constraining the twomey effect from satellite observations ...j. quaas et al.: twomey from satellite...

21
Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020 © Author(s) 2020. This work is distributed under the Creative Commons Attribution 4.0 License. Constraining the Twomey effect from satellite observations: issues and perspectives Johannes Quaas 1 , Antti Arola 2 , Brian Cairns 3 , Matthew Christensen 4 , Hartwig Deneke 5 , Annica M. L. Ekman 6 , Graham Feingold 7 , Ann Fridlind 3 , Edward Gryspeerdt 8 , Otto Hasekamp 9 , Zhanqing Li 10 , Antti Lipponen 2 , Po-Lun Ma 11 , Johannes Mülmenstädt 11 , Athanasios Nenes 12,13 , Joyce E. Penner 14 , Daniel Rosenfeld 15 , Roland Schrödner 5 , Kenneth Sinclair 3,16 , Odran Sourdeval 17 , Philip Stier 4 , Matthias Tesche 1 , Bastiaan van Diedenhoven 3 , and Manfred Wendisch 1 1 Leipzig Institute for Meteorology, Universität Leipzig, Leipzig, Germany 2 Finnish Meteorological Institute, Kuopio,Finland 3 NASA Goddard Institute for Space Studies, New York, USA 4 Department of Physics, University of Oxford, Oxford, UK 5 Leibniz Institute for Tropospheric Research, Leipzig, Germany 6 Department of Meteorology and Bolin Centre for Climate Research, Stockholm University, Stockholm, Sweden 7 NOAA Earth System Laboratories, Chemical Science Laboratory, Boulder, USA 8 Space and Atmospheric Physics Group, Imperial College London, UK 9 SRON Netherlands Institute for Space Research, Utrecht, the Netherlands 10 Earth System Science Interdisciplinary Center and Department of Atmospheric and Oceanic Science, University of Maryland, College Park, USA 11 Pacific Northwest National Laboratory, Richland, USA 12 School of Architecture, Civil & Environmental Engineering, École Polytechnique Fédérale de Lausanne, Lausanne, Switzerland 13 Institute of Chemical Engineering Sciences, Foundation for Research and Technology Hellas, Patras, Greece 14 Department of Climate and Space Sciences and Engineering, University of Michigan, Ann Arbor, USA 15 Institute of Earth Sciences, Hebrew University of Jerusalem, Jerusalem, Israel 16 Department of Earth and Environmental Engineering, Universities Space Research Association (USRA), Columbia, MD 21046, USA 17 Université de Lille, CNRS, UMR 8518 – LOA – Laboratoire d’Optique Atmosphérique, Lille, France Correspondence: Johannes Quaas ([email protected]) Received: 24 March 2020 – Discussion started: 15 May 2020 Revised: 24 September 2020 – Accepted: 8 October 2020 – Published: 4 December 2020 Abstract. The Twomey effect describes the radiative forcing associated with a change in cloud albedo due to an increase in anthropogenic aerosol emissions. It is driven by the per- turbation in cloud droplet number concentration (1N d, ant ) in liquid-water clouds and is currently understood to exert a cooling effect on climate. The Twomey effect is the key driver in the effective radiative forcing due to aerosol–cloud interactions, but rapid adjustments also contribute. These adjustments are essentially the responses of cloud fraction and liquid water path to 1N d, ant and thus scale approxi- mately with it. While the fundamental physics of the influ- ence of added aerosol particles on the droplet concentration (N d ) is well described by established theory at the particle scale (micrometres), how this relationship is expressed at the large-scale (hundreds of kilometres) perturbation, 1N d, ant , remains uncertain. The discrepancy between process under- standing at particle scale and insufficient quantification at the climate-relevant large scale is caused by co-variability of aerosol particles and updraught velocity and by droplet sink processes. These operate at scales on the order of tens of me- Published by Copernicus Publications on behalf of the European Geosciences Union.

Upload: others

Post on 09-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Atmos. Chem. Phys., 20, 15079–15099, 2020https://doi.org/10.5194/acp-20-15079-2020© Author(s) 2020. This work is distributed underthe Creative Commons Attribution 4.0 License.

    Constraining the Twomey effect from satellite observations:issues and perspectivesJohannes Quaas1, Antti Arola2, Brian Cairns3, Matthew Christensen4, Hartwig Deneke5, Annica M. L. Ekman6,Graham Feingold7, Ann Fridlind3, Edward Gryspeerdt8, Otto Hasekamp9, Zhanqing Li10, Antti Lipponen2,Po-Lun Ma11, Johannes Mülmenstädt11, Athanasios Nenes12,13, Joyce E. Penner14, Daniel Rosenfeld15,Roland Schrödner5, Kenneth Sinclair3,16, Odran Sourdeval17, Philip Stier4, Matthias Tesche1,Bastiaan van Diedenhoven3, and Manfred Wendisch11Leipzig Institute for Meteorology, Universität Leipzig, Leipzig, Germany2Finnish Meteorological Institute, Kuopio,Finland3NASA Goddard Institute for Space Studies, New York, USA4Department of Physics, University of Oxford, Oxford, UK5Leibniz Institute for Tropospheric Research, Leipzig, Germany6Department of Meteorology and Bolin Centre for Climate Research, Stockholm University, Stockholm, Sweden7NOAA Earth System Laboratories, Chemical Science Laboratory, Boulder, USA8Space and Atmospheric Physics Group, Imperial College London, UK9SRON Netherlands Institute for Space Research, Utrecht, the Netherlands10Earth System Science Interdisciplinary Center and Department of Atmospheric and Oceanic Science,University of Maryland, College Park, USA11Pacific Northwest National Laboratory, Richland, USA12School of Architecture, Civil & Environmental Engineering, École Polytechnique Fédéralede Lausanne, Lausanne, Switzerland13Institute of Chemical Engineering Sciences, Foundation for Research and Technology Hellas, Patras, Greece14Department of Climate and Space Sciences and Engineering, University of Michigan, Ann Arbor, USA15Institute of Earth Sciences, Hebrew University of Jerusalem, Jerusalem, Israel16Department of Earth and Environmental Engineering, Universities Space Research Association (USRA),Columbia, MD 21046, USA17Université de Lille, CNRS, UMR 8518 – LOA – Laboratoire d’Optique Atmosphérique, Lille, France

    Correspondence: Johannes Quaas ([email protected])

    Received: 24 March 2020 – Discussion started: 15 May 2020Revised: 24 September 2020 – Accepted: 8 October 2020 – Published: 4 December 2020

    Abstract. The Twomey effect describes the radiative forcingassociated with a change in cloud albedo due to an increasein anthropogenic aerosol emissions. It is driven by the per-turbation in cloud droplet number concentration (1Nd, ant)in liquid-water clouds and is currently understood to exerta cooling effect on climate. The Twomey effect is the keydriver in the effective radiative forcing due to aerosol–cloudinteractions, but rapid adjustments also contribute. Theseadjustments are essentially the responses of cloud fractionand liquid water path to 1Nd, ant and thus scale approxi-

    mately with it. While the fundamental physics of the influ-ence of added aerosol particles on the droplet concentration(Nd) is well described by established theory at the particlescale (micrometres), how this relationship is expressed at thelarge-scale (hundreds of kilometres) perturbation, 1Nd, ant,remains uncertain. The discrepancy between process under-standing at particle scale and insufficient quantification atthe climate-relevant large scale is caused by co-variability ofaerosol particles and updraught velocity and by droplet sinkprocesses. These operate at scales on the order of tens of me-

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

  • 15080 J. Quaas et al.: Twomey from satellite

    tres at which only localised observations are available and atwhich no approach yet exists to quantify the anthropogenicperturbation. Different atmospheric models suggest diversemagnitudes of the Twomey effect even when applying thesame anthropogenic aerosol emission perturbation. Thus, ob-servational data are needed to quantify and constrain theTwomey effect. At the global scale, this means satellite data.There are four key uncertainties in determining 1Nd, ant,namely the quantification of (i) the cloud-active aerosol – thecloud condensation nuclei (CCN) concentrations at or abovecloud base, (ii) Nd, (iii) the statistical approach for inferringthe sensitivity of Nd to aerosol particles from the satellitedata and (iv) uncertainty in the anthropogenic perturbationto CCN concentrations, which is not easily accessible fromobservational data. This review discusses deficiencies of cur-rent approaches for the different aspects of the problem andproposes several ways forward: in terms of CCN, retrievalsof optical quantities such as aerosol optical depth suffer froma lack of vertical resolution, size and hygroscopicity infor-mation, non-direct relation to the concentration of aerosols,difficulty to quantify it within or below clouds, and the prob-lem of insufficient sensitivity at low concentrations, in ad-dition to retrieval errors. A future path forward can includeutilising co-located polarimeter and lidar instruments, ide-ally including high-spectral-resolution lidar capability at twowavelengths to maximise vertically resolved size distributioninformation content. In terms ofNd, a key problem is the lackof operational retrievals of this quantity and the inaccuracy ofthe retrieval especially in broken-cloud regimes. As for theNd-to-CCN sensitivity, key issues are the updraught distribu-tions and the role of Nd sink processes, for which empiricalassessments for specific cloud regimes are currently the bestsolutions. These considerations point to the conclusion thatpast studies using existing approaches have likely underesti-mated the true sensitivity and, thus, the radiative forcing dueto the Twomey effect.

    1 Introduction

    Cloud droplets in liquid-water clouds form on cloud con-densation nuclei (Aitken, 1880), a subset of the atmo-spheric aerosol particle population. The formation of clouddroplets in thermodynamic equilibrium is established text-book knowledge (Köhler, 1936). Whether an aerosol parti-cle acts as cloud condensation nuclei (CCN) at a given su-persaturation depends on its size and chemical composition,which determine the particle hygroscopicity (Dusek et al.,2006; Ma et al., 2013). If CCN concentrations at one su-persaturation level are known, CCN concentrations at othersupersaturation levels approximately scale with it accordingto the Twomey (1959) parameterisation. Here, we implic-itly consider a supersaturation level of 0.2 % unless other-wise stated. Supersaturation is generated in the large majority

    of clouds by updraughts. The rare exceptions are formationdue to radiative cooling (mainly fog events) or the mixing ofcold and dry with warm and moist air masses. Cloud-scaleupdraughts originate in most cases from turbulence, con-vection or gravity waves. Updraught velocity, w, exhibits alarge heterogeneity across temporal and spatial scales (Tont-tila et al., 2011; Moeng and Arakawa, 2012). For a givenprobability density function (PDF) of updraughts, in an adia-batic air parcel with no active collision and coalescence, theaddition of extra CCN will generally lead to a monotonic in-crease in cloud droplet number concentration, Nd (Twomeyand Warner, 1967). The approximate functional form of thedependence of Nd on CCN concentration is then logarith-mic, since the increase in Nd associated with activation ofadditional aerosol leads to a depletion of the maximum su-persaturation (Twomey, 1959).

    The CCN concentration in the atmosphere is increasedby anthropogenic emission of aerosols and aerosol precursorgases (Boucher et al., 2013). This leads to enhanced Nd, un-less aerosol particle concentrations are high and updraughtsweak (Ghan et al., 1998; Feingold et al., 2001; Reutter et al.,2009). In turn, cloud albedo (αc, the fraction of solar radiativeenergy reflected back to space by clouds in relation to thatincident at the cloud top) increases, as it is a monotonicallyincreasing function of Nd. Following Platnick and Twomey(1994) and Ackerman et al. (2000),

    ∂αc

    ∂ lnNd=

    13αc (1−αc) , (1)

    a formulation which relies on (i) a two-stream radiative trans-fer approximation and (ii) the assumption that clouds obeyvertical stratification that scales with an adiabatic one andthat is horizontally homogeneous. Equation (1) is expressedas a partial derivative: other quantities – notably cloud waterpath – are considered constant.

