Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Climate Dynamics (2019) 52:4371–4392

https://doi.org/10.1007/s00382-018-4384-z

Assessment of aerosol–cloud–radiation correlations in satellite


observations, climate models and reanalysis
F. A.‑M. Bender1 · L. Frey1 · D. T. McCoy2 · D. P. Grosvenor3 · J. K. Mohrmann4

Received: 20 February 2018 / Accepted: 3 August 2018 / Published online: 21 August 2018
© The Author(s) 2018

Abstract
Representing large-scale co-variability between variables related to aerosols, clouds and radiation is one of many aspects of
agreement with observations desirable for a climate model. In this study such relations are investigated in terms of temporal
correlations on monthly mean scale, to identify points of agreement and disagreement with observations. Ten regions with
different meteorological characteristics and aerosol signatures are studied and correlation matrices for the selected regions
offer an overview of model ability to represent present day climate variability. Global climate models with different levels of
detail and sophistication in their representation of aerosols and clouds are compared with satellite observations and reanalysis
assimilating meteorological fields as well as aerosol optical depth from observations. One example of how the correlation
comparison can guide model evaluation and development is the often studied relation between cloud droplet number and
water content. Reanalysis, with no parameterized aerosol–cloud coupling, shows weaker correlations than observations, indi-
cating that microphysical couplings between cloud droplet number and water content are not negligible for the co-variations
emerging on larger scale. These observed correlations are, however, not in agreement with those expected from dominance
of the underlying microphysical aerosol–cloud couplings. For instance, negative correlations in subtropical stratocumulus
regions show that suppression of precipitation and subsequent increase in water content due to aerosol is not a dominating
process on this scale. Only in one of the studied models are cloud dynamics able to overcome the parameterized dependence
of rain formation on droplet number concentration, and negative correlations in the stratocumulus regions are reproduced.

Keywords Aerosol–cloud–radiation interaction · GCM-evaluation · Satellite observations · Reanalysis · Volcanic aerosol

1 Introduction

Systematic variations in aerosol amount and composi-


tion over time give rise to a forcing on the climate system,
through aerosol interaction with clouds and with radia-
tion. The estimate of aerosol forcing, from pre-industrial
Electronic supplementary material The online version of this (PI) to present day (PD), is at present highly uncertain
article (https​://doi.org/10.1007/s0038​2-018-4384-z) contains (Boucher et al. 2013) and to achieve the goal of constrain-
supplementary material, which is available to authorized users. ing the estimate of aerosol forcing, including that from aer-
* F. A.‑M. Bender osol-interactions with clouds, a better understanding of the
frida@misu.su.se relations between aerosols, cloud microphysics and cloud
macrophysics is necessary, although not sufficient.
1
Department of Meteorology and Bolin Centre for Climate In lack of long-term observations of variations in aerosol
Research, Stockholm University, Stockholm, Sweden
and cloud, it is customary to relate present-day, short-term
2
School of Earth and Environment, Institute of Climate relations between aerosol and cloud properties to those on
and Atmospheric Science, University of Leeds, Leeds, UK
the time scale of anthropogenic forcing. It has been shown,
3
National Centre for Atmospheric Science, University however, that present-day relations and their uncertainties
of Leeds, Leeds, UK
may not be representative of those dominating the PI–PD
4
Department of Atmospheric Sciences, University changes, and the use of present-day variability to constrain
of Washington, Seattle, USA

13
Vol.:(0123456789)
4372 F. A.-M. Bender et al.

PI–PD forcing has been questioned (Penner et al. 2011; droplet number concentration ( Nd ), the former instantane-
Carlsaw et al. 2013; Ghan et al. 2016; Lee et al. 2016). ously and the latter as a rapid adjustment. As the variability
Ghan et al. (2016) decompose the dependence of cloud in Nd has been found to be driven mainly by sulfate aerosol
radiative forcing on aerosol into individual sensitivity fac- (Boucher and Lohmann 1995; McCoy et al. 2017), sulfate
tors, and find from AeroCom (Aerosol Comparisons between ( SO4 ) is the aerosol species we focus on.
Observations and Models) models that determining these Cloud brightening occurs when, all else equal, increased
relations from recent variability typically offers poor con- aerosol number yields more numerous and smaller cloud
straint on the anthropogenic forcing, and in some cases droplets (Twomey 1974, 1977). Cloud thickening in turn is
even results in sensitivities of opposite signs. Penner et al. an effect of the precipitation efficiency in such small-drop-
(2011) suggest that satellite-derived sensitivity of cloud let clouds being reduced, leading to a build-up of conden-
droplet number to aerosol optical depth strongly underes- sate both in terms of prolonged cloud lifetime and larger
timates the actual change in cloud droplet number from PI fractional cloud cover, and in terms of cloud geometrical
to PD, while Gryspeerdt et al. (2017) demonstrate that by thickness (Albrecht 1989; Pincus and Baker 1994; Bren-
accounting for multiple predictors of cloud droplet number, guier et al. 2000). Thereby a priori, the 1st indirect effect act-
the radiative forcing from aerosol–cloud interactions can ing in isolation would lead to positive correlations between
be well constrained by PD relationships, provided that the Nd and cloud albedo (𝛼cloud ), and similarly the 2nd indirect
anthropogenic aerosol component is known. Carlsaw et al. effect acting in isolation would yield positive correlations
(2013) ascribe a large fraction of the uncertainty in anthro- between Nd and cloud liquid water path (L) and/or cloud
pogenic aerosol forcing to uncertainty in the PI state, that fraction ( fc).
may not necessarily be reduced by further constraint of PD However, neither of the two effects can be isolated
conditions, and Lee et al. (2016) point at the fact that model from the other, nor from other processes. Just the fact that
uncertainty allows for multiple different models agreeing L-changes driven by the 2nd indirect effect will violate the
sufficiently well with PD aerosol observations, but still pro- conditions of the 1st indirect effect illustrates this difficulty,
ducing a large range of forcing estimates. see e.g. McComiskey and Feingold (2012).
In this study, we do not attempt to constrain aerosol As discussed by Ackerman et al. (2004), Jiang and Fein-
forcing, per se. Rather, we set out to test relations between gold (2006), Stevens and Feingold (2009), Bretherton et al.
various aerosol-, cloud-, and radiation-related variables on (2007), Wood (2007), Small et al. (2009), Jiang et al. (2011),
climatologically relevant scale, proposing a framework for Rosenfeld et al. (2014) and Neubauer et al. (2017) among
evaluation, and thereby providing guidance to better model others, there are also other counteracting mechanisms related
representation of these relations and continued model to aerosol and cloud dynamics, and these correlations do not
improvement. Even if agreement with PD observations is necessarily protrude when all processes are allowed to act
not sufficient for constraining model estimates of past or together, as must be expected to typically be the case in
future forcing, a reasonable representation of the state, vari- the real atmosphere. Therefore, a comparison of correlation
ability and co-variability of observable quantities is desired strengths and directions can be of aid, indicating what mat-
and required. It remains essential that global climate models ters and what does not, on a larger scale.
(GCMs) used for process-studies as well as for future projec- In a preceding attempt to compare observed and mod-
tions are able to realistically capture the relations between elled relations between Nd and L in particular, i.e. focusing
climatologically relevant aerosol and cloud properties, on the 2nd indirect effect, Michibata et al. (2016) find that
including those emerging from aerosol influence on cloud the cloud susceptibility to aerosol perturbation is region
water content, which is central to aerosol–cloud interactions dependent, and also that the MIROC GCM overestimates
in warm clouds. the increase in L with increasing Nd as compared to sat-
In terms of aerosol–cloud interactions, we focus on ellite observations. They ascribe the discrepancy to a too
the effects of aerosols on cloud albedo (referred to as the dominant role of microphysics in this model, not allowing
Twomey effect, the cloud albedo effect, cloud brightening for macrophysical feedbacks affecting the cloud water. It
or the 1st indirect effect) and cloud amount (referred to as is noteworthy here, that the Nd and L retrievals utilized by
the Albrecht effect, the cloud lifetime effect, cloud thick- Michibata et al. (2016) may be biased in regions of broken
ening or the 2nd indirect effect). Although the classical cloud cover (Cho et al. 2015; Seethala and Hovath 2010).
nomenclature of aerosol indirect effects, as opposed to direct Prior to this, Quaas et al. (2009) used observed statistical
effects, has largely been abandoned (Boucher et al. 2013), relationships between aerosol optical depth ( 𝜏a ) and vari-
it is instructive to separate the effects on cloud albedo and ous cloud parameters to estimate aerosol forcing, based on
cloud amount. their finding that the same relationships in an ensemble of
Both the cloud brightening (1st indirect effect) and the models are actually skilful in predicting simulated forcing.
cloud thickening (2nd indirect effect) act via the cloud Quaas et al. (2009), similar to Michibata et al. (2016), draw

