

Karcher, B., Hendricks, J., and Lohmann, U.: Physicallybased parameterization of cirrus cloud formation for use inglobal atmospheric models, J. Geophys. Res., 111, D01205,doi:01210.01029/02005JD006219, 2006.

Kay, J. E., Baker, M., and Hegg, D.: Physical controls on oro-graphic cirrus inhomogeneity, Atmos. Chem. Phys., 7, 3771–3781,doi:10.5194/acp-7-3771-2007, 2007.

Khvorostyanov, V. I., Morrison, H., Curry, J. A., Baumgard-ner, D. G., and Lawson, P.: High supersaturation andmodes of ice nucleation in thin tropopause cirrus: simula-tion of the 13 July 2002 cirrus regional study of tropicalanvils and cirrus layers case, J. Geophys. Res., 111, D02201,doi:02210.01029/02004JD005235, 2006.

Kim, Y.-J., Eckermann, S. D., and Chun, H.-Y.: An overview of thepast, present, and future of gravity-wave drag parameterizationfor numerical climate and weather prediction models, Atmos.-Ocean, 41, 65–98, 2003.

Koop, T., Luo, B., Tslas, A., and Peter, T.: Water activity as the de-terminant for homogeneous ice nucleation in aqueous solutions,Nature, 406, 611–614, 2000.

Korolev, A. V. and Mazin, I. P.: Supersaturation of water vapor inclouds, J. Atmos. Sci., 60, 2957–2974, 2003.

Kramer, M., Schiller, C., Afchine, A., Bauer, R., Gensch, I., Man-gold, A., Schlicht, S., Spelten, N., Sitnikov, N., Borrmann, S.,

de Reus, M., and Spichtinger, P.: Ice supersaturations and cir-rus cloud crystal numbers, Atmos. Chem. Phys., 9, 3505–3522,doi:10.5194/acp-9-3505-2009, 2009.

Lawson, R. P., Pilson, B., Baker, B., Mo, Q., Jensen, E., Pfister,L., and Bui, P.: Aircraft measurements of microphysical prop-erties of subvisible cirrus in the tropical tropopause layer, At-mos. Chem. Phys., 8, 1609–1620,doi:10.5194/acp-8-1609-2008,2008.

Liou, K.: Influence of cirrus clouds on weather and climate pro-cesses: a global perspective, Mon. Weather Rev., 114, 1167–1199, 1986.

Lohmann, U. and Feichter, J.: Global indirect aerosol effects: a re-view, Atmos. Chem. Phys., 5, 715–737,doi:10.5194/acp-5-715-2005, 2005.

Magee, N., Moyle, A. M., and Lamb, D.: Experimental de-termination of the deposition coefficient of small cirrus-likeice crystals near−50◦C, Geophys. Res. Lett., 33, L17813,doi:17810.11029/12006GL026665, 2006.

Murray, B. J.: Inhibition of ice crystallisation in highly viscousaqueous organic acid droplets, Atmos. Chem. Phys., 8, 5423–5433,doi:10.5194/acp-8-5423-2008, 2008.

Murphy, D. M. and Koop, T.: Review of the vapour pressures of iceand supercooled water for atmospheric applications, Q. J. Roy.Meteorol. Soc., 131, 1539–1565, 2005.

Murray, B. J., Knopf, D. A., and Bertram, A. K.: The formation ofcubic ice under conditions relevant to Earth’s atmosphere, Na-ture, 434, 202–205, 2005.

Murray, B. J., Wilson, T. W., Dobbie, S., Cui, Z., Al-Jumur, S. M.R. K., Mohler, O., Schnaiter, M., Wagner, R., Benz, S., Niemand,M., Saathoff, H., Ebert, V., Wagner, S., and Karcher, B.: Hetero-geneous nucleation of ice particles on glassy aerosol under cirrusconditions, Nature Geosci., 3, 233–237, 2010.

Peter, T.: Microphysics and heterogeneous chemistry of polarstratospheric clouds Annu. Rev. Phys. Chem., 48, 785–822,1997.

Peter, T., Marcolli, C., Spichtinger, P., Corti, T., Baker, B., andKoop, T.: When dry air is too humid, Science, 314, 1399–1401,2006.

Pruppacher, H. R. and Klett, J. D.: Microphysics of clouds and pre-cipitation 2nd ed., Kluwer Academic Publishers, Boston, MA,954 pp., 1997.

Ramanathan, V. and Collins, W.: Thermodynamic regulation ofocean warming by cirrus clouds deduced from observations of1987 El Nino, Nature, 351, 27–32, 1991.

Sato, K.: Vertical wind disturbances in the troposphere and lowerstratosphere observed by the MU radar, J. Atmos. Sci., 47, 2803–2817, 1990.

Seinfeld, J. H.: Clouds, contrails and climate, Nature, 391, 837–838, 1998.

Seinfeld, J. H. and Pandis, S. N.: Atmospheric Chemistry andPhysics, John Wiley and Sons, New York, NY, USA, 1998.

Selkirk, H. B., Vomel, H., Valverde Canossa, J. M., Pfister, L.,Diaz, J. A., Fernandez, W., Amador, J., Stolz, W., and Peng,G. S.: Detailed structure of the tropical upper troposphere andlower stratosphere as revealed by balloon sonde observations ofwater vapor, ozone, temperature, and winds during the NASATCSP and TC4 campaigns, J. Geophys. Res., 115, D00J19,doi:10.1029/2009jd013209, 2010.

Shilling, J. E., Fortin, T. J., and Tolbert, M. A.: Depositional ice nu-

Atmos. Chem. Phys., 11, 3757–3771, 2011 www.atmos-chem-phys.net/11/3757/2011/

D. Barahona and A. Nenes: Dynamical states of low temperature cirrus 3771

cleation on crystalline organic and inorganic solids, J. Geophys.Res., 111, D12204; doi:12101.11029/12005JD006664, 2006.

Spichtinger, P. and Gierens, K. M.: Modelling of cirrus clouds– Part 2: Competition of different nucleation mechanisms, At-mos. Chem. Phys., 9, 2319–2334,doi:10.5194/acp-9-2319-2009,2009a.

Spichtinger, P. and Gierens, K. M.: Modelling of cirrus clouds –Part 1b: Structuring cirrus clouds by dynamics, Atmos. Chem.Phys., 9, 707–719,doi:10.5194/acp-9-707-2009, 2009b.

Wise, M. E., Baustian, K. J., and Tolbert, M. A.: Inter-nally mixed sulfate and organic particles as potential icenuclei in the tropical tropopause region, 107, 6693–6698,doi:6610.1073/pnas.0913018107, 2010.

www.atmos-chem-phys.net/11/3757/2011/ Atmos. Chem. Phys., 11, 3757–3771, 2011


Recommended