    These two facts – Nd is a monotonic function of CCN andαc in the partial-derivative sense is a monotonic function ofNd – imply that the anthropogenic increase in CCN concen-trations causes a negative (cooling) radiative forcing due toaerosol–cloud interactions, RFaci (Boucher et al., 2013), de-noted as Faci (Bellouin et al., 2020b). It can be approximately(neglecting absorption in the column above the cloud afterscattering at cloud top) written as (Quaas et al., 2008; Bel-louin et al., 2020b)

    Faci = F↓s ·∂αc

    ∂ lnNd·∂ lnNd∂ lna

    ·1 lnaant, (2)

    with the downward solar radiative flux density (irradiance)above clouds, F↓s , and a quantitative description of CCN de-noted here as a. The relative anthropogenic perturbation toa is denoted 1 lnaant. This formulation assumes (i) that onlythe solar spectrum is relevant, which is well justified for theoptically thick, liquid water clouds considered here, since anNd perturbation only marginally changes the cloud radiative

    Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020

  • J. Quaas et al.: Twomey from satellite 15081

    effect in the terrestrial spectrum of an optically thick cloudand (ii) that there is one liquid water cloud layer that deter-mines the effect so that the problem can be considered purelyhorizontal in space. In contrast to the formulation by Bellouinet al. (2020b), we consider the problem as horizontally vari-able in space (x, y) and in time (t), i.e. Faci = Faci(x, y, t).If Eq. (2) is assessed from temporally sparse satellite data,a proper integration over temporally varying solar zenith an-gles and cloud diurnal cycles is necessary.

    RFaci is often referred to as the “Twomey effect”(Twomey, 1974) and also called the “(first) aerosol indi-rect effect” or “cloud albedo effect” (Lohmann and Feichter,2001). Atmospheric models simulate a large range for RFaci(Gryspeerdt et al., 2020; Smith et al., 2020). It is, thus, nec-essary to constrain the Twomey effect quantitatively basedon observations. Only satellites can provide global observa-tional data that could be used to quantify the global RFaci(Stephens et al., 2019).

    The Twomey effect has been assessed in many studies(starting with Bréon et al., 2002) in terms of cloud dropleteffective radius, re, rather than using Nd. This is plausibleas, for idealised vertical profiles of droplet size distributions(e.g. vertically constant or adiabatically increasing profiles),cloud optical depth and cloud albedo are easily expressedin terms of re (Hansen and Travis, 1974; Stephens, 1978).Given that re is closely related to light-scattering proper-ties of clouds in the visible and near-infrared, this quantityis operationally retrieved from remote-sensing observations(Nakajima and King, 1990). However, re is not just a func-tion of Nd but also varies with cloud liquid water path, L(Brenguier et al., 2000). It is thus necessary to formulate theproblem for constant L, which is difficult to realise in dataanalysis from observations that are limited in time and space,or for selected cloud scenarios, so that datasets stratified byL become too small for meaningful analysis (Quaas et al.,2006; McComiskey and Feingold, 2012; Liu and Li, 2019).Specifically, in Eq. (2), the middle term, ∂ lnNd

    ∂ lna , would be for-mulated as ∂ lnre

    ∂ lna , in which case the evaluation of the partialderivative requires stratification by L, in addition to the up-draught regime, which adds substantial complexity.

    Among the four factors on the right-hand side of Eq. (2),the first one, F↓s , is well quantified for each given latitude,longitude and time. The second one, ∂αc/∂ lnNd, can beevaluated using Eq. (1) (Bellouin et al., 2020b; Hasekampet al., 2019a), or alternatively by radiative-transfer simula-tions (Mülmenstädt et al., 2019). This implies that the twokey problems in determining RFaci are the quantificationof the anthropogenic perturbation of CCN, 1 lnaant, andthe sensitivity of Nd to CCN perturbations, β = ∂Nd/∂ lna(Feingold et al., 2001). Taken together, this is the distribu-tion of the anthropogenic perturbation of Nd (here expressedin absolute, not relative, terms):

    1Nd, ant =∂Nd

    ∂ lna

    ∣∣∣∣w

    ·1 lnaant = β(w) ·1 lnaant. (3)

    The plausible range of the sensitivity is 0≤ β ≤ 1, except forheavily polluted situations (where it may become negative;Feingold et al., 2001), or when giant CCN play an importantrole (Ghan et al., 1998; Morales Betancourt and Nenes, 2014;Gryspeerdt et al., 2016; McCoy et al., 2017) where com-petition for water vapour during droplet formation is at itsstrongest. Such conditions represent a significant challengeto models and parameterisations of the process (Morales Be-tancourt and Nenes, 2014).

    The aerosol forcing has to be evaluated at a scale muchlarger than an individual cloud. One of the key reasons forthis is that there is currently no way to use satellite data todetermine the anthropogenic fraction of the CCN populationfor a single air parcel. Methods applying model information,or data-tied approaches such as Bellouin et al. (2013) insteaduse the scale of model resolution or aggregate data resolutionwhich is typically of the order of 1◦ × 1◦ (or about 100×100 km2). The problem formulated in Eq. (3) then has to bereformulated, using an overbar to denote the averaging overa 1◦× 1◦ grid box as

    1Nd, ant =

    ∞∫w=−∞

    ∂Nd

    ∂ lna

    ∣∣∣∣w

    P(w)P(a)dw

    1 lnaant= β ·1 lnaant, (4)

    which considers the mean sensitivity of Nd to CCN, β, giventhe probability density function (PDF) of cloud base up-draught velocity, w in the grid box, P(w); the PDF of CCNat cloud base within the scene, P(a); and the anthropogenicperturbation of the CCN concentration at the grid-box scale,1 lnaant. Note in the above equation, β is assumed indepen-dent of lnaant, which assumes that P(w) is independent ofcloud properties (primarily, liquid water content), which ap-plies to stratus clouds (Morales and Nenes, 2010) but not ingeneral. Similarly, the covariance of P(w) and P(a)may notbe zero (e.g. Kacarab et al., 2020 – in addition to Bougiatiotiet al., 2020). All of the above suggest that observation of β ata cloud parcel scale is not directly transferrable to the largescale for an assessment of the Twomey effect. Rather, β hasto be estimated.

    Beyond RFaci, aerosol–cloud interactions also lead torapid adjustments: once cloud droplet size distributions arealtered due to anthropogenic CCN, cloud microphysical anddynamical processes are modified as well (Albrecht, 1989;Ackerman et al., 2000; Wang et al., 2003; Heyn et al., 2017;Mülmenstädt and Feingold, 2018). Aerosols can induce tran-sitions between cloud regimes, for instance by changing driz-zle behaviour (Rosenfeld et al., 2006; Feingold et al., 2010;Wood et al., 2011). The direction and magnitude of thesechanges depends on the cloud state and regime, because re-sponses to aerosol changes occur due to processes spanning arange from microphysics to the mesoscale (Christensen andStephens, 2012; Kazil et al., 2011; Wang et al., 2011). Theseprocesses include precipitation suppression (Albrecht, 1989),

    https://doi.org/10.5194/acp-20-15079-2020 Atmos. Chem. Phys., 20, 15079–15099, 2020

  • 15082 J. Quaas et al.: Twomey from satellite

    rapid feedbacks involving cloud-top entrainment (Ackermanet al., 2004; Bretherton et al., 2007; Hill et al., 2009; Bula-tovic et al., 2019) and rapid feedbacks involving cloud lateralentrainment (Xue and Feingold, 2006; Small et al., 2009) aswell as responses in dynamics (Xue et al., 2008; Stevens andFeingold, 2009; Wang and Feingold, 2009). If one also con-siders deep clouds, further intricate cloud adjustments mayoccur that are not considered here (e.g. Ekman et al., 2011;Fan et al., 2013; Yan et al., 2014). As a result of these adjust-ment processes, cloud horizontal extent (Gryspeerdt et al.,2016) and liquid water path (Gryspeerdt et al., 2019) re-spond to perturbations in Nd. The sum of RFaci and the ra-diative effects of these adjustments is the effective radiativeforcing due to aerosol–cloud interactions, ERFaci (Boucheret al., 2013). Based on modelling and data analysis, it is ev-ident that the adjustments and, thus, also ERFaci scale with1Nd, ant (Bellouin et al., 2020b; Gryspeerdt et al., 2020; Mül-menstädt et al., 2019). Analysis of model data shows thatthe rapid adjustments due to other contributions (small-scaleto mesoscale circulation changes, thermodynamic changes)are small (Heyn et al., 2017; Mülmenstädt et al., 2019).Even so, thermodynamic and dynamic adjustments to aerosolchanges can still have an important impact on droplet forma-tion – especially under conditions where droplet formationis largely velocity-limited (Kacarab et al., 2020; Bougiatiotiet al., 2020).

    Despite the fact that the activation of an individual CCNto form a droplet is well understood in thermodynamic equi-librium (Köhler, 1936), it is not clear how Nd responds toperturbations of CCN at the scale of a cloudy air parcel, anentire cloud or of a cloud field up to the large scale of theorder of 1◦× 1◦ as used in Eq. (4). A one-to-one relation-ship between CCN in the updraught below cumulus and Ndabove the cloud base within the cumulus has been observed(Werner et al., 2014); although even at the cloud updraughtscale, this relationship could be a convolution of the effect ofCCN on droplet number, vertical velocity variability and lat-eral entrainment (Morales et al., 2011). At a larger scale, thisrelation is less pronounced (Boucher and Lohmann, 1995),consistent with the expectation from Eq. (4). In turn, theremay be co-variability of updraughts and aerosol concentra-tions that lead to larger β compared to situations with con-stant w (Kacarab et al., 2020; Bougiatioti et al., 2017, 2020).

    Ground-based remote-sensing methods provide data to in-fer the sensitivity term β from long-term observations (Fein-gold et al., 2003; McComiskey et al., 2009; Schmidt et al.,2015; Liu and Li, 2018). However, this approach is limitedto individual sites and cloud regimes. In consequence, wheninvestigating the global radiative forcing relevant for climatestudies, the sensitivity term necessarily is derived from satel-lite remote sensing (Nakajima and Schulz, 2009).

    This leads to a number of problems and challenges dis-cussed in more detail in the following sections.

    - Retrieval of CCN. The first issue is the missing co-incidence of cloud and aerosol retrievals. Usually, noaerosol is retrieved below or within clouds. It is thusquestionable how representative aerosol in cloudlessscenes is for (neighbouring) cloud base CCN. The sec-ond issue is the imperfect nature of proxies for CCN.Often the aerosol optical depth (AOD; see below) or avariant thereof is used, which can only imperfectly berelated to CCN due to differences in sensitivity and thelack of vertical resolution.

    - Retrieval of Nd . There are (i) retrieval errors and biasesin Nd, which depend on cloud regimes, and (ii) oneneeds to consider the link between Nd as formed byCCN activation at cloud base and the retrieved cloud-topNd. Cloud-topNd (Nd, top) is the one that determinesthe scattering of sunlight and, thus, is relevant for thetop-of-atmosphere cloud radiative effect. It differs fromcloud base Nd (Nd, base) in conditions where Nd sinkssuch as precipitation or mixing play a role. When usingre rather thanNd the additional problem of stratificationby retrieved L arises.