13
Aerosol-cloud-radiation correlations 4373

the conclusion that the implementation of the 2nd indirect (Sc) regions (Californian, Peruvian, Namibian, Australian
effect in models by delayed autoconversion from cloud drop- and Canarian, as defined by Klein and Hartmann (1993)), a
let to drizzle at larger droplet number concentration, results category that serves as a reference, as aerosol–cloud–radia-
in a too strong sensitivity of L to aerosol both in terms of Nd tion relations in these regions have been studied rather exten-
and 𝜏a . Neubauer et al. (2017) investigate in detail the influ- sively (Wood 2007; Bender et al. 2016; Frey et al. 2017, e.g.)
ence of aerosol water uptake on the co-variability between and have been indicated as particularly susceptible to aerosol
aerosol and cloud water and find that dry aerosol better pre- influence, at least in climate models (Kirkevåg et al. 2013;
dicts the observed variability in L, and that water uptake, wet Carlsaw et al. 2013).
scavenging and cloud processing should all be minimized to We add to the previous literature in several ways:
isolate the susceptibility of cloud water to aerosol changes
and compute aerosol forcing. • Focus regions are selected based on aerosol signature
Several studies, e.g. Gassó (2008) and Toll et al. (2017) as well as dynamical regime, motivated by the regional
have made use of the natural experiment supplied by vol- variability in aerosol–cloud relations discussed by e.g.
canic eruptions and their emissions of sulfate aerosol pre- Michibata et al. (2016), Zhang et al. (2016) and Neu-
cursors, namely sulfur dioxide (SO2 ), to study potential bauer et al. (2017).
effects on cloud properties. McCoy and Hartmann (2015) • Compared to Quaas et al. (2009), Michibata et al.
specifically found that during the 2014–2015 eruption of (2016) and Toll et al. (2017) we include additional
Holohraun in Iceland, the cloud droplet size (represented years and sensors in the satellite data analysis, which
by effective radius, re ) significantly decreased in the area provides an indication of retrieval biases and influence
downwind of the volcano. Analysing the same volcanic of temporal variability and sensitivity to time period.
eruption, Malavelle et al. (2017) indicate that re is actually • Monthly mean time scale is used, rather than daily
the only cloud-variable that responds to the increased SO2 (Quaas et al. 2009) or instantaneous (Neubauer et al.
and SO4 . Malavelle et al. (2017) also suggest that the indif- 2017). This is to emphasize not the process level, but
ference of L to aerosol perturbation can be used to constrain the relations emerging at climatologically relevant time
models in a general sense; models with a significant 2nd scales, which are not necessarily identical, cf. Bender
indirect effect, i.e. in which L increases when aerosol and et al. (2016), Konsta et al. (2016).
droplet number increases, should be considered less realistic. • We perform a multi-model comparison, rather than a
Toll et al. (2017) found that in volcano tracks, i.e. spatially single model evaluation or sensitivity experiments with
limited, linear cloud modifications due to underlying vol- one model as Neubauer et al. (2017), Michibata et al.
canic emissions analogous to ship tracks, i.e. similar cloud (2016) or Toll et al. (2017), providing an evaluation of
features due to ship emissions, the L response to the aerosol the subset of those models in the CMIP5 ensemble (see
perturbation varies and is on average close to zero, while the Section 2.3) that supply the necessary output variables.
HadGEM model simulates a consistently positive response As pointed out in the model intercomparison made by
in L to aerosol. Quaas et al. (2009), models evolve and therefore evalu-
To investigate the emerging relations between aerosol and ations need to be continually updated.
cloud properties in an array of models and observations, • We include MERRA-2 (see Sect. 2.2), an independent
we here look at several regions, in three main categories. set of reanalysis data that is useful because it includes
Following recent literature (McCoy and Hartmann 2015; dynamical and meteorological feedbacks but has no
Mace and Abernathy 2016; Malavelle et al. 2017; McCoy parameterized link between aerosol and cloud micro-
et al. 2018) we include three regions of volcanic influence physics, and can therefore be used as a “no-indirect-
(Iceland, Vanuatu, Hawaii). While the Vanuatu and Hawaii effect” reference case.
regions have been assessed to be among the largest sources • We focus not only on Nd –L relations (as Michibata
of volcanic SO2 degassing in the past decade (Carn et al. et al. 2016; Malavelle et al. 2017; Neubauer et al. 2017)
2017), the Iceland region is not degassing in the same way, or on sensitivities to 𝜏a perturbations (as Quaas et al.
but has experienced eruptive events during the time period (2009)) or on the specific relations behind the micro-
studied (Grìmsvötn May 2011, Eyjafjallajökull March–June physical interactions assumed to drive the 1st and 2nd
2010, Holohraun Aug 2014–Feb 2015), as has Hawaii indirect effect (as Ghan et al. 2016), but give a broader
(Kilauea June–August 2008). In contrast to these natural picture of relations between macro- and microphysical
sulfate-dominated regions we add two regions of strong cloud and aerosol properties, that can be used to test
anthropogenic influence (US, China). The Iceland region and evaluate models.
with its less remote location compared to the other two vol- • Rather than “susceptibility” i.e. linear regression coef-
canic regions may also be expected to be anthropogenically ficients, we study correlations (with statistical sig-
influenced. Finally, we study five subtropical stratocumulus nificance estimates) and hereby avoid the problem of

13
4374 F. A.-M. Bender et al.