    - Cloud-regime dependence. Cloud base droplet concen-tration, Nd, base, is a function of both CCN and up-draught, and Nd, top is further a function of Nd sinkssuch as precipitation formation and entrainment mixing.Thus, one needs to understand how the characteristics ofw and its PDF, as well as precipitation and mixing pro-cesses, depend on cloud regime and how this may beused for an empirical estimation of β.

    - Aggregation scale. The relation of aggregate quantitiesis not the same as the aggregate relation, and, thus,one needs to determine how to derive β optimally fromremote-sensing data (Grandey and Stier, 2010; Mc-Comiskey and Feingold, 2012).

    In practical terms, one further needs to assess to which ex-tent a simple scalar sensitivity metric is sufficient, or whethera joint-PDF approach is preferable (McComiskey and Fein-gold, 2012; Gryspeerdt et al., 2017).

    Beyond these questions which are discussed in the follow-ing sections, it is necessary to quantify the anthropogenicperturbation to CCN, 1 lnaant, which is not easily quanti-fied from observations. The key problem is that there is lit-tle potential to observe an atmosphere unperturbed by an-thropogenic emissions (Carslaw et al., 2013, 2017). Somestudies attempt to quantify the anthropogenic perturbation tothe column aerosol light extinction, or aerosol optical depth(AOD; τa), in a data-tied approach (Kaufman et al., 2005;Bellouin et al., 2005, 2013; Kinne, 2019). Such approachesrely on simplifying parameterisations, such as the assump-tion that small-mode aerosol particles are predominantly an-thropogenic. The other option is to estimate it from simula-tions (Quaas et al., 2009b; Gryspeerdt et al., 2017). There are

    Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020

  • J. Quaas et al.: Twomey from satellite 15083

    some indirect ways to infer the anthropogenic impacts on Nd(Quaas, 2015), such as from trends (Krüger and Graßl, 2002;Bennartz et al., 2011) or periodicity in anthropogenic emis-sions such as the weekly cycle (Quaas et al., 2009a). Hence,models are involved in determining an anthropogenic pertur-bation of CCN concentrations, which can even be attemptedfor individual weather events (Schwartz et al., 2002). In anycase, it seems impossible to know the anthropogenic pertur-bation to the aerosol at the scale of an air parcel; rather it ispossible only at larger, aggregate scales. The remainder ofthis review will focus on the sensitivity term β.

    2 Remote sensing of CCN concentrations

    The aerosol quantity most accessible to passive satellite re-mote sensing is AOD (Kaufman et al., 2002). It is derivedfrom the multi-spectral reflectance of the Earth–atmospheresystem using the incident solar radiation and retrieving or as-suming surface albedo characteristics as well as aerosol ab-sorption coefficient and scattering phase functions. There arefour key issues with using the retrieved AOD for estimatingtheNd to CCN sensitivity, which will be discussed in the fol-lowing subsections.

    - AOD is the vertical integral of the extinction coefficient.For the sensitivity of Nd to the aerosol, one needs toknow the vertical distribution of the CCN concentration,most importantly the CCN at cloud base.

    - AOD is an optical integral and does not provide infor-mation on the aerosol size distribution and its hygro-scopicity. The use of AOD does not isolate aerosol par-ticles that have the size and chemical composition toserve as CCN. It is also affected by aerosol swelling dueto hygroscopic growth.

    - AOD can be derived only for pixels determined to becloud-free. The degree to which this correlates withthe CCN at the base of (neighbouring) clouds is ques-tionable. In addition, retrieved AOD can show a posi-tive bias due to enhanced reflectance from neighbouringcloudy pixels or due to the lack of detecting spuriousclouds in a retrieval scene.

    - The optical signal is very weak at low concentrations.Therefore, retrievals become more and more uncertainbelow a certain aerosol load, especially over land and insituations with variable or uncertain surface albedo.

    At aggregate scales, i.e. for monthly averages over re-gions, AOD from ground-based remote-sensing retrievals(AERONET; Holben et al., 2001) correlates well with CCNsurface measurements (Andreae, 2009; Shen et al., 2019).Similar results were also reported for aircraft measurements(Clarke and Kapustin, 2010; Shinozuka et al., 2015). How-ever, at shorter timescales or less spatial aggregation, there

    are significant deviations from a perfect correlation (Liu andLi, 2014). AOD due to aerosol light extinction is determinedby the vertical integral of the extinction cross section, pro-portional to the vertical integral of the second moment of theaerosol size distribution. In turn, for a given chemical com-position of aerosol particles, the CCN concentration is thezeroth moment of the size distribution for particles exceed-ing a size threshold that depends on supersaturation. In thefollowing, the different problems are discussed in more de-tail, together with options for a better proxy for CCN fromsatellite remote sensing.

    2.1 Vertical co-location

    Stier (2016) investigated the correlation between AOD andCCN as represented in a climate model. He confirmed amostly positive correlation of the temporal variability of thetwo quantities, although in some regions the correlation islow or even negative. A key reason for the partly low cor-relation is the fact that AOD is a vertically integrated quan-tity and may include aerosol layers that are not interactingwith clouds. A similar result was reported from a statisti-cal analysis of satellite data: cloud microphysical parame-ters correlate well with aerosol properties only if the verticalalignment of the aerosol and cloud layers is accounted for(Costantino and Bréon, 2010, 2013). More recently, Paine-mal et al. (2020) demonstrate a much higher correlation be-tween Nd and aerosol extinction coefficients below cloudtop sampled from satellite lidar compared to Nd vs. AOD.Ship measurements of CCN and microwave-retrieved Nd atcloud base between Los Angeles and Hawaii show a weakerβ metric as the boundary layer deepens, thus indicating thatsurface aerosol measurements become less representative foraerosol variability at cloud base as the boundary layer deep-ens (Painemal et al., 2017), or that the updraughts becomehigh enough to activate smaller aerosols than the accumula-tion mode. In situ observations suggest that AOD may evenbe anticorrelated with CCN at cloud base (Kacarab et al.,2020).

    A way forward is the use of spaceborne vertically resolvedobservations such as lidar measurements (Shinozuka et al.,2015; Stier, 2016). The Cloud-Aerosol Lidar and InfraredPathfinder Satellite Observations (CALIPSO; Winker et al.,2009) lidar retrieves aerosol backscatter profiles and thus iscapable of identifying aerosol layers (Costantino and Bréon,2010). Profiles of aerosol particle extinction are inferredfrom these backscatter profiles by using typical extinction-to-backscatter ratios based on aerosol type. However, the signalis not sensitive to smaller aerosol concentrations, which ham-pers a quantitative analysis at the large scale (Watson-Parriset al., 2018; Ma et al., 2018). For situations with sufficientaerosol loading for reliable CALIPSO aerosol profile ob-servations, methods for retrieving CCN concentrations fromground-based lidar measurements can be adapted (Feingoldand Grund, 1994; Lv et al., 2018; Haarig et al., 2019). These

    https://doi.org/10.5194/acp-20-15079-2020 Atmos. Chem. Phys., 20, 15079–15099, 2020

  • 15084 J. Quaas et al.: Twomey from satellite

    methods apply empirical extinction-to-particle-concentrationrelationships to obtain input for CCN concentrations for dif-ferent aerosol types (Mamouri and Ansmann, 2016). In thefuture, the EarthCARE satellite mission currently scheduledfor launch in 2022 (Illingworth et al., 2015; Hélière et al.,2017) shows promise to extend and improve upon the successof the CALIPSO mission. Its Atmospheric Lidar (ATLID)is a linearly polarised high-spectral-resolution lidar (HSRL)operating at a wavelength of 355 nm. The instrument allowsthe direct inference of profiles of aerosol backscatter andextinction coefficients, thereby substantially increasing theretrieval accuracy. The direct retrieval of the extinction-to-backscatter (lidar) ratio (Müller et al., 2007) with ATLID(compared to the use of pre-set values in the CALIPSO re-trieval; Kim et al., 2018) and the large difference betweenlidar ratios of aerosols (20–80 sr) and clouds (20–30 sr) arealso expected to provide better distinction between opticallythin cirrus clouds and aerosols than CALIPSO (Reverdyet al., 2015). While a similar sensitivity to aerosol load is ex-pected for ATLID and CALIOP observations during night-time, ATLID promises a better daytime sensitivity. Earth-CARE is also expected to provide better distinction betweenoptically thin clouds and aerosols than CALIPSO (Reverdyet al., 2015). Airborne measurements have shown that furtherutilising HSRL at more than one wavelength (extending be-yond ATLID) would provide substantial additional informa-tion content for retrieving vertically resolved aerosol param-eters, especially when combined with polarimeter measure-ments (Burton et al., 2016). From the passive-remote-sensingperspective, promising results have been obtained for re-trievals of aerosol vertical information from near-ultravioletpolarimetry (Wu et al., 2016), although the quality degradesfor small aerosol concentrations. Passive observations withhigh spectral resolution within the oxygen A absorption bandaround 760 nm can also be used to infer aerosol layer height(Hollstein and Fischer, 2014; Geddes and Bösch, 2015).In particular, an operational aerosol layer height productis now available from the Tropospheric Monitoring Instru-ment (TROPOMI) flown on the Sentinel-5p mission (Sanderset al., 2015). Also, a recent study presents promising re-sults based on Orbiting Carbon Observatory 2 (OCO-2) ob-servations (Zeng et al., 2020). In particular, a combinationof such approaches, e.g. passive polarimetry and active lidarobservations (Stamnes et al., 2018) or multi-angle polarime-try and oxygen A band observations as planned for NASA’sPlankton, Aerosol, Cloud, ocean Ecosystem (PACE) mission(Remer et al., 2019), shows potential. Retrievals could alsocombine observations and model adjoints to constrain below-cloud aerosol number, which is directly relevant for aerosol–cloud interactions (Saide et al., 2012).

    In summary, the lack of vertical co-location between re-trieved CCN proxy and clouds leads to an underestimate inNd–CCN sensitivity (Costantino and Bréon, 2010). Modelstudies suggest that this bias may be approximately cancelledby a corresponding bias in the anthropogenic component of

    the cloud base CCN (Gryspeerdt et al., 2017). However, theextent of this cancellation in current observational studies isunknown and requires further investigation. For an accurateestimation of β, the use of lidar retrievals seems to be thebest way forward, while additional information on the ver-tical distribution of aerosol can also be gained from presentand upcoming passive satellite instruments.

    2.2 Horizontal co-location

    In studies examining β from satellite data, spatial aggregatesare considered (i.e. β as in Eq. 4), in which the aerosol re-trievals in the cloud-free pixels are averaged at a coarse res-olution (such as 1◦) and taken to define the relation with Ndretrievals in the same grid box (Quaas et al., 2008). Thisassumes that the aerosol population is horizontally homo-geneous at such large scales. According to Anderson et al.(2003), this is often the case. It has been confirmed from air-craft data for the stratocumulus cases investigated by Shi-nozuka et al. (2020). However, CCN is consumed whendroplets activate, and aerosol is scavenged when clouds pre-cipitate. Hence, the assumption of aerosol concentration hor-izontal homogeneity is questionable, at least in precipitatingclouds.