potentially assigning a large susceptibility to a relation SSF (Single Scanner Footprint) Edition 4, level 3 product,
between variables that are in fact not well correlated. available from the NASA Langley Research Center Atmos-
We provide correlation matrices for individual regions pheric Science Data Center. Observed estimates of albedo
that can readily be re-produced for the same or different ( 𝛼 ), clear-sky albedo ( 𝛼clear ), fc , L, re and 𝜏a averaged to
regions, and used in the process of model development, 1◦ × 1◦ resolution on monthly mean time scale are analyzed.
tuning and evaluation. In the SSF data set, clear-sky is determined by the
CERES-MODIS cloud mask algorithm (Minnis et al. 2009)
applied to the MODIS pixels within the CERES footprint.
2 Methods and data Each MODIS pixel (size at nadir ranging from 250 to 1000
m depending on wavelength) within the CERES footprint
We make use of three main data sources: satellite observa- (20 km nominal resolution) is determined as either clear or
tions, nudged reanalysis and climate model output. With this cloudy and the cloud fraction is the ratio of cloudy pixels
data selection we illustrate the use of correlations between to the total number of pixels. MODIS cloud properties (col-
various micro- and macrophysical cloud and aerosol proper- lection 5) are derived from five channels in the visible and
ties, as a way to evaluate models. infrared (Minnis et al. 2009, 2011), and like the MODIS 𝜏a
Correlation does not imply causation, and e.g. Eng- (Remer et al. 2005, 2008) are mapped to the lower CERES
ström and Ekman (2010) illustrate specifically how correla- resolution.
tions between cloud and aerosol quantities can be strongly CERES and MODIS are carried by the polar orbiting
affected by other factors than microphysical relations. But Aqua and Terra satellites, with local equator crossing times
conversely, if a specific microphysical coupling were strong of 10.30 AM and 1.30 PM respectively. Observations from
enough to dominate on a climatologically relevant scale, it Aqua are used in the present study, following Malavelle et al.
would give rise to a correlation. Thereby it follows that a (2017). The gridded radiative fluxes are derived using angu-
lack of correlation indicates that no single microphysical lar distribution models as described by Loeb et al. (2005),
coupling is dominating. and diurnally averaged assuming constant meteorology
The study regions are defined in Table 1. All analysis is between the satellite overpasses, after which monthly means
performed on monthly averaged data, corrected for a clima- are created.
tological seasonal cycle, i.e. values are given as anomalies MODIS Nd is calculated as a function of cloud optical
relative to the regional mean climatology. Details on each of thickness ( 𝜏c ) and cloud top re on daily mean scale at 1◦ ×
the data sources are provided in the following Sects. 2.1–2.3. 1◦ resolution, using level 2 MODIS swath data filtered to
include only low, liquid clouds in cases where daily grid
2.1 Satellite observations cloud fraction exceeds 80%, and to exclude pixels with
solar zenith angle (𝜃 ) greater than 65◦ and other problem-
Observed top-of-atmosphere shortwave (0.2–5 μm) radiative atic retrievals, as according to Grosvenor and Wood (2014).
fluxes from CERES [(Clouds and the Earth’s Radiant Energy The daily data are then averaged to monthly means. See also
System (Wielicki et al. 1996)] and cloud and aerosol proper- McCoy et al. (2018) for a closer description and evaluation
ties from MODIS (MODerate resolution Imaging Spectro- of the same Nd data set.
radiometer (Barnes et al. 1998) are taken from the CERES In addition to MODIS L we make use of the Multisensor
Advanced Climatology (MAC) L product (Elsaesser et al.
2016) that is based on the combination of data from SSM/I
Table 1  Study regions, geographically specified and categorized and SSMIS (Special Sensor Microwave Imager/Sounder),
based on aerosol signature or cloud regime AMSR-E and AMSR-2 (Advanced Microwave Scanning
Region Category Longitude–latitude Radiometers), TRMM (Tropical Rainfall Measuring Mis-
sion) and GPM (Global Precipitation Measurement) micro-
Iceland Volcanic 30–10◦ W , 50–70◦ N
wave imagers and WindSat satellite sensors. The compila-
Vanuatu Volcanic 155–175◦ E, 25–5◦ S
tion, inter-calibration and bias-correction of the data sets
Hawaii Volcanic 170–150◦ W , 10–30◦ N
into a monthly gridded (1◦ × 1◦ resolution) global ocean
China Anthropogenic 110–130◦ E, 20–40◦ N
L climatology is described by Elsaesser et al. (2017). An
US Anthropogenic 80–60◦ W , 30–50◦ N
advantage of the MAC L is that it avoids the retrieval failures
Californian Stratocumulus 130–120◦ W , 20–30◦ N
in regions of broken clouds, that may lead to sampling biases
Peruvian Stratocumulus 90–80◦ W , 20–10◦ S
in the MODIS L (Seethala and Hovath 2010; Grosvenor and
Australian Stratocumulus 95–105◦ E, 35–25◦ S
Wood 2014; Cho et al. 2015). The microwave measurements
Namibian Stratocumulus 0–10◦ N, 20–10◦ S
provide a grid-box average L, and to make this estimate con-
Canarian Stratocumulus 35–25◦ S, 15–25◦ N
sistent with that from MODIS, and more directly relatable

13
Aerosol-cloud-radiation correlations 4375

to in-cloud Nd estimates, MAC L is divided by the MODIS IPSL-CM5B-LR, MIROC5, MIROC-ESM, MRI-CGCM3
fc, yielding an in-cloud L estimate. The two L estimates will and MRI-ESM1. The two versions of IPSL-CM5A dif-
be referred to as LMODIS and LMAC respectively, but we note fer in resolution only, whereas the third model from the
that through the conversion to in-cloud L, the LMAC is not same institute, IPSL-CM5B-LR, has significantly differ-
independent of MODIS. ent physics, including boundary layer, cloud and convec-
All the observational data used span the time period Janu- tion schemes (Dufresne et al. 2013; Hourdin et al. 2013),
ary 2003 through December 2015, i.e. 13 years. as further evaluated by Konsta et al. (2016). MIROC5 and
MIROC-ESM both use the same aerosol scheme, but differ-
2.2 Reanalysis ent cloud schemes; MIROC5 is based on Wilson and Ballard
(1999) and MIROC-ESM is based on Treut and Li (1991).
MERRA-2 (Modern-Era Retrospective analysis for Research Further, MIROC-ESM in addition to the ocean, land sur-
and Applications, version 2) is an atmospheric reanaly- face and chemistry models includes atmospheric chemistry,
sis dataset, based on the Goddard Earth Observing Sys- ocean- and land ecosystem models with dynamic vegetation
tem Model, version 5 (GEOS-5) (Rienecker et al. 2011; (Watanabe et al. 2010, 2011). Similarly, MRI-CGCM3 and
Molod et al. 2015). In addition to assimilation of meteoro- MRI-ESM1 are based on the same atmosphere-ocean core
logical and cloud variables including rain-rate, water vapour component, but the ESM has a greater complexity including
path and wind speed, MERRA-2 also assimilates 𝜏a from sat- a sophisticated representation of the carbon cycle (Yuki-
ellite (AVHRR, MODIS, MISR) and surface based remote moto et al. 2012). An error in the microphysics has been
sensing (Aeronet), as described in Randles et al. (2016). identified in the CMIP5-archived historical simulations of
MERRA-2 also implements an on-line aerosol chemistry, MRI, related to the prognostic equations for cloud droplet
radiation and transport model (Colarco et al. 2010). Hourly number concentration and yielding unrealistically large Nd
averages of the assimilated reanalysis product at 0.5◦ × values (Kawai et al. 2017). As a reference we therefore also
0.625◦ resolution are averaged to a monthly mean 1◦ × 1◦ include analysis of an MRI-CGCM3 simulation in which
grid, in the case of all variables except the sulfate mass con- this error has been rectified. This is an AMIP-type fixed-
centration for which monthly means are created from daily SST simulation with pre-industrial greenhouse gas forcing
averages of the instantaneous 3-h reanalysis output. We note and year 2000 aerosol fields, and it will be referred to as
that the agreement with 𝜏a observations may vary during MRI-fixed.
the analysis cycle, as a result of biases in model forecast Whereas all the investigated models include a parameteri-
compared to the assimilated MODIS observations, but the zation of the 1st indirect effect, they differ in their represen-
average product are in close agreement with the observed tation of Nd , and the 2nd indirect effect, as summarized in
fields (Randles et al. 2016). Table 2. MIROC5 and MIROC-ESM include parameteri-
In MERRA-2, the aerosol module is coupled to the mete- zations of the 1st and 2nd indirect effects, prognostic Nd
orology so that e.g. wet removal and humidification affect dependent on supersaturation as well as aerosol number and
the aerosol distribution, but there is no parameterized micro- composition, and re dependent on Nd and cloud water mixing
physical link between aerosols and clouds, and MERRA-2 ratio (Watanabe et al. 2010, 2011; Takemura et al. 2005).
in our study therefore acts as a “no-indirect-effect” reference MRI-CGCM3 and MRI-ESM1 similarly include the 1st
case. L is not directly assimilated in MERRA-2, and the only and 2nd indirect effects, explicit treatment of aerosol acti-
way for potential real-world aerosol effects to influence the vation into cloud droplets, and re dependent on cloud drop-
reanalysis L would be via the assimilated water vapour path let number density and cloud water (Yukimoto et al. 2012).
or rain rates, both weak links. MERRA-2 further has no Nd
estimate, but McCoy et al. (2017) have shown that SO4 from
MERRA-2 is a good predictor of observed Nd , and hence we Table 2  CMIP5 models used in the study, and their atmospheric reso-
use reanalysis SO4 as a proxy for Nd , with global coverage. lutions and representation of Nd and indirect effects
This metric will be referred to as NdMERRA. MERRA-2 sup- Name Atm. resolution Nd-param. IE
plies a grid-box average L, that is divided by fc to yield an
IPSL-CM5A-LR 96 × 95 × 39 Diagnostic 1st only
in-cloud estimate of L, consistent with MODIS observations.
IPSL-CM5A-MR 144 × 143 × 39 Diagnostic 1st only
IPSL-CM5B-LR 96 × 95 × 39 Diagnostic 1st only
2.3 Coupled climate models
MIROC5 T85L40 Prognostic 1st and 2nd
MIROC-ESM T42L80 Prognostic 1st and 2nd
We include seven models from the Coupled Model Inter-
MRI-CGCM3 TL159L48 Prognostic 1st and 2nd
comparison Project, phase 5 (CMIP5, (Taylor et al. 1996))
MRI-ESM1 TL159L48 Prognostic 1st and 2nd
that provide the required output fields in their historical
MRI-fixed TL159L48 Prognostic 1st and 2nd
simulations, namely IPSL-CM5A-LR, IPSL-CM5A-MR,

13
4376 F. A.-M. Bender et al.

IPSL-CM5A-LR, IPSL-CM5A-MR and IPSL-CM5B-LR 2.5 Data aggregation issues


include the 1st indirect effect but not the 2nd indirect effect,
diagnose Nd from total mass of water soluble aerosol, and Aerosol influence on clouds span a range of spatiotemporal
re from aerosol concentration (Dufresne et al. 2013; Hour- scales, and relations between relevant variables are scale-
din et al. 2013). dependent (Bender et al. 2016; Konsta et al. 2016; Feingold
The CMIP time period analyzed differs from that of et al. 2016).
observations and reanalysis. With the exception of MIROC5 Estimates of radiative forcing from aerosol–cloud interac-
whose historical simulation reaches year 2012, the histor- tions have also been found to vary depending on the spatial
ical experiments in the models end in year 2005 and for scale chosen in both models and observations, but in yet
consistency between models here we use a 10-year period inconclusive ways (Grandey and Stier 2010; McComiskey
1996–2005, also avoiding large volcanic eruptions (like and Feingold 2012; Possner et al. 2016).
the Pinatubo eruption in 1992 that is parameterized in the In comparisons between models and observations, the
MRI models). The Holohraun, Eyjafjallajökull and Vanuatu differences in sampling and spatial and temporal resolution
eruptions in the 2010’s are hence not included in the model may introduce large biases, especially in attempts to evaluate
simulations. models against point measurements. Such issues are dis-
The grid-box average model L provided is divided by fc , cussed in detail by Schutgens et al. (2016).
for consistency with in-cloud observations and estimates of The approach taken here is to aggregate data to a common
Nd , cf. Jiang et al. (2012). spatial scale of regional averages, and a common temporal
The model Nd analyzed (cldncl) is the droplet number scale of monthly means. Spatio-temporal averaging reduces
concentration at cloud top, considering liquid clouds only, the influence of sampling errors (Schutgens et al. 2016), and
and weighted by total liquid cloud fraction at each time step, allows for investigation of the ability of models to represent
before being averaged to a monthly mean. This is the droplet relations on observable and climatologically relevant scales.
number measure that is available for the largest number of
models, and is also most directly comparable to observa-
tions. A few models supply temporally evolving 3D-fields 3 Results
of Nd (cdnc) and others provide only a vertically integrated
droplet number concentration (cldnvi) that is inherently 3.1 Regional characteristics
related to L, via cloud geometrical thickness.
The focus regions listed in Table 1 are indicated in Fig. 1,
2.4 Limitations to estimates of 𝜏a , Nd and L also showing the 13-year average global distribution of 𝜏a
from MODIS.
As pointed out by e.g. Malavelle et al. (2017), MODIS 𝜏a is Figure 2 shows regional mean time series of observed
challenging as it is only retrieved under cloud-free condi- aerosol, cloud and radiation variables (𝛼 , 𝛼cloud , 𝛼clear , fc ,
tions. This means that grid-box averaged 𝜏a values are calcu- 𝜏a , L, re, Nd ), over the same time period. There are several
lated assuming that the cloud-free retrievals are representa- notable regional differences in absolute values and variabil-
tive for the entire grid box. This is standard procedure, (see ity of both aerosol and cloud metrics.
e.g. Quaas et al. (2009), Bender et al. (2016)), but in cases
of very large cloud cover, as frequently encountered over 3.1.1 Stratocumulus and anthropogenically influenced
the mid-latitudes in winter, no 𝜏a retrievals can be made. regions
MERRA-2 does not have this problem, but as MERRA-2
is nudged to satellite-observed 𝜏a , corrected for swelling of The Sc regions are similar in dynamical cloud regime, and
aerosol near cloud, any residual MODIS 𝜏a-issues may indi- typically have high fc and low L, indicating thin clouds,
rectly affect the MERRA-2 data. which is also in agreement with the relatively low 𝛼cloud in
On the other hand, we require a minimum cloud fraction these regions. At the same time the regions differ in aerosol
of 80% for the Nd retrievals to be considered valid (Bennartz signature. The Canarian region, shows the greatest 𝜏a and 𝜏a
et al. 2011). In addition, reliable Nd retrievals cannot be variability, related to the dominance of desert dust and the 𝜏a
made at high latitudes in winter months, because the low sun is also high in the Namibian region, that has a comparatively
angle breaks the plane-parallel radiative transfer assump- strong signal of black carbon from biomass burning (cf. Frey
tions (Grosvenor and Wood 2014). The large 𝜃 afflicts also et al. (2017)).
the MODIS L-retrievals, and for these reasons, the analyzed The anthropogenically influenced regions (US, China)
time series for the Iceland region exclude the monthly means have distinctively larger Nd than the other regions. Their
for November through February for 𝜏a and Nd as well as L. re is also comparatively low, in particular in relation to the
more remote volcanic regions. The 𝜏a is high, particularly

13
Aerosol-cloud-radiation correlations 4377

°
90 N

1
°
45 N
5
6 4
10
3

°
0
°
180 W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E
7 9
2
8

45 ° S

90 ° S

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4


τ
a

Fig. 1  Global ocean 𝜏a distribution, averaged from 2003 through 2015. Regions are (1) Iceland, (2) Vanuatu, (3) Hawaii, (4) US, (5) China, (6)
Californian, (7) Peruvian, (8) Australian, (9) Namibian and (10), in all cases excluding land-covered grid points