    It is the aerosol in air masses before cloud particlesform that is relevant to compute the aerosol impact on Nd(Gryspeerdt et al., 2015). In one of the early aerosol–cloudinteraction studies from satellite data Bréon et al. (2002) usedtrajectories to identify cloudless situations in which aerosolretrievals were possible for air masses that later formedclouds. This is a promising solution but it requires muchmore effort than the simpler co-location assumptions. It alsorequires reliable, high-resolution information about atmo-spheric trajectories. Another complication is that the forma-tion rate of secondary aerosol is enhanced by aqueous phasereactions, potentially enhancing aerosol concentrations in thevicinity of clouds (Jeong and Li, 2010). Such trajectory ap-proaches are particularly useful when they exploit the hightemporal resolution that is available from geostationary satel-lites. Aerosol retrievals from geostationary satellites may becombined using trajectory modelling to link these to cloudsthat form in these air masses (Kikuchi et al., 2018), or alsothe aerosol retrieval from a polar orbiter could be related toclouds retrieved from geostationary satellites that form in thesame air masses (Christensen et al., 2020).

    Altogether, the lack of horizontal co-location may implysomewhat too low β due to the potential de-correlation ofCCN concentrations and Nd in situations with spatially het-erogeneous aerosol. The consideration of backward trajec-tory analysis seems the best option to address the issue sincethere is no solution yet to retrieve aerosols below or withinclouds from satellite.

    Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020

  • J. Quaas et al.: Twomey from satellite 15085

    2.3 Hygroscopic growth of aerosol particles

    The extinction of solar radiation by aerosol particles is astrong function of the hygroscopic growth of the particles.Haze particles attenuate much more sunlight compared tothe same aerosol particle ensemble in dry conditions. AODis thus heavily influenced by the variability of relative hu-midity. The light extinction caused by dry particles (at rel-ative humidities below 30 %) is much better correlated toCCN concentrations than the extinction of particles at am-bient relative humidity (Shinozuka et al., 2015). Liu and Li(2018) showed that using total AOD compared to dry AODas a CCN proxy when estimating β from measurements atdifferent Atmospheric Radiation Measurements (ARM) sitesresulted in a 23 % underestimate. A way forward is to ap-ply parameterisations in terms of retrievals of relative hu-midity to account for the aerosol swelling. These need infor-mation about aerosol hygroscopicity and relative humidityat the appropriate scale. Hygroscopicity information couldrely on the kappa-Köhler parameterisation approach (Pettersand Kreidenweis, 2007; Pringle et al., 2010), and a param-eterisation of small-scale to mesoscale humidity variabilitycould make use of approaches exploited in general circula-tion models (GCMs) (Quaas, 2012; Petersik et al., 2018). An-other alternative would be to retrieve the amount of aerosolwater, making use of the real part of the refractive index(Schuster et al., 2009). This would allow the translation ofthe size distribution of humidified aerosol particles to thecorresponding dry size distribution. In the near future, accu-rate refractive index retrievals are expected from polarime-ters such as the SPEXone instrument on the NASA PACEmission (Hasekamp et al., 2019b; Werdell et al., 2019), to belaunched in 2022.

    Summarising, using AOD as a proxy for CCN resultsin low-biased estimates of β due to aerosol swelling. Ap-proaches to parameterise the dry aerosol properties on thebasis of the humidified one can help alleviate the problem.

    2.4 Approaches using aerosol index, column-CCN,reanalysis or cloud base updraught

    The aerosol index (AI1) is defined as the product of AODand the Ångström exponent (Deuzé et al., 2001). This latterquantity is the slope of the spectral variation in AOD and istypically larger for smaller particles (Ångström, 1929). AIis more weighted towards smaller particles, which makesit better suited as a proxy for CCN concentration at typi-cal supersaturations than AOD. For log-normal size distribu-tions, AI is approximately proportional to the column aerosolnumber concentration (Nakajima et al., 2001). Studies us-ing models concluded that AI is a better predictor for CCN

    1The difference in the measured radiance in the ultraviolet spec-tral range from a purely Rayleigh-scattering atmosphere is alsocalled the UV-AI (Torres et al., 1998), but the UV-AI is differentfrom the AI as used in this review.

    (Stier, 2016) and that AI–Nd relationships are better suited topredict 1Nd, ant than AOD–Nd relationships (Penner et al.,2011; Gryspeerdt et al., 2017). However, retrievals of theÅngström exponent, and thus of AI, over land are not re-ported in operational products such as the MODIS dark tar-get algorithm and are in general not as reliable as they areover ocean (Lee and Chung, 2013; Sayer et al., 2013).

    Further refining this idea, Hasekamp et al. (2019a) aimedto retrieve the column CCN concentrations over oceans. Theanalysis of polarimetric observations allowed us to accountfor some aspects of the aerosol particle size distribution, andfor particle sphericity, which is related to particle hygroscop-icity. This column-CCN retrieval implied larger β, increasingthe resulting RFaci by almost 50 %. It is an example of howadditional information from polarimetry is useful for study-ing the CCN-to-Nd relationship.

    However, neither the approach of Hasekamp et al. (2019a)nor the use of AI overcomes the problem of lack of hori-zontal and vertical coincidence of CCN and Nd retrievals.An option to overcome this problem is to make use of ad-ditional model information. Satellite-retrieved AOD is as-similated into aerosol models, e.g. in the Copernicus Atmo-sphere Monitoring Service (CAMS, Benedetti et al., 2009;Inness et al., 2019) or the Modern-Era Retrospective Anal-ysis for Research and Applications (version 2; MERRA-2Gelaro et al., 2017). The model predictions are applied toobtain aerosol information beneath clouds. Such aerosol re-analysis information has been used for assessing RFaci inseveral studies (Bellouin et al., 2013; McCoy et al., 2017;Bellouin et al., 2020a). However, assessing the validity ofmodel results requires extensive and rigorous evaluation, es-pecially for coarsely resolved models with regard to aerosolscavenging below clouds. For this, independent data are re-quired such as from ground-based observations or satelliteobservations from sensors other than those that are assimi-lated.

    Yet another solution initially proposed by Feingold et al.(1998) and applied to satellite retrievals by Rosenfeld et al.(2016) is to parameterise the cloud base updraught, w, on thebasis of cloud retrievals, rather than to retrieve the aerosol.For convective clouds, Zheng et al. (2015) suggested thatw scales with cloud base altitude, which can be retrievedfrom satellites. For stratocumulus clouds, Zheng et al. (2016)proposed that updraught is a function of cloud-top radiativecooling, and that this can be computed by radiative trans-fer modelling on the basis of cloud quantities retrieved frompassive sensors and thermodynamic profiles from meteoro-logical re-analyses. The retrieved profiles of re together withderivations of supersaturation as a function of w and Nd(Rosenfeld et al., 2016) then allow the parameterisation ofthe CCN concentration at any given supersaturation. This ap-proach does not suffer from the problem of a lower detec-tion limit. However, it has not yet been used to quantify theTwomey effect.

    https://doi.org/10.5194/acp-20-15079-2020 Atmos. Chem. Phys., 20, 15079–15099, 2020

  • 15086 J. Quaas et al.: Twomey from satellite

    Concluding, all four approaches alleviate many problemsencountered when using AOD. An ideal solution may be thecombination of several of these by also assimilating, in ad-dition to AOD, polarimetric satellite observations, as well aslidar measurements, into the analysis of the atmospheric statein high-resolution models.

    3 Remote sensing of cloud droplet concentrations

    The problem of the remotely sensed Nd as used to estimateβ has three different facets to it, which will be discussed inthis section, namely the following.

    - Consideration of re rather than Nd in aerosol–cloud in-teraction studies. In many studies, the droplet effectiveradius, re, is used, and the datasets are stratified with re-spect to L in order to estimate β. This is very difficultto perform adequately and leads to biases.

    - Biases in the retrieved Nd . For the assessment of sen-sitivity, systematic (rather than random) errors in re-trievedNd are relevant. Also,Nd is not retrieved in stan-dard operational procedures, so that inconsistencies be-tween the retrieval of standard components and in thecomputation of Nd on the basis of retrievals can lead toadditional errors.

    - Relationship of Nd formed at activation with retrievedand radiation-relevant Nd, top. Retrieved Nd, top refersto the drop concentration within the top one to two opti-cal depths of the clouds, and it is Nd, top that is relevantfor determining the cloud radiative effect. Nd sink pro-cesses such as coagulation imply that Nd, top is smallerthan the one resulting from activation above cloud base,Nd, base.

    Nd is vertically constant for single-layer, purely liquid-waterclouds with (i) a vertically homogeneous droplet size spec-trum, (ii) for adiabatically stratified clouds or (iii) for sub-adiabatic clouds in which mixing is homogeneous. However,in many situations, precipitation formation or entrainmentcan lead to reduction of Nd above cloud base. In such sit-uations, it is Nd, top that is relevant to determine the cloudradiative effect (cloud albedo in Eq. 2). Building on Eq. (4)thus gives

    1Nd, top,ant

    =dNd, topdNd, base

    ·

    ∞∫w=−∞

    ∂Nd, base

    ∂ lna

    ∣∣∣∣w

    P(w)P(a)dw

    ·1 lnaant = β̂ ·1 lnaant . (5)

    When estimating β as a regression coefficient from, for ex-ample, satellite-retrieved Nd and a proxy for CCN such asAOD, it is thus this β̂ that is inferred.

    3.1 Considering re rather than Nd

    Many past studies have used operationally retrieved re ratherthan Nd in aerosol–cloud interaction studies. However, re isa function of bothNd and L. This introduces the requirementfor stratifying the data with respect to L in order to estimateβ̂. To further complicate matters, Nd and L have been foundto be correlated (e.g. Michibata et al., 2016; Gryspeerdt et al.,2019). A precise estimation of β̂ is thus only possible fora large amount of data combined with suitable binning byL. Errors in this approach that are related to a lack of dataincrease at aggregated scales (McComiskey and Feingold,2012). Using derived Nd is therefore preferable to avoid un-necessary complications.

    3.2 Biases in the Nd retrieval

    Satellite retrievals of Nd were extensively reviewed byGrosvenor et al. (2018). Since Nd currently is not retrievedby operational algorithms and new developments to retrieveNd (e.g. from polarimetry) are still in their infancy, the mostfrequently used method is to infer Nd from retrieved re andcloud optical depth, τc, using the relationship

    Nd = γ · τ12

    c · r−

    52

    e , (6)

    where γ ≈ 1.37× 10−5 m−0.5 is a parameter provided as aconstant here but more realistically depending on cloud basetemperature and pressure, the adiabatic fraction, and the dropsize distribution breadth (Boers et al., 2006; Quaas et al.,2006; Grosvenor et al., 2018). The relationship in Eq. (6) as-sumes that clouds are adiabatic or nearly adiabatic (i.e. adia-batic clouds or sub-adiabatic clouds with homogeneous mix-ing only; Brenguier et al., 2000). The most common methoduses a bispectral approach to retrieve re and τc (Nakajimaand King, 1990). Various error sources lead to an overall re-trieval error forNd (Grosvenor et al., 2018; Wolf et al., 2019).As can be deduced form Eq. (6), the most important contri-butions are from retrieval errors in re. Other error sourcesare the uncertainty in sub-adiabatic factor, the cloud modelused in the retrieval, and the droplet size distribution width.Satellite retrievals of the vertical profile of cloud droplet sizemay help to improve the retrieval (Chang and Li, 2002; Chenet al., 2008). Grosvenor et al. (2018) identified biases of re-trievedNd in particular for broken cloud regimes and at largesolar zenith angles. In stratocumulus, it was suggested thatthe retrieval yields the most trustworthy results when con-sidering only the brightest pixels (Zhu et al., 2018). For theideal case of homogeneous, low-latitude stratiform clouds,relative errors in the Nd retrieval at pixel scale are quanti-fied as 78 % (Grosvenor et al., 2018). In such cases, the errorwas assumed to be random. However, systematic errors oc-cur in particular in broken cloud regimes and for large solarzenith angles, leading to an underestimation (broken cloudi-ness) and overestimation (large solar zenith angles), respec-tively, ofNd. Painemal et al. (2020) addressed theNd bias for

    Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020

  • J. Quaas et al.: Twomey from satellite 15087

    broken clouds by only samplingNd retrieved for large clouds(larger than 5 km× 5 km) to find that the relation betweenNdand aerosols is substantially enhanced.