for China. Both regions display decreasing trends in 𝜏a and This can not, however, be considered a reliable signal, due
Nd during the given time period, consistent with emission to the high 𝜃 biasing the few data points that are reported.
reductions (Krotkov et al. 2016; Zhao et al. 2017; McCoy The MAC-derived L, which is not afflicted with the same
et al. 2018). retrieval problems at high 𝜃 does not indicate a similar
The MAC L estimates are typically higher than those anomaly in L at the time of the eruption.
from MODIS, which is likely in part related to MODIS McCoy and Hartmann (2015) focus on the months of
retrieval biases in regions of broken clouds (Seethala and September–November, finding a negative re anomaly that is
Hovath 2010; Cho et al. 2015). more persistent in space and time. In that case the attribu-
tion to volcanic aerosol may rather be complicated by the
3.1.2 Volcanic regions fact that the flow pattern varies throughout the period, and
is not consistently coincident with observed anomalies (see
For the Iceland region, re drops at the time of the Holohraun e.g. trajectory analysis in McCoy and Hartmann (2015)).
eruption, as discussed in detail by McCoy and Hartmann Despite the locally large L anomalies and the re drop that
(2015) and also confirmed by Malavelle et al. (2017). The appear during the time of the Holohraun eruption, no signal
MODIS and MAC L estimates agree well for this region, and can be seen in 𝛼cloud or 𝛼 , indicating that the actual radia-
neither of them indicate a peak at the time of the Holohraun tive effect of the cloud property anomalies is weak. Corre-
eruption, in agreement with Malavelle et al. (2017). The sponding drops in re at the time of the eruptions of Eyjafjal-
spatial variability in L is large, however, and Malavelle et al. lajökull (Iceland March–June 2010) and Kilauea (Hawaii
(2017) average over an area where positive anomalies in the June–August 2008), the latter described by Yuan et al.
south-west are balanced by negative anomalies in the north- (2012), Malavelle et al. (2017), are less marked and within
east. Our Iceland domain is smaller, but still its average L the noise level of the monthly regional mean time series, and
does not show an increase during the months of September L peaks are not discernible at these points in time.
to October, when the volcanic emissions were at their peak We note that Malavelle et al. (2017) convert the MODIS
(Schmidt et al. 2015). During the following winter months, in-cloud L estimate to a grid-box average, which contributes
most prominently December, see Figure S1, the MODIS data to smaller L values than the in-cloud estimates shown here.
indicate an anomalously high L over the Iceland domain.

13
4378 F. A.-M. Bender et al.

Californian
Peruvian
Iceland Australian
Volcanic regions Vanuatu
Anthropogenic regions China
Stratocumulus regions Namibian
Hawaii US Canarian
α 0.5 0.5 0.5
α cloud α
α clear
0 0 0
2003 2006 2009 2012 2015 2003 2006 2009 2012 2015 2003 2006 2009 2012 2015
1 1 1

0.5 0.5 0.5


c
f

0 0 0
2003 2006 2009 2012 2015 2003 2006 2009 2012 2015 2003 2006 2009 2012 2015

0.5 0.5 0.5


a
τ

0 0 0
2003 2006 2009 2012 2015 2003 2006 2009 2012 2015 2003 2006 2009 2012 2015

200 200 200


L [gm -2]

MODIS
MAC 100 100 100

0 0 0
2003 2006 2009 2012 2015 2003 2006 2009 2012 2015 2003 2006 2009 2012 2015
20 20 20
r [µm]

10 10 10
e

0 0 0
2003 2006 2009 2012 2015 2003 2006 2009 2012 2015 2003 2006 2009 2012 2015

Eyjafjallajökull
N [cm-3]

200 Kilauea Holohraun 200 200


d

0 0 0
2003 2006 2009 2012 2015 2003 2006 2009 2012 2015 2003 2006 2009 2012 2015
Year Year Year

Fig. 2  Observed monthly mean time series of albedo (𝛼 ), cloud frac- (solid lines), clear-sky albedo (dashed lines) and cloud albedo (dotted
tion ( fc), aerosol optical depth (𝜏a), liquid water path (L), effective lines) time series are shown separately, and L time series are shown
radius (re) and droplet number concentration ( Nd ) (corrected for cli- for both MODIS (solid lines) and MAC (dashed lines). Eruptions of
matological seasonal cycle) for three volcanic regions (Iceland, Vanu- Kilauea (Hawaii) and Eyjafjallajökull and Holohraun (Iceland) are
atu, Hawaii), two anthropogenically influenced regions (China, US) marked with grey in the left-most column. Corresponding time series
and five subtropical stratocumulus regions (Californian, Peruvian, for MERRA-2 (with SO 4 rather than Nd and re) are shown in Supple-
Austalian Namibian, Canarian) as marked in Fig. 1. Total albedo mentary Figure S2

MERRA-2 explicitly includes eruptive volcanoes up until while lack of correlation can be indicative of a theoretical
the end of 2010, but not later, and for degassing volcanoes, causal relation not being strong enough to protrude on
the pattern from year 2010 is repeated for the years fol- the given scale. Spatial correlations, i.e. the correlations
lowing (Randles et al. 2016). Hence, in MERRA-2 data, a between temporally averaged latitude-by-longitude maps
Holohraun effect is expected to be seen only as a response for each region may also be of interest, but more difficult
to assimilated 𝜏a observations, whereas Eyjafjallajökull and to separate from meteorologically driven co-variability,
Kilauea aerosol emissions are explicitly included in the in terms of persisting cloud regime differences within
model. As in the observations, however, the effects are too the regions chosen. For instance, spatial anti-correlations
localized in space and time to show in the monthly mean between aerosol and cloud properties in coastal regions
regional mean time series (Figure S2). may reflect near-land areas being more polluted (high Nd )
and having thinner clouds (low L) because of local mete-
3.2 Correlations orology and transport respectively (Grosvenor et al. 2017).
Observations Figure 3 shows correlation matrices for
We focus here on temporal correlations, i.e. the correla- observed monthly mean time series of 𝛼 , 𝛼clear , 𝛼cloud , fc ,
tions between regional mean time series, corrected for a 𝜏a , L, re and Nd for each of the ten regions. As discussed in
climatological seasonal variation. We emphasize again Sect. 2.4 𝜏a, Nd and L from MODIS are excluded for months
that correlation does not imply causation, and we also November through February for the Iceland region.
emphasize that correlations may be due to meteorological The observed correlations among macrophysical quanti-
co-variation rather than driven by microphysics or cloud ties are largely in agreement with expectations from pre-
dynamics. The temporal correlation coefficients can indi- viously established relations. The correlation between
cate long-term co-variability as well as co-variations on albedo and cloud fraction, R(𝛼, fc ), is strong and positive,
shorter time scale that are frequent and uniform enough confirming the first order dependence of albedo on cloud
to create a signal in the monthly mean regional average, fraction (Cess 1976; Loeb et al. 2007; George and Wood

13
Aerosol-cloud-radiation correlations 4379

Iceland Vanuatu Hawaii


Nd Nd Nd
re re re
LMAC LMAC LMAC
L L L
MODIS MODIS MODIS
τa τa τa
f f f
αc αc αc
cloud cloud cloud
α clear α clear α clear
α α α
α α α f τ L L r N α α α f τ L L r N α α α f τ L L r N
cle cloc a M M e
O AC d cle cloc a M M e
O AC d cle clo c a M M e
O AC d
ar ud DI ar ud DI ar ud DI
S S S
China US
Nd Nd
re re
LMAC LMAC
L L
MODIS MODIS
τa τa
f f
αc αc
cloud cloud
α α
clear clear
α α
α α α f τ L L r N α α α f τ L L r N
cle cloc a M M e
O AC d cle cloc a M M e
O AC d
ar ud DI ar ud DI
S S
Californian Peruvian Australian
Nd Nd Nd
r r r
e e e
LMAC LMAC LMAC
LMODIS LMODIS LMODIS
τa τa τa
fc fc fc
α α α
cloud cloud cloud
α α α
clear clear clear
α α α
α α α f τ L L r N α α α f τ L L r N α α α f τ L L r N
cle cloc a M M e
O AC d cle cloc a M M e
O AC d cle clo c a M M e
O AC d
ar ud DI ar ud DI ar ud DI
S S S
Namibian Canarian 1
N N
d d
r r 0.6
e e
LMAC LMAC
L L 0.2
MODIS MODIS
τa τa
f f -0.2
c c
α cloud α cloud
α α -0.6
clear clear
α α -1
α α α f τ L L r N α α α f τ L L r N
cle cloc a M M e
O AC d cle cloc a M M e
O AC d
ar ud DI ar ud DI
S S

Fig. 3  Observational correlation-matrices for albedo (𝛼 ), cloud frac- 2003–2015, with data from CERES, MODIS and MAC, respectively,
tion ( fc), aerosol optical depth (𝜏a), liquid water path (L), effective are shown. Correlations significant at 95% confidence level (using a
radius (re) and droplet number concentration ( Nd ). Temporal corre- student t-test) are marked with an asterisk
lations of monthly mean, de-seasonalized time series for the period

13
4380 F. A.-M. Bender et al.

2010; Bender et al. 2011; Engström et al. 2015). R(𝛼, fc ) re, correlations are also generally weak in these volcanic and
is stronger than R(𝛼, 𝛼cloud ) or R(𝛼, 𝛼clear ) for instance, the anthropogenic aerosol regions, but Hawaii, China and US
former significantly positive for all regions, and the latter show positive correlations R(L, re ) for one of the L data sets.
significant at the 95% level only in one region. 𝛼clear has a Turning to the Sc regions, Fig. 3 shows that R(L, re )
significant positive correlation with 𝜏a in all regions except is consistently positive and R(L, Nd ) is negative, with the
the Australian region. exception of the Australian region where R(L, Nd ) is not sig-
The importance of L for 𝛼cloud in Sc regions, discussed nificant, and the Canarian region that shows a significant
e.g. by Bender et al. (2016) and Frey et al. (2017), is con- negative R(L, Nd ) only for LMAC . The negative R(L, Nd) and
firmed by positive correlations R(L, 𝛼cloud ) in all five Sc positive R(L, re ) in these regions is contrary to the expecta-
regions. R(L, 𝛼cloud ) is further positive for all regions except tion from a dominating 2nd indirect effect. Wood (2007)
with LMAC in the US region. 𝛼cloud is here calculated based has shown that the suppressed precipitation is in competi-
on the method derived in Bender et al. (2011) utilizing the tion with enhanced entrainment in the Sc regions, and found
near-linear dependence of 𝛼 on fc on regional monthly mean that the sign of the L response to aerosol is dependent on
scale. environmental conditions including ambient humidity and
The co-variation between satellite-retrieved 𝜏a and fc dis- stability as well as time scale. Other previous studies that
cussed in detail e.g. by Quaas et al. (2009), Gryspeerdt et al. have sought a causal explanation for a negative co-variation,
(2016) and Christensen et al. (2017), but also Bender et al. such as that found here, have suggested competing effects of
(2016) and Neubauer et al. (2017), that is likely to be partly precipitation on cloud-top entrainment and sub-cloud mixing
spurious, is manifest as positive R(𝜏a , fc ) in all but one of the (Stevens and Feingold 2009), and increased cloud top cool-
Sc and volcanic regions, but not in the anthropogenically ing in optically thicker clouds leading to enhanced entrain-
influenced regions. ment (Neubauer et al. 2017).
Not unexpectedly, correlations involving microphysi- Replacing LMODIS with LMAC apparently changes the
cal parameters are more difficult to interpret. At constant picture only slightly for the Sc regions, whereas the differ-
L, Nd can be expected to be closely negatively correlated ences are somewhat more noticeable for the volcanic and
with re. This is confirmed by temporal correlations between anthropogenically influenced regions. Figure 3 also gives an
MODIS-derived re and Nd , for all regions except China. On indication of the general level of agreement between the two
the other hand, R(𝛼cloud , Nd ), to which the 1st indirect effect L data sets; R(LMODIS , LMAC ) is positive in all regions with
is expected to contribute positively (more droplets leading values of 0.5 and above.
to brighter clouds, given a constant L), is typically weak, The emerging correlations are not primarily determined
significantly positive only in the Iceland, US and Peruvian by the presence of volcanic aerosol during parts of the
regions. This is in agreement with the suggestion of a weak period studied. Exclusion of the months with volcanic erup-
signal from the 1st indirect effect in satellite observations, tions, marked in the time series in Fig. 2, makes marginal
made by Bender et al. (2016); there are simply too many difference to the correlation matrices; there are no changes
other factors affecting cloud albedo for the influence of Nd of correlation signs in the affected regions, supplementary
to come through on this scale. The link between 𝜏a and Nd Figure S3), although, the positive correlation R(acloud , Nd )
that is implicitly assumed by e.g. Bender et al. (2016) and and R(𝜏a , 𝛼clear ) in Iceland, is lost when volcano-years are
found on global scale by Quaas et al. (2009), is also not con- excluded.
spicuous from the monthly and regional mean observational Reanalysis Correlation matrices based on MERRA-2 rea-
correlation matrix; R(𝜏a , Nd ) is positive for non-Sc regions, nalysis are displayed in Fig. 4, and show some interesting
but not significant in Sc regions. This illustrates the influ- differences from the observational correlations. Contrary to
ence of e.g. updrafts, cloud processes and humidity on Nd the satellite observations and the findings by Bender et al.
and 𝜏a variation. (2016) and Frey et al. (2017), R(L, 𝛼cloud ) is not significantly
For a dominating 2nd indirect effect, considered in isola- positive in any of the ten regions, but actually negative for
tion, a positive correlation between Nd and L is expected Iceland, Vanuatu, China, Peruvian and Namibian. R(L, fc ) is
(more droplets leading to less rain and more L). However, also in contrast with observations negative for all regions.
none of the volcanic regions (Iceland, Vanuatu, Hawaii) This bias does not appear for MERRA-2 grid-box average
show a significant correlation, which is consistent with the L and hence these discrepancies involving L may be related
findings of Malavelle et al. (2017) that L is insensitive to to biases in the MERRA-2 fc , used to calculated in-cloud L.
Nd variations in these regions. Correlations between L and R(𝜏a , fc ) is not pronounced here as in the satellite obser-
Nd are also mostly non-significant for the regions affected vations; the correlations are in some cases positive and in
by anthropogenic aerosol (China, US); the US region dis- some cases negative, consistent with this correlation stem-
plays a weakly positive R(L, Nd ) for LMODIS , while China ming from the near-cloud aerosol swelling, that biases
shows a weakly negative correlation with LMAC . For L and

13
Aerosol-cloud-radiation correlations 4381

Iceland Vanuatu Hawaii


S0 S0 S0
4 4 4
L L L
τ τ τ
a a a
f f f
αccloud αccloud αccloud
α α α
clear clear clear
α α α
α α α f τ L S0 α α α f τ L S0 α α α f τ L S0
c c c a
le lo 4 c c c a
le lo 4 c c c a
le lo 4
ar ud ar ud ar ud

China US
S0 S0
4 4
L L
τa τa
fc fc
αcloud αcloud
α α
clear clear
α α
α α α f τ L S0 α α α f τ L S0
c c c a
le lo 4 c c c a
le lou 4
ar ud ar d

Californian Peruvian Australian


S0 S0 S0
4 4 4
L L L
τa τa τa
fc fc fc
αcloud αcloud αcloud
α α α
clear clear clear
α α α
α α α f τ L S0 α α α f τ L S0 α α α f τ L S0
c c c a
le lo 4 c c c a
le lou 4 c c c a
le lo 4
ar ud ar d ar ud

Namibian Canarian
S0 S0 1
4 4
L L 0.6
τa τa
fc fc 0.2
α
cloud
α
cloud
-0.2
α α -0.6
clear clear
α α
-1
α α α f τ L S0 α α α f τ L S0
c c c a
le lo 4 c c c a
le lo 4
ar ud ar ud

Fig. 4  Reanalysis correlation-matrices for albedo (𝛼 ), cloud frac- de-seasonalized time series for the period 2003–2015, with data from
tion ( fc), aerosol optical depth (𝜏a), liquid water path (L) and sulfate MERRA-2 are shown. Correlations significant at 95% confidence
mass concentration (SO 4 ). Temporal correlations of monthly mean, level (using a student t-test) are marked with an asterisk

observations but is corrected for when assimilated into the Nd is not available from MERRA-2, and as described in
reanalysis (Randles et al. 2016; Chin et al. 2002). Sect. 2.2, we use MERRA-2 SO4 as proxy for the droplet
number concentration. SO4 is positively correlated with 𝜏a

13
4382 F. A.-M. Bender et al.