    For improvements in estimates of Nd, it would be bene-ficial to formulate a retrieval in terms of Nd directly ratherthan in terms of re and τc. It is also possible to reduce un-certainties in retrievals of re and τc, or to reduce uncertain-ties related to assumptions of the vertical structure of thecloud and particle size distribution shape. Approaches toquantify and partly correct for retrieval biases as discussedin Grosvenor et al. (2018) include accounting for cloud het-erogeneity by using those channels in passive imagers thatprovide spatial resolution that exceeds the one at which thestandard retrieval products are provided (Zhang et al., 2016).The combination of passive observations with radar may fur-ther improve the retrieval (Posselt et al., 2017). Substan-tially more accurate retrievals of re and additional relevantinformation about droplet size distributions may also comefrom multi-angular polarimetric measurements (Alexandrovet al., 2012a, b; Shang et al., 2019), which will be possi-ble from orbit at pixel level from the Hyper-Angular Rain-bow Polarimeter-2 (HARP-2) on the NASA PACE mission(Martins et al., 2018; McBride et al., 2019). Polarimetricretrievals allow the inference of the spectral width or gen-eral shape of the droplet size distribution at cloud top (Huet al., 2007). This approach is not substantially sensitive tosub-pixel cloudiness, mixed-phase conditions and 3D radia-tive effects (Alexandrov et al., 2012b). The sensitivity of de-rived Nd to uncertainties in re from polarimetric retrievalsmay further be reduced by additionally inferring cloud phys-ical thickness. In this case, Nd can be inferred to be linearin τc and inversely linear in geometrical thickness and meandroplet extinction cross section at cloud top (Sinclair et al.,2019). The geometrical thickness may also be inferred fromtotal and/or polarised reflectances measured in oxygen or wa-ter vapour absorption bands (Desmons et al., 2013; Sanghaviet al., 2015; Richardson et al., 2019; Sinclair et al., 2019)or by retrieving cloud base using lidar (Mülmenstädt et al.,2018) or using multi-angle observations (Böhm et al., 2019).When exploiting passive observations together with lidar,Ndat cloud top can be robustly inferred as the ratio of in-cloudextinction (lidar) and extinction cross section (passive). Aslightly less direct approach using depolarisation to estimateextinction and effective radius to estimate extinction crosssection has been presented by Hu et al. (2007).

    3.3 Relationship between Nd formed at CCN activationand retrieved radiation-relevant Nd

    In stratiform clouds, droplets form in updraughts near cloudbase which is where Nd most closely relates to CCN. Inconvective clouds, updraught in some cases increases withheight above cloud base. Hence, additional CCN may acti-vate above cloud base and lead to vertically increasing Nd inthe lower third of the cloud with a decrease further up (Endo

    et al., 2015). However, in most cumulus clouds, and in strat-iform clouds, Nd is found to be largest at cloud base and toslightly decrease above it (Jiang et al., 2008; Small et al.,2009; vanZanten et al., 2011). In the approach discussed byGrosvenor et al. (2018), the retrieved Nd is representative ofthe cloud-top reflectance, and thus the relevant proxy for theNd that matters for cloud albedo and RFaci (Platnick, 2000).To which extent the microphysical structure of lower partsof a cloud exactly impacts radiation (weighting function)depends on the multiple scattering and thus on the verticalstructure of Nd itself (Platnick, 2000; Krisna et al., 2018).For vertically constant Nd, the retrieved Nd represents thedroplet concentration formed by CCN activation. However,there are Nd sinks, in particular due to collision and coales-cence (in liquid clouds, the autoconversion and accretion, or“warm rain” processes) that lead to droplet depletion. Wood(2006) demonstrated that the depletion is exponential in pre-cipitation rate and estimated a loss in Nd of 100 cm−3 d−1

    for precipitation rates of 1 mm d−1. There may also be lat-eral and vertical mixing (of heterogeneous type; Lehmannet al., 2009) of cloud air with environmental cloud-free airthat can lead to the full evaporation of droplets. In both sinksfor Nd, the one due to precipitation formation and the onedue to mixing, the retrievedNd is expected to be smaller thanthe Nd formed at activation of CCN. In an aged cloud, how-ever, updraughts may have decayed such that no additionaldroplets are formed, while existing droplets persist, or maybe advected from elsewhere. Also, in case they are very large,raindrops may break up into droplets, in which case Nd is in-creased. Arguably, it is the right choice to relate the retrievedNd, as the radiation-relevant one, to CCN, i.e. to use β̂, whencomputing the Nd-to-CCN sensitivity with the aim to con-strain RFaci.

    Cloud-resolving models are a good tool to investigatethese interpretations (McComiskey and Feingold, 2012). Fig-ure 1 shows an analysis of a large-domain large-eddy sim-ulation with the ICON-LEM model (Heinze et al., 2017;Costa-Surós et al., 2020). CCN concentrations in these simu-lations are relaxed towards pre-computed spatially and tem-porally varying fields and are consumed at activation. In the22 million grid columns, the droplet concentration at cloudtop (what is retrieved from satellites) is compared to themaximum droplet concentration (approximately the concen-tration of activated CCN divided by formed droplets). Thisdemonstrates that there is a link between the droplet concen-tration formed at activation and Nd determining the cloud ra-diative effect at its top. These two quantities correlate ratherwell in the joint histogram, though that link is far from one toone. The second plot (Fig. 1b) assesses the possibility to in-fer cloud-top Nd from cloud-top re and τc (Grosvenor et al.,2018). For this, the MODIS simulator (Pincus et al., 2012)that is part of the Cloud Feedback Model IntercomparisonProject (CFMIP) Observational Simulator Package (COSP;Bodas-Salcedo et al., 2011) is applied to the model outputto compute cloud-top re and τc. From these, Nd is computed

    https://doi.org/10.5194/acp-20-15079-2020 Atmos. Chem. Phys., 20, 15079–15099, 2020

  • 15088 J. Quaas et al.: Twomey from satellite

    as in Eq. (6). This approach mimics the satellite retrieval butassumes no retrieval errors; i.e. the comparison is a lowerbound on the accuracy of the retrievedNd in representing theactual Nd at cloud top. There is a meaningful co-variationin the two quantities, but it is far from perfect. In particu-lar, there is a systematic overestimation of Nd in the retrievalapproach, especially at low Nd. The relative error even is afunction of Nd, with larger relative errors at low Nd.

    In conclusion, the fact that cloud-topNd is in general lowerthan Nd at activation height implies that β̂ is indeed some-what smaller than unity. This is not a problem but rather a de-sired analysis result when studying the Twomey effect. How-ever, Nd obtained from retrieval products is biased high forlow values of Nd, top. This relative error, which is a functionof Nd, implies that the regression between satellite-derivedNd and CCN yields a sensitivity that is too weak.

    4 Cloud-regime dependence

    Aerosol–cloud interactions depend on cloud regime (Stevensand Feingold, 2009; Mülmenstädt and Feingold, 2018).When it comes to RFaci, there are three reasons for this:(i) the radiative sensitivity (Oreopoulos and Platnick, 2008;Alterskjær et al., 2012), i.e. the first two terms on the right-hand-side of Eq. (2) (in particular the sensitivity expressedin Eq. 1); (ii) the updraught dependence of β̂; and (iii) thedependence of the relation of cloud-top to cloud base Ndto characteristics of turbulence and rain. The latter two areof interest here. “Cloud regime” thus here means, a clus-

    ter of clouds with similar P(w) and similar dNd, topdNd, base

    inEq. (5). When considering CCN at a certain supersaturationlevel, β̂ is larger at larger updraught, w (MacDonald et al.,2020). Broadly, cumulus clouds have largerw than stratiformclouds. In addition, clouds over land usually have larger wthan clouds over ocean. Building on Eq. (5), this suggests aregime-based analysis expressed as

    1Nd, top,ant =dNd, top

    dlna

    ∣∣∣∣∣regime

    ·1 lnaant . (7)

    Figure 2 shows the spatial distribution of the Nd – AI regres-sion coefficient from its temporal variability within 1◦× 1◦

    grid boxes. The large spatial heterogeneity is not straightfor-ward to interpret. Some problems may be due to the lack ofaerosol retrieval sensitivity (e.g. in regions with low CCNconcentrations such as the southern oceans) or lack of ver-tical or horizontal co-incidence (e.g. in regions with het-erogeneous aerosol and large cloud coverage such as mid-latitude storm tracks). However, aspects of the geographicalheterogeneity may indeed be attributable to physical and rel-evant reasons. However, it is difficult to determine any at-tributable factors in the spatial and cloud-regime variationsin β̂ (Gryspeerdt and Stier, 2012) before retrieval errors areremedied.

    In precipitating situations, the two-way interactions canlead to large challenges in determining the β̂ term (Ekmanet al., 2011). Precipitation scavenges aerosol and, in certainsituations, the interplay between aerosol, droplet concentra-tions and precipitation determines both aerosol and dropletconcentrations. This may yield bifurcations between situa-tions with large Nd in which no drizzle forms and very lowNd and cloud dissolution when precipitation forms (e.g. Ya-maguchi et al., 2017). In such situations, it is particularlychallenging to identify the Nd–CCN concentration sensitiv-ity.

    5 Aggregation scale

    The impact of aggregation scale on estimates of β has beendiscussed in detail by McComiskey and Feingold (2012).Their key conclusion is that at scales larger than the cloudvariability scale of about 1 to 10 km, aerosol and cloud databecome de-correlated so that the diagnosed β becomes lessand less representative for individual cloud parcels. In turn,Sekiguchi et al. (2003) computed β̂ for different aggregationscales and demonstrated that it actually increases with largerscales. An analysis of spatio-temporal vs. temporal-only co-variability of Nd and AOD by Grandey and Stier (2010)found that β̂ is larger when considering spatio-temporal vari-ability over entire regions compared to only temporal vari-ability at individual 1◦× 1◦ grid boxes. These results areopposite to those expected from the process-based conclu-sions of McComiskey and Feingold (2012). A possible prob-lem in the Sekiguchi et al. (2003) study is their use of rerather than Nd, and the subsequent need to stratify by L.McComiskey and Feingold (2012) demonstrated that this ap-proach becomes more problematic with increasing aggrega-tion scale. However, their analysis suggested a low-bias inβ at coarser scales due to stratification by L. Reduced β̂ atsmall scales could occur if aerosol conditions become too ho-mogeneous to diagnose the full range of co-variability due tosmaller sample sizes at smaller scales.

    Concluding, from a process point of view, aggregationover larger scales is expected to lead to a decrease in esti-mated β̂. In turn, to study the large-scale Twomey effect, anaggregate Nd–CCN relationship is desired as it is the large-scale 1Nd, ant that matters for the radiation perturbation andbecause the anthropogenic aerosol perturbation can only beinferred at a large scale. The often adopted choice of a 1◦×1◦

    gridding is somewhat motivated by the suggestion that this isa scale at which aerosol concentrations are considered ho-mogeneous (Anderson et al., 2003) and loosely (to withina factor of about 2 in each horizontal direction; re-analysesare to closer ∼ 50 km scales, and many general circulationmodels still are as coarse as 200 km) related to the scale atwhich models infer the anthropogenic perturbation of CCN.A rigorous study on the scale dependency of β̂ and the con-sequences thereof for RFaci would be desirable.

    Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020

  • J. Quaas et al.: Twomey from satellite 15089

    Figure 1. Analysis of Nd in the “virtual reality” of a cloud-resolving simulation: droplet number concentration (cm−3) from the ICONlarge-eddy simulation (156 m horizontal resolution) over the domain of Germany for 2 May 2013 (Heinze et al., 2017), for the overpasstimes of the Terra and Aqua satellites for which the swath of the MODIS instrument covered the domain (twice around 10:30 local solartime for Terra, twice around 13:30 for Aqua) even if no actual data are used in this analysis (Costa-Surós et al., 2020). Joint histograms,normalised along the y axis as in Gryspeerdt et al. (2016) for (a) column-maximum (proxy for activated CCN) vs. cloud-top Nd (taken atτc = 1 integrated from cloud top) and (b) Nd derived from re and τc as in Grosvenor et al. (2018) vs. cloud-top Nd, where both quantitiesare computed as seen from a satellite using the Cloud Feedback Model Intercomparison project (CFMIP) Observational Simulator PackageCOSP (Bodas-Salcedo et al., 2011). The blue line is the mean in each bin for cloud-top Nd.

    Figure 2. Regression coefficients ofNd computed on the basis of retrievals of the MODerate Resolution Imaging Spectroradiometer (MODIS;Platnick et al., 2017) as in Grosvenor et al. (2018) and AI from MODIS (Levy et al., 2013) from the daily temporal variability in grid boxesof 1◦× 1◦.

    6 Quantification for the regression coefficient

    When sensitivities are approximated by linear regression co-efficients from an ordinary least-squares (OLS) line fittingmethod, rather than derived in the form of joint histograms,the problem of regression dilution arises to the extent thatthe aerosol quantity shows errors: the regression coefficientbecomes gradually smaller as the stochastic error increases(Cantrell, 2008; Pitkänen et al., 2016; Wu and Yu, 2018).Regression dilution, also known as regression attenuation, isa problem if the independent variable (x axis) in the regres-sion is subject to a statistical error. If the regression methoddoes not take the statistical error into account, which is oftenthe case (for example in OLS), the regression coefficient isalways systematically biased low. In turn, statistical error on

    the dependent variable (y axis) only causes uncertainty in theregression coefficient but no systematic bias. This is quanti-fied for the column-CCN vs. Nd sensitivity evaluated as aregression coefficient in Fig. 3. Due to the regression dilu-tion, the sensitivity decreases by factors of 2 to 3 as the errorin column CCN increases when considering relative errorsof 50 %. This can to a large extent be remedied by ignor-ing data points at low CCN concentrations from the regres-sion (Fig. 3b). However, this solution is limited to regionsnot dominated by low aerosol concentrations. Figure 3 alsoillustrates that an absolute bias in the data translates to rela-tive bias in logarithmic scale. Therefore, if no bias correctionis applied, an absolute bias in the data will cause a bias inthe sensitivity estimates. As shown by Pitkänen et al. (2016),the regression dilution in turn becomes weaker at coarser ag-

    https://doi.org/10.5194/acp-20-15079-2020 Atmos. Chem. Phys., 20, 15079–15099, 2020

  • 15090 J. Quaas et al.: Twomey from satellite

    Figure 3. Nd–column CCN sensitivity as a function of the stochastic error in column CCN (absolute additive error) in an emulated analysisas in Hasekamp et al. (2019a), for different relative (multiplicative) errors, for (a) the full range of data, including low NCCN values and(b) excluding NCCN < 107 cm−2. Hasekamp et al. (2019a) suggest a realistic error is about 0.2 ·NCCN+ 4× 106 cm−2.

    gregation scales in cases of auto-correlated data, which isthe case for aerosol concentrations. This is of relevance inthe case of both temporal and spatial aggregation. In otherwords, the systematic low-bias in the sensitivity is reduced ifdata are aggregated. This could partly explain some previousfindings of increasing sensitivity with decreasing resolution(see discussion in the previous section), in addition to the ac-tual bias due to the aggregation over a smaller scale of cloudprocesses. These considerations imply that it is necessary toeither analyse the full variability of aerosol–cloud interac-tions, e.g. in the form of joint histograms, or to account forthe regression dilution using established mathematical ap-proaches that properly consider measurement uncertainties,as discussed in Mikkonen et al. (2019), for instance.

    7 Conclusions

    The radiative forcing due to aerosol–cloud interactions, orthe Twomey effect, requires quantification based on observa-tional data, since models are associated with large uncertain-ties. At a large scale, this calls for satellite retrievals. Thereare, however, large challenges when using satellite data andthis review summarises these challenges and suggests somepotential ways forward. The key data-related question isthe sensitivity of droplet concentration, Nd, to perturbationsin the cloud-active aerosol, i.e. the cloud condensation nu-clei (CCN) concentration at or above cloud base. The mostwidely used proxy of the cloud base CCN concentration isthe aerosol optical depth (AOD), or alternatively the aerosolindex (AI), taken from cloud-free pixels in the vicinity ofthe locations of the cloud retrievals. The four main caveatswith AOD are the lack of vertical resolution, the additionalinfluence of hygroscopic swelling, the fact that the detectedaerosol might be not active as CCN nd the impossibility toretrieve it below clouds. In terms of the vertical resolution,satellite-based lidar offers help. However, current lidar re-trievals are even more constrained to large aerosol concentra-tions than passive AOD retrievals. EarthCARE’s ATLID lidarwill allow direct inference of the ratio of backscatter to ex-

    tinction, enabling greatly improved retrievals of aerosol ex-tinction profile. Adding a second wavelength with ATLID ca-pabilities and combining it with polarimetric measurementswould substantially extend vertically resolved aerosol infor-mation content. In terms of horizontal co-location, trajectorycomputations may help to identify the aerosol representativeof that affecting specific clouds. However, this requires extraeffort and reliable information about trajectories. The hygro-scopic swelling can be addressed by parameterisations thatuse retrievals and ancillary data to compute the swelling. Fur-ther relevant information is possible from polarimetric mea-surements.

    Cloud droplet number concentration, Nd, is only indi-rectly available from current operational satellite retrievals.It is generally computed from retrieved cloud-top droplet ef-fective radius, re, and cloud optical thickness, τc, leadingto substantial biases in comparison to the cloud-top dropletnumber concentration, especially in inhomogeneous, brokenand/or precipitating cloud regimes. Sink processes for Ndand variability due to atmospheric dynamics, including tur-bulent mixing, imply that the radiatively relevant cloud-topNd relates imperfectly to the Nd formed by CCN activation.In addition, at a given CCN concentration, the updraughtvariability also leads to sensitivities of Nd to CCN that aremuch less than 1. These latter two facts are not problematicwhen assessing the Nd to aerosol sensitivity from data forthe estimation of the Twomey effect. In fact, it is desirable toquantify at a large scale the net impact of aerosol perturba-tions of the (radiatively relevant) cloud-top Nd that accountsfor updraught and Nd sink variability. However, it is neces-sary to operationally retrieve Nd, rather than to indirectlycompute it from re and τc retrievals. It is also necessary toimprove these retrievals in particular for low droplet concen-trations and broken cloud conditions. In addition, these re-trievals should take into account additional information, e.g.about the onset of drizzle.

    Regression dilution influences the statistically inferredsensitivity as a result of stochastic retrieval errors in CCNconcentration. On the one hand, at aggregate scales, this

    Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020

  • J. Quaas et al.: Twomey from satellite 15091

    problem becomes less relevant due to the autocorrelationof the aerosol concentrations. The relationship between Nd,which varies at cloud-dynamics scales, and CCN proxies be-comes weaker at aggregate scales. Relative retrieval errors inNd that depend on actual Nd (with larger high-biases at lowtrue Nd) lead to a further reduction in the estimated sensi-tivity. It is thus necessary to account for the impact of CCNerrors in the statistics and to optimise the resolution of Ndand CCN retrievals towards cloud-scale resolutions.

    The recent study by Hasekamp et al. (2019a) made useof polarimetric satellite measurements to suggest a global-ocean average Nd-to-CCN sensitivity of 0.66. This, com-bined with anthropogenic column-CCN concentrations andradiative sensitivities, translates into a global Twomey ef-fect of −1.1 W m−2. The net effect of the remaining prob-lems laid out above suggests that this likely is still too lowan estimate for the Nd–CCN sensitivity, implying a strongerTwomey effect. However, the estimate is in line with an inde-pendent observation-based estimate of McCoy et al. (2020)that used differences in Nd between pristine and polluted re-gions in combination with GCM results as an emergent con-straint. In any case, it is desirable to add the extra steps toimprove the quantification supported by data for process un-derstanding as well as for evaluating and improving climatemodels.

    In situ and ground-based observations, as well as analysisof cloud-resolving dynamical models, may be a path forwardfor the evaluation of critical aspects in the satellite-basedanalysis. Important steps would be the quantification of up-draught PDFs for different cloud regimes and the assessmentof horizontal homogeneity of aerosol concentrations.

    Data availability. The data used in Fig. 1 are the simulation data asdescribed in Costa-Surós et al. (2020, https://doi.org/10.5194/acp-20-5657-2020) and available upon request due to the largeamount of data; it is securely saved in tape archives at theDeutsches Klimarechenzentrum (DKRZ), which will be accessi-ble for 10 years. The MODIS data used in Fig. 2 was down-loaded from the Level-1 and Atmosphere Archive & DistributionSystem (LAADS) Distributed Active Archive Center (DAAC), lo-cated in the Goddard Space Flight Center in Greenbelt, Maryland(https://ladsweb.nascom.nasa.gov/, last access: 19 November 2020,LAADS DAAC, 2020). The data used for Fig. 3 are archived andavailable at https://cera-www.dkrz.de/WDCC/ui/cerasearch/entry?acronym=DKRZ_LTA_1002_ds00001 (last access: 19 November2020, Hasekamp and Quaas, 2020).

    Author contributions. JQ led the writing of the manuscript with sig-nificant contributions from all authors.

    Competing interests. The authors declare there are no competinginterests.

    Acknowledgements. The work of Johannes Quaas, Annica M. L.Ekman, Athanasios Nenes and Philip Stier was supported by theEuropean Union via its Horizon 2020 project FORCeS. The work ofAthanasios Nenes was further supported by the European ResearchCouncil via the project PyroTRACH. Philip Stier was also sup-ported by the European Research Council (ERC) project constRain-ing the EffeCts of Aerosols on Precipitation (RECAP). This revieworiginated from discussions at the 2019 Nanjing workshop of theAerosols-clouds-precipitation and climate (ACPC) initiative (http://acpcinitiative.org/, last access: 13 November 2020) and benefitedfrom discussions within the group “Study of aerosol–cloud inter-actions based on satellite observations of the terrestrial underlyingsurface–atmosphere system: a new frontier of atmospheric science”,hosted by the International Space Science Institute (ISSI). We thankthe German Climate Computing Centre (Deutsches Klimarechen-zentrum, DRKZ) and the German Federal Ministry of Educationand Research (BMBF) within the framework programme “Researchfor Sustainable Development (FONA)”, https://www.fona.de/ (lastaccess: 13 November 2020) for making the ICON-LEM simula-tions available. Po-Lun Ma and Johannes Mülmenstädt were sup-ported by the U.S. Department of Energy, Office of Science, Of-fice of Biological and Environmental Research, Earth System Mod-eling program. Johannes Quaas is grateful to the NASA GoddardInstitute for Space Studies, New York, for hospitality during a re-search stay. We thank Andrew Ackerman and Patrick Chuang forconstructive discussions. We thank the three anonymous reviewersand David Painemal for helpful comments on the earlier version ofthe manuscript. We acknowledge support from Leipzig Universityfor Open Access Publishing.