for all regions, but R(𝛼cloud , SO4 ), to be compared with the by Bender et al. (2016), but it is clear that this relation does
observed R(𝛼cloud , Nd ), is generally weak. Correlations are not necessarily come through with the simple correlation
significantly positive only in the Iceland region, while nega- approach taken, as many factors, not least L, influence the
tive for Hawaii and China. Hence, as in the case of satellite co-variation. Even for R(𝛼cloud , 𝜏a ), that offers a more direct
observations, the correlations do not support a dominating comparison with Bender et al. (2016), negative values are
1st indirect effect on this temporal and spatial scale. With no seen for the Namibian region in IPSL-CM5A-MR, for the
parameterized coupling between aerosol and cloud micro- US region in IPSL-CM5B-LR, the Namibian region in
physics in MERRA-2 this insensitivity to aerosol concentra- MIROC5, the Vanuatu region in MIROC-ESM and China
tion is not unexpected. and the Australian region in MRI-fixed.
Similarly, R(L, SO4 ) is generally weak. Significant nega- The procedure in Bender et al. (2016), accounting for
tive correlations are seen in the Iceland and Australian both seasonal and regional co-variability and investigating
regions, and positive correlations in the China and Vanuatu 𝜏a-related albedo-variation at a given cloud fraction can bet-
regions. Hence, taking SO4 as a proxy for Nd , the obser- ter distil the actual aerosol-𝛼cloud coupling than mere cor-
vational negative correlations R(L, Nd ) in several of the Sc relations. Another caveat is that here we are using monthly
regions are not reproduced in the MERRA-2 data set. Again, mean data to calculate a monthly resolved cloud albedo,
lacking a parameterized aerosol–cloud coupling in MERRA- rather than producing an average value for a longer period,
2, a weak correlation between aerosol ( SO4 ) and cloud (L) as originally done by Bender et al. (2011). Each regional
is not unexpected. mean cloud albedo value is hence based on albedo and cloud
Excluding specific volcano-months (see markings in fraction observations in as many data points as there are
Fig. 2 time series) makes marginal difference with MERRA- grid boxes in the region, i.e. significantly fewer points than
2. Negative correlations between L and 𝛼 and 𝛼cloud respec- available for a temporal mean 𝛼cloud calculation. Further, the
tively are lost when specific volcano-affected months are condition of linearity between albedo and cloud fraction, and
excluded (Figure S3). hence the cloud albedo estimate, is most reliable in the Sc
Models We now compare the observational and reanaly- regions, cf. Bender et al. (2011).
sis based correlation matrices with those from the CMIP5 For consistency with MERRA-2 we also add in SO4 in
models listed in Table 2. Correlation matrices for each of the the model correlation matrices. We note that the sulfate
models are found in Supplementary Figures S4–S11. mass measures are not the same in models and reanalysis;
Differences from observations are particularly related to MERRA-2 SO4 represents mass concentration ( μgm−3 )
microphysical quantities, and specifically include the rela- and CMIP5 SO4 the integrated mass loading (kg m−2 ), but
tions R(Nd , re ), R(acloud , Nd ) and R(L, Nd ). surface and column sulfate can be expected to be tightly
Contrary to observations, R(Nd , re ) is strongly positive for correlated. R(SO4 , Nd ) is typically positive, in models with
all regions in IPSL-CM5A-LR, IPSL-CM5A-MR and IPSL- diagnostic as well as prognostic Nd , and so is R(𝜏a , SO4 ).
CM5B-LR. For these models re is a function of aerosol mass The correlation R(L, Nd ) that in observations is found to
concentration (Hourdin et al. 2013). In the MRI models on be negative for Sc regions, and in other regions weaker and
the other hand, R(Nd , re ) is negative throughout when sig- varying in sign, varies among models as well. For the IPSL
nificant, and in the MIROC models slightly more variety is models, which in contrast with the other studied models do
seen. These model families, as opposed to the IPSL models, not include the 2nd indirect effect, negative correlations
parameterize re as a function of droplet number concentra- appear in the Californian region (CM5A-LR, CM5A-MR),
tion and liquid water content, which gives a more realistic re and positive correlations are seen for Hawaii and Peruvian
estimate. The geographical distribution of Nd is in general (CM5A-LR, CM5A-MR, CM5B-LR), Iceland (CM5A-
agreement among all the models, but the IPSL models are LR, CM5B-LR), China (CM5A-LR, CM5A-MR) and US,
the only ones that display a global scale positive correla- Namibian, Canarian regions (CM5B-LR).
tion R(Nd , re ) (not shown), rendering their re estimates less MIROC5 produces negative R(L, Nd ) in the Californian,
trustworthy. Australian and Canarian regions, and also for the Iceland,
R(𝛼cloud , Nd ) is in many cases positive when significant, Vanuatu and Hawaii regions. MIROC-ESM on the contrary
but there are several exceptions. MIROC5 shows nega- has weaker correlations, significantly positive only for the
tive R(𝛼cloud , Nd ) for the Vanuatu, Hawaii, Californian and Vanuatu, Hawaii and Californian regions. MRI-CGCM
Namibian regions, and IPSL-CM5B-LR for Iceland and shows significant negative correlations in the Vanuatu and
Australian and IPSL-CM5A-MR for Californian, Austral- US regions, and positive R(L, Nd ) in the Namibian and
ian and Iceland and IPSL-CM5A-LR for the Californian, Canarian regions. This is similar to MRI-ESM1 which has
Australian and Namibian regions. From the parameterized negative R(L, Nd ) in Vanuatu and US, and positive R(L, Nd )
cloud albedo effect in the models, a positive correlation in the Peruvian, Namibian and Canarian regions. For MRI-
between Nd and 𝛼cloud is expected, as was also confirmed fixed the picture is somewhat changed; correlations are

13
Aerosol-cloud-radiation correlations 4383

negative for the Hawaii, US, Californian and Australian to be calculated. This can be compared with the correlation
regions. matrices (Sect. 3.2), where larger regional means are used,
and correlations can be defined even in areas that are here
screened out.
3.2.1 Global distribution of R(L, Nd ) Following McCoy et al. (2017), we use MERRA-2 SO4
to derive a global gridded Nd-approximation. The temporal
The variation in R(L, Nd ) between the selected regions sup- correlation between this Nd-proxy ( NdMERRA) and L derived
ports previous studies that have indicated that the suscepti- from MODIS is shown in Fig. 5b. The general pattern of
bility of cloud water to changes in aerosol, dln(L)∕dln(Nd ), negative correlations at low latitudes and positive correla-
varies between geographical regions (Quaas et al. 2009) in tions at high latitudes are in agreement with the results of
response to variations in environmental regimes in terms Michibata et al. (2016) for both precipitating and non-pre-
of, for example, humidity, stability and the occurrence of cipitating clouds, with Nd calculated from gridded MODIS 𝜏
precipitation (Michibata et al. 2016; Neubauer et al. 2017). and re. As in the case of L and Nd both derived from MODIS
We here look further at the R(L, Nd ) and its geographical (Fig. 5a) there is, however, indication that the Sc regions,
distribution and relation to dynamical regime. display negative correlations.
Figure 5a shows that with MODIS Nd and L, R(L, Nd ) In Fig. 5c, the correlation distribution is shown for
is positive in large areas, especially at mid-latitudes, while N dMERRA and LMAC , but the MAC L is in this case grid-aver-
near-coast areas, particularly the subtropical Sc regions, aged rather than in-cloud, i.e. L is not divided by fc , which
display negative correlations, as also seen in the correla- means it is sensitive to changes in cloud coverage as well as
tion matrices (Fig. 3). The MODIS Nd has limited spatial cloud thickness. As seen earlier, the conversion to in-cloud
coverage due to the restriction to fc >80% and unreliable L gave rise to some biases in the comparison with obser-
retrievals at high 𝜃 (see Sect. 2.1). Here we set a threshold vations. The LMAC estimate also includes broken or scat-
value so that at least 70% of the points in the monthly mean tered clouds in which the MODIS cloud property retrieval
time series at each grid point must be valid, for a correlation

R(LMODIS ,NdMODIS ) R(LMODIS ,NdMERRA)


(a) 90 ° N (b) 90 ° N

° °
45 N 45 N

° °
0 0
°
180 W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E °
180 W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E

45 ° S 45 ° S

° °
90 S 90 S

(c) R(LMAC ,NdMERRA)


(d) R(LMERRA,NdMERRA)
° °
90 N 90 N

° °
45 N 45 N

° °
0 0
° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° °
180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E

45 ° S 45 ° S

90 ° S 90 ° S

-1 -0.6 -0.2 0.2 0.6 1


R(L,N d)

Fig. 5  Correlation L and Nd a in-cloud L and Ndfrom MODIS, b in-cloud L from MODIS and Nd derived from MERRA-2 ­SO4, c grid-average L
from MAC and Nd derived from MERRA-2 ­SO4, d grid-average L from MERRA-2 and Nd derived from MERRA-2 ­SO4

13
4384 F. A.-M. Bender et al.