    Financial support. This research has been supported by the Euro-pean Union’s Horizon 2020 research and innovation programmevia H2020-EU.3.5.1. – Fighting and adapting to climate change(FORCeS (grant agreement no. 821205)), the European ResearchCouncil (ERC) via H2020-EU.1.1. – Excellent Science program(PyroTRACH (grant agreement no. 726165) and RECAP (grantagreement no. 724602)), the U.S. Department of Energy, Office ofScience, Office of Biological and Environmental Research, EarthSystem Modeling program by the “Enabling Aerosol-cloud interac-tions at GLobal convection-permitting scalES (EAGLES) (projectno. 74358), and the Pacific Northwest National Laboratory is oper-ated for the U.S. Department of Energy by Battelle Memorial Insti-tute (contract no. DE-AC05-76RL01830).

    Review statement. This paper was edited by Frank Dentener andreviewed by three anonymous referees.

    References

    Ackerman, A., Toon, O., Taylor, J., Johnson, D., Hobbs,P., and Ferek, R.: Effects of aerosols on cloud albedo:Evaluation of Twomey’s parameterization of cloud sus-ceptibility using measurements of ship tracks, J. At-mos. Sci., 57, 2684–2695, https://doi.org/10.1175/1520-0469(2000)0572.0.CO;2, 2000.

    https://doi.org/10.5194/acp-20-15079-2020 Atmos. Chem. Phys., 20, 15079–15099, 2020

    https://doi.org/10.5194/acp-20-5657-2020https://doi.org/10.5194/acp-20-5657-2020https://ladsweb.nascom.nasa.gov/https://cera-www.dkrz.de/WDCC/ui/cerasearch/entry?acronym=DKRZ_LTA_1002_ds00001https://cera-www.dkrz.de/WDCC/ui/cerasearch/entry?acronym=DKRZ_LTA_1002_ds00001http://acpcinitiative.org/http://acpcinitiative.org/https://www.fona.de/https://doi.org/10.1175/1520-0469(2000)0572.0.CO;2https://doi.org/10.1175/1520-0469(2000)0572.0.CO;2

  • 15092 J. Quaas et al.: Twomey from satellite

    Ackerman, A., Kirkpatrick, M., Stevens, D., and Toon, O.:The impact of humidity above stratiform clouds on in-direct aerosol climate forcing, Nature, 432, 1014–1017,https://doi.org/10.1038/nature03174, 2004.

    Aitken, J.: On dust, fogs, and clouds, Proceedings of the Royal So-ciety of Edinburgh, 11, 122–126, 1880.

    Albrecht, B.: Aerosols, cloud microphysics, andfractional cloudiness, Science, 245, 1227–1230,https://doi.org/10.1126/science.245.4923.1227, 1989.

    Alexandrov, M., Cairns, B., and Mishchenko, M.: RainbowFourier transform, J. Quant. Spectrosc. Ra., 113, 2521–2535,https://doi.org/10.1016/j.jqsrt.2012.03.025, 2012a.

    Alexandrov, M. D., Cairns, B., Emde, C., Ackerman, A. S., andvan Diedenhoven, B.: Accuracy assessments of cloud dropletsize retrievals from polarized reflectance measurements by theresearch scanning polarimeter, Remote Sens. Environ., 125, 92–111, 2012b.

    Alterskjær, K., Kristjánsson, J. E., and Seland, Ø.: Sensitivityto deliberate sea salt seeding of marine clouds – observationsand model simulations, Atmos. Chem. Phys., 12, 2795–2807,https://doi.org/10.5194/acp-12-2795-2012, 2012.

    Anderson, T. L., Charlson, R. J., Winker, D. M., Ogren, J. A.,and Holmen, K.: Mesoscale variations of tropospheric aerosols,J. Atmos. Sci., 60, 119–136, https://doi.org/10.1175/1520-0469(2003)0602.0.CO;2, 2003.

    Andreae, M. O.: Correlation between cloud condensation nu-clei concentration and aerosol optical thickness in remoteand polluted regions, Atmos. Chem. Phys., 9, 543–556,https://doi.org/10.5194/acp-9-543-2009, 2009.

    Ångström, A.: On the Atmospheric Transmission of Sun Radia-tion and on Dust in the Air, Geografiska Annaler, 11, 156–166,https://doi.org/10.1080/20014422.1929.11880498, 1929.

    Bellouin, N., Boucher, O., Haywood, J., and Reddy,M. S.: Global estimate of aerosol direct radiative forc-ing from satellite measurements, Nature, 438, 1138–1141,https://doi.org/10.1038/nature04348, 2005.

    Bellouin, N., Quaas, J., Morcrette, J.-J., and Boucher, O.: Estimatesof aerosol radiative forcing from the MACC re-analysis, At-mos. Chem. Phys., 13, 2045–2062, https://doi.org/10.5194/acp-13-2045-2013, 2013.

    Bellouin, N., Davies, W., Shine, K. P., Quaas, J., Mülmenstädt, J.,Forster, P. M., Smith, C., Lee, L., Regayre, L., Brasseur, G., Su-darchikova, N., Bouarar, I., Boucher, O., and Myhre, G.: Radia-tive forcing of climate change from the Copernicus reanalysis ofatmospheric composition, Earth Syst. Sci. Data, 12, 1649–1677,https://doi.org/10.5194/essd-12-1649-2020, 2020a.

    Bellouin, N., Quaas, J., Gryspeerdt, E., Kinne, S., Stier, P., Watson-Parris, D., Boucher, O., Carslaw, K., Christensen, M., Da-niau, A.-L., Dufresne, J.-L., Feingold, G., Fiedler, S., Forster,P., Gettelman, A., Haywood, J. M., Malavelle, F., Lohmann,U., Mauritsen, T., McCoy, D., Myhre, G., Mülmenstädt, J.,Neubauer, D., Possner, A., Rugenstein, M., Sato, Y., Schulz, M.,Schwartz, S. E., Sourdeval, O., Storelvmo, T., Toll, V., Winker,D., and Stevens, B.: Bounding global aerosol radiative forc-ing of climate change, Rev. Geophys., 58, e2019RG000660,https://doi.org/10.1029/2019RG000660, 2020b.

    Benedetti, A., Morcrette, J.-J., Boucher, O., Dethof, A., Engelen,R. J., Fisher, M., Flentje, H., Huneeus, N., Jones, L., Kaiser,J. W., Kinne, S., Mangold, A., Razinger, M., Simmons, A. J.,

    and Suttie, M.: Aerosol analysis and forecast in the EuropeanCentre for Medium-Range Weather Forecasts Integrated ForecastSystem: 2. Data assimilation, J. Geophys. Res., 114, D13205,https://doi.org/10.1029/2008JD011115, 2009.

    Bennartz, R., Fan, J., Rausch, J., Leung, L. R., and Heidinger, A. K.:Pollution from China increases cloud droplet number, suppressesrain over the East China Sea, Geophys. Res. Lett., 38, L09704,https://doi.org/10.1029/2011GL047235, 2011.

    Bodas-Salcedo, A., Webb, M. J., Bony, S., Chepfer, H., Dufresne,J.-L., Klein, S. A., Zhang, Y., Marchand, R., Haynes, J. M., Pin-cus, R., and John, V. O.: COSP: Satellite simulation softwarefor model assessment, B. Am. Meteorol. Soc., 92, 1023–1043,https://doi.org/10.1175/2011BAMS2856.1, 2011.

    Boers, R., Acarreta, J. R., and Gras, J. L.: Satellite monitoringof the first indirect aerosol effect: Retrieval of the droplet con-centration of water clouds, J. Geophys. Res., 111, D22208,https://doi.org/10.1029/2005JD006838, 2006.

    Böhm, C., Sourdeval, O., Mülmenstädt, J., Quaas, J., and Crewell,S.: Cloud base height retrieval from multi-angle satellite data, At-mos. Meas. Tech., 12, 1841–1860, https://doi.org/10.5194/amt-12-1841-2019, 2019.

    Boucher, O. and Lohmann, U.: The sulfate-CCN-cloud albedoeffect, Tellus B, 47, 281–300, https://doi.org/10.1034/j.1600-0889.47.issue3.1.x, 1995.

    Boucher, O., Randall, D., Artaxo, P., Bretherton, C., Feingold, G.,Forster, P., Kerminen, V.-M., Kondo, Y., Liao, H., Lohmann, U.,Rasch, P., Satheesh, S., Sherwood, S., Stevens, B., and Zhang,X.: Clouds and Aerosols, in: Climate Change 2013: The Phys-ical Science Basis. Contribution of Working Group I to theFifth Assessment Report of the Intergovernmental Panel on Cli-mate Change, edited by: Stocker, T., Qin, D., Plattner, G.-K.,Tignor, M., Allen, S., Boschung, J., Nauels, A., Xia, Y., Bex,V., and Midgley, P., Cambridge University Press, Cambridge,United Kingdom and New York, NY, USA, chap. 7, 571–658,https://doi.org/10.1017/CBO9781107415324.016, 2013.

    Bougiatioti, A., Argyrouli, A., Solomos, S., Vratolis, S., Eleft-heriadis, K., Papayannis, A., and Nenes, A.: CCN activ-ity, variability and influence on droplet formation duringthe HygrA-CD campaign in Athens, Atmosphere, 8, 108,https://doi.org/10.3390/atmos8060108, 2017.

    Bougiatioti, A., Nenes, A., Lin, J. J., Brock, C. A., de Gouw, J.A., Liao, J., Middlebrook, A. M., and Welti, A.: Drivers ofcloud droplet number variability in the summertime in the south-eastern United States, Atmos. Chem. Phys., 20, 12163–12176,https://doi.org/10.5194/acp-20-12163-2020, 2020.

    Brenguier, J.-L., Pawlowska, H., Schüller, L., Preusker, R., Fis-cher, J., and Fouquart, Y.: Radiative properties of boundary layerclouds: Droplet effective radius versus number concentration, J.Atmos. Sci., 57, 803–821, 2000.

    Bréon, F.-M., Tanré, D., and Generoso, S.: Aerosol effect on clouddroplet size monitored from satellite, Science, 295, 834–838,2002.

    Bretherton, C. S., Blossey, P. N., and Uchida, J.: Clouddroplet sedimentation, entrainment efficiency, and subtropi-cal stratocumulus albedo, Geophys. Res. Lett., 34, L03813,https://doi.org/10.1029/2006GL027648, 2007.

    Bulatovic, I., Ekman, A. M. L., Savre, J., Riipinen, I.,and Leck, C.: Aerosol Indirect Effects in Marine Stra-tocumulus: The Importance of Explicitly Predicting Cloud

    Atmos. Chem. Phys., 20, 15079–15099, 2020 https://doi.org/10.5194/acp-20-15079-2020

    https://doi.org/10.1038/nature03174https://doi.org/10.1126/science.245.4923.1227https://doi.org/10.1016/j.jqsrt.2012.03.025https://doi.org/10.5194/acp-12-2795-2012https://doi.org/10.1175/1520-0469(2003)0602.0.CO;2https://doi.org/10.1175/1520-0469(2003)0602.0.CO;2https://doi.org/10.5194/acp-9-543-2009https://doi.org/10.1080/20014422.1929.11880498https://doi.org/10.1038/nature04348https://doi.org/10.5194/acp-13-2045-2013https://doi.org/10.5194/acp-13-2045-2013https://doi.org/10.5194/essd-12-1649-2020https://doi.org/10.1029/2019RG000660https://doi.org/10.1029/2008JD011115https://doi.org/10.1029/2011GL047235https://doi.org/10.1175/2011BAMS2856.1https://doi.org/10.1029/2005JD006838https://doi.org/10.5194/amt-12-1841-2019https://doi.org/10.5194/amt-12-1841-2019https://doi.org/10.1034/j.1600-0889.47.issue3.1.xhttps://doi.org/10.1034/j.1600-0889.47.issue3.1.xhttps://doi.org/10.1017/CBO9781107415324.016https://doi.org/10.3390/atmos8060108https://doi.org/10.5194/acp-20-12163-2020https://doi.org/10.1029/2006GL027648

  • J. Quaas et al.: Twomey from satellite 15093

    Droplet Activation, Geophys. Res. Lett., 46, 3473–3481,https://doi.org/10.1029/2018GL081746, 2019.

    Burton, S. P., Chemyakin, E., Liu, X., Knobelspiesse, K., Stamnes,S., Sawamura, P., Moore, R. H., Hostetler, C. A., and Ferrare,R. A.: Information content and sensitivity of the 3β + 2α li-dar measurement system for aerosol microphysical retrievals, At-mos. Meas. Tech., 9, 5555–5574, https://doi.org/10.5194/amt-9-5555-2016, 2016.

    Cantrell, C. A.: Technical Note: Review of methods for lin-ear least-squares fitting of data and application to atmo-spheric chemistry problems, Atmos. Chem. Phys., 8, 5477–5487,https://doi.org/10.5194/acp-8-5477-2008, 2008.

    Carslaw, K. S., Lee, L. A., Reddington, C. L., Pringle, K. J., Rap,A., Forster, P. M., Mann, G. W., Spracklen, D. V., Woodhouse,M. T., Regayre, L. A., and Pierce, J. R.: Large contribution ofnatural aerosols to uncertainty in indirect forcing, Nature, 503,67–71, https://doi.org/10.1038/nature12674, 2013.

    Carslaw, K. S., Gordon, H., Hamilton, D. S., Johnson, J. S., Re-gayre, L. A., Yoshioka, M., and Pringle, K. J.: Aerosols in thePre-industrial Atmosphere, Curr. Clim. Change Rep., 3, 1–15,https://doi.org/10.1007/s40641-017-0061-2, 2017.

    Chang, F.-L. and Li, Z.: Estimating the vertical variation of clouddroplet effective radius using multispectral near-infrared satel-lite measurements, J. Geophys. Res., 107, AAC 7-1–AAC 7-12,https://doi.org/10.1029/2001JD0007666, 2002.

    Chen, R., Wood, R., Li, Z., Ferraro, R., and Chang, F.-L.: Study-ing the vertical variation of cloud droplets effective radius usingship and space-borne remote sensing data, J. Geophy. Res., 113,D00A02, https://doi.org/10.1029/2007/JD009596, 2008.

    Christensen, M., Jones, W. K., and Stier, P.: Aerosols En-hance Cloud Lifetime and Brightness along the Stratus-to-Cumulus Transition, P. Natl. Acad. Sci. USA, 117, 17591–17598, https://doi.org/10.1073/pnas.1921231117, 2020.

    Christensen, M. W. and Stephens, G. L.: Microphysical and macro-physical responses of marine stratocumulus polluted by underly-ing ships: 2. Impacts of haze on precipitating clouds, J. Geophys.Res., 117, D11203, https://doi.org/10.1029/2011JD017125,2012.

    Clarke, A. and Kapustin, V.: Hemispheric aerosol vertical profiles:anthropogenic impacts on optical depth and cloud nuclei, Sci-ence, 330, 1047–1047, 2010.

    Costantino, L. and Bréon, F.-M.: Analysis of aerosol-cloud inter-action from multi-sensor satellite observations, Geophys. Res.Lett., 37, L11801, https://doi.org/10.1029/2009GL041828, 2010.

    Costantino, L. and Bréon, F.-M.: Aerosol indirect effect onwarm clouds over South-East Atlantic, from co-located MODISand CALIPSO observations, Atmos. Chem. Phys., 13, 69–88,https://doi.org/10.5194/acp-13-69-2013, 2013.

    Costa-Surós, M., Sourdeval, O., Acquistapace, C., Baars, H., Car-bajal Henken, C., Genz, C., Hesemann, J., Jimenez, C., König,M., Kretzschmar, J., Madenach, N., Meyer, C. I., Schröd-ner, R., Seifert, P., Senf, F., Brueck, M., Cioni, G., En-gels, J. F., Fieg, K., Gorges, K., Heinze, R., Siligam, P. K.,Burkhardt, U., Crewell, S., Hoose, C., Seifert, A., Tegen, I., andQuaas, J.: Detection and attribution of aerosol–cloud interactionsin large-domain large-eddy simulations with the ICOsahedralNon-hydrostatic model, Atmos. Chem. Phys., 20, 5657–5678,https://doi.org/10.5194/acp-20-5657-2020, 2020.

    Desmons, M., Ferlay, N., Parol, F., Mcharek, L., and Van-bauce, C.: Improved information about the vertical locationand extent of monolayer clouds from POLDER3 measurementsin the oxygen A-band, Atmos. Meas. Tech., 6, 2221–2238,https://doi.org/10.5194/amt-6-2221-2013, 2013.

    Deuzé, J. L., Bréon, F.-M., Devaux, C., Goloub, P., Herman, M.,Lafrance, B., Maignan, F., Marchand, A., Nadal, F., Perry, G.,and Tanré, D.: Remote sensing of aerosols over land surfacesfrom POLDER-ADEOS-1 polarized measurements, J. Geophys.Res., 106, 4913–4926, https://doi.org/10.1029/2000jd900364,2001.

    Dusek, U., Frank, G. P., Hildebrandt, L., Curtius, J., Schneider, J.,Walter, S., Chand, D., Drewnick, F., Hings, S., Jung, D., Bor-rmann, S., and Andreae, M. O.: Size Matters More Than Chem-istry for Cloud-Nucleating Ability of Aerosol Particles, Science,312, 1375–1378, https://doi.org/10.1126/science.1125261, 2006.

    Ekman, A. M. L., Engström, A., and Söderberg, A.: Im-pact of Two-Way Aerosol–Cloud Interaction and Changes inAerosol Size Distribution on Simulated Aerosol-Induced DeepConvective Cloud Sensitivity, J. Atmos. Sci., 68, 685–698,https://doi.org/10.1175/2010JAS3651.1, 2011.

    Endo, S., Fridlind, A. M., Lin, W., Vogelmann, A. M., Toto,T., Ackerman, A. S., McFarquhar, G. M., R.C.Jackson, Jon-sson, H. H., and Liu, Y.: RACORO continental bound-ary layer cloud investigations: 2. Large-eddy simulations ofcumulus clouds and evaluation with in situ and ground-based observations, J. Geophys. Res.-Atmos., 120, 5993–6014,https://doi.org/10.1002/2014JD022525, 2015.

    Fan, J., Leung, L. R., Rosenfeld, D., Chen, Q., Li, Z.,Zhang, J., and Yan, H.: Microphysical effects determinemacrophysical response for aerosol impacts on deep convec-tive clouds, P. Natl. Acad. Sci. USA, 110, E4581–E4590,https://doi.org/10.1073/pnas.1316830110, 2013.

    Feingold, G. and Grund, C. J.: Feasibility of Us-ing Multiwavelength Lidar Measurements to Mea-sure Cloud Condensation Nuclei, J. Atmos. Ocean.Tech., 11, 1543–1558, https://doi.org/10.1175/1520-0426(1994)0112.0.CO;2, 1994.

    Feingold, G., Yang, S., Hardesty, R. M., and Cotton, W. R.: Retriev-ing cloud condensation nucleus properties from Doppler cloudradar, microwave radiometer, and lidar, J. Atmos. Ocean. Tech.,15, 1189–1196, 1998.

    Feingold, G., Remer, L. A., Ramaprasad, J., and Kaufman, Y.: Anal-ysis of smoke impact on clouds in Brazilian biomass burningregions: An extension of Twomey’s approach, J. Geophys. Res.,106, 22907–22922, https://doi.org/10.1029/2001JD000732,2001.

    Feingold, G., Eberhard, W. L., Veron, D. E., and Previdi,M.: First measurements of the Twomey indirect effect usingground-based remote sensors, Geophys. Res. Lett., 30, 1287,https://doi.org/10.1029/2002GL016633, 2003.

    Feingold, G., Koren, I., Wang, H. L., Xue, H. W., and Brewer, W. A.:Precipitation-generated oscillations in open cellular cloud fields,Nature, 466, 849–852, 2010.

    Geddes, A. and Bösch, H.: Tropospheric aerosol profile informationfrom high-resolution oxygen A-band measurements from space,Atmos. Meas. Tech., 8, 859–874, https://doi.org/10.5194/amt-8-859-2015, 2015.

    https://doi.org/10.5194/acp-20-15079-2020 Atmos. Chem. Phys., 20, 15079–15099, 2020

    https://doi.org/10.1029/2018GL081746https://doi.org/10.5194/amt-9-5555-2016https://doi.org/10.5194/amt-9-5555-2016https://doi.org/10.5194/acp-8-5477-2008https://doi.org/10.1038/nature12674https://doi.org/10.1007/s40641-017-0061-2https://doi.org/10.1029/2001JD0007666https://doi.org/10.1029/2007/JD009596https://doi.org/10.1073/pnas.1921231117https://doi.org/10.1029/2011JD017125https://doi.org/10.1029/2009GL041828https://doi.org/10.5194/acp-13-69-2013https://doi.org/10.5194/acp-20-5657-2020https://doi.org/10.5194/amt-6-2221-2013https://doi.org/10.1029/2000jd900364https://doi.org/10.1126/science.1125261https://doi.org/10.1175/2010JAS3651.1https://doi.org/10.1002/2014JD022525https://doi.org/10.1073/pnas.1316830110https://doi.org/10.1175/1520-0426(1994)0112.0.CO;2https://doi.org/10.1175/1520-0426(1994)0112.0.CO;2https://doi.org/10.1029/2001JD000732https://doi.org/10.1029/2002GL016633https://doi.org/10.5194/amt-8-859-2015https://doi.org/10.5194/amt-8-859-2015

  • 15094 J. Quaas et al.: Twomey from satellite

    Gelaro, R., McCarty, W., Suárez, M. J., Todling, R., Molod, A.,Takacs, L., Randles, C. A., Darmenov, A., Bosilovich, M. G., Re-ichle, R., Wargan, K., Coy, L., Cullather, R., Draper, C., Akella,S., Buchard, V., Conaty, A., da Silva, A. M., Gu, W., Kim, G.-K., Koster, R., Lucchesi, R., Merkova, D., Nielsen, J. E., Par-tyka, G., Pawson, S., Putman, W., Rienecker, M., Schubert, S. D.,Sienkiewicz, M., and Zhao, B.: The Modern-Era RetrospectiveAnalysis for Research and Applications, Version 2 (MERRA-2),J. Climate, 30, 5419–