°
IPSL-CM5A-LR °
IPSL-CM5A-MR °
IPSL-CM5B-LR
90 N 90 N 90 N

° ° °
45 N 45 N 45 N

0° 0° 0°
° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° °
180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E

° ° °
45 S 45 S 45 S

° °
90 S 90 S 90 ° S

MIROC5 MIROC-ESM
90 ° N 90 ° N
1
° °
45 N 45 N
0.6

° ° 0.2
0
180 ° W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E
0
180 ° W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E
R(L,N )
d
-0.2
45 ° S 45 ° S
-0.6
° °
90 S 90 S
-1

°
MRI-CGCM3 °
MRI-ESM °
MRI-fixed
90 N 90 N 90 N

45 ° N 45 ° N 45 ° N

0° 0° 0°
180 ° W 135 ° W °
90 W
°
45 W 0
° °
45 E
°
90 E
°
135 E 180 E
° ° °
180 W 135 W
°
90 W
°
45 W 0
° °
45 E
°
90 E
°
135 E 180 E
° ° °
180 W 135 W
°
90 W
°
45 W 0
° °
45 E
°
90 E
° °
135 E 180 E

° ° °
45 S 45 S 45 S

° °
90 S 90 S 90 ° S

Fig. 6  Spatial distribution of temporal correlation between L and Nd , for 8 CMIP5 models. De-seasonalized monthly means for 10 years are used
for each model. As in Figs. 5c, d, L is here given as grid box average

is limited, but the picture for R(LMAC , NdMODIS ) is consistent in their Nd -parameterization, and for comparison we also
with R(LMODIS , NdMERRA ) (Fig. 5b). show results for MRI-fixed. It is clear that this reduces the
Figure 5d shows the correlation between MERRA- strong sensitivity of L to Nd in the Sc regions, and in terms
derived Nd and MERRA grid-averaged L, and it is clear that of magnitude of correlation coefficient, MRI-fixed shows
compared to the observational and combined observation- the best general agreement with observations, whereas the
reanalysis distributions of R(L, Nd ) , the pure reanalysis other models typically overestimate the magnitudes (the
correlations are weaker and show a much less pronounced colour scale is the same for Figs. 5 and 6.) IPSL-CM5A-
pattern. As before, MERRA-2 can be seen as a “no-indi- LR and IPSL-CM5A-MR are as expected very similar,
rect-effect” reference, and hence this difference suggests but IPSL-CM5B-LR shows a quite different pattern with
that meteorological factors alone, without microphysical positive correlations between L and Nd nearly everywhere.
coupling between aerosols and clouds, are not sufficient for Neither of the IPSL models include the 2nd indirect effect,
producing the observed relations between L and Nd. and hence in these cases the correlation between Nd and
The CMIP5 models display great variation in geographi- L is not driven by a lifetime effect, but is affected by the
cal distribution of R(L, Nd ), see Fig. 6. Several of the mod- differences in physics that distinguish the IPSL-CM5A and
els show a region of strong positive correlations over the IPSL-CM5B models (see Sect. 2.3). MIROC5 in contrast
Southern Ocean, leading to a much more pronounced lati- shows largely negative correlations, with exceptions of the
tudinal gradient than in observations. In the IPSL models high latitude oceans and regions of continental outflow, like
and MIROC-ESM strong positive correlations are also found China. Michibata et al. (2016) show that MIROC5 has a
over the tropics, in a way that is not supported by observa- positive L to Nd susceptibility more or less globally, but this
tions. With the exception of the two MIROC models, par- result can only be reproduced if cloud-top Nd is replaced
ticularly MIROC5, the observational pattern of negative with vertically integrated Nd (see Sect. 1 for CMIP5 variable
correlations over the subtropical stratocumulus regions is definitions), in which case variations in cloud-geometrical
reversed, with pronounced positive correlations over these thickness result in a positive correlation between Nd and L
regions in the IPSL and MRI models. As stated previously, globally. MIROC5 here also shows some indication of local
MRI ESM and MRI CGCM3 have a documented error negative maxima coinciding with Sc regions, which is in line

13
Aerosol-cloud-radiation correlations 4385

0.8
MODIS,MERRA
MAC,MERRA
MERRA,MERRA
0.6 IPSL-CM5A-LR
IPSL-CM5A-MR
IPSL-CM5B-LR
MIROC5
0.4 MIROC-ESM
MRI-CGCM3
MRI-ESM
0.2 MRI-fixed
R(L,Nd)

-0.2

-0.4

-0.6
-0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08
-1
ω [Pas ]
500

Fig. 7  R(L, Nd ) for global observations, reanalysis and CMIP5 models Coloured lines relate model L to model Nd , binned by model 𝜔500.
as a function of vertical velocity. Solid black line relates MODIS L to Correlations have been averaged in bins of width 0.02 Pas−1, and are
MERRA-2 Nd , dashed lack line MAC L to MERRA-2 Nd and dotted displayed excluding bins containing less than 0.15% of the total num-
black line MERRA-2 L to MERRA-2 Nd , all binned by ERA 𝜔500. ber of cases

with the observational pattern. Although the two MIROC The CMIP5 models typically show stronger correlations,
models share the same aerosol model, their separate cloud when binned in the same way by the model vertical velocity
schemes give rise to pronounced differences in the relation at 500 hPa. None of the models reproduce the weak correla-
between L and Nd. tions that observations and reanalysis indicate, and the pat-
To investigate a possible dependence of aerosol–cloud terns are also quite different from that seen in observations.
interaction on meteorological regime, adding to the findings In agreement with Zhang et al. (2016) who studied L
of environmental regime dependence by Michibata et al. sensitivity to concentration of cloud condensation nuclei in
(2016), Zhang et al. (2016) and Neubauer et al. (2017), we a different set of GCMs, the model diversity is largest in
investigate the dependence of R(L, Nd ) to vertical velocity regions of strong ascent (negative 𝜔500 ), where correlations
in the upper troposphere. range from strongly positive (IPSL-CM5A) to strongly nega-
Figure 7 shows the correlation R(L, Nd ) as a function of tive (MIROC5). Aerosol-cloud interaction in convection is
vertical velocity at 500 hPa, 𝜔500 . Relating MODIS and not represented in the models, and to the extent Nd-L rela-
MAC L respectively to MERRA Nd , R(L, Nd ) is weakly tions in these clouds are driven by such interactions a certain
positive at moderate ascent and subsidence. Stronger dis-agreement with observations is expected. As biases are
ascent around the Intertropical Convergence Zone (ITCZ) both negative and positive it is not clear in which direction
coincides with negative correlations, which may be an the lack of aerosol-aware convection affects the correlation.
effect of efficient aerosol removal by precipitation. For It is interesting to note the close agreement between the
these cases, however, McCoy et al. (2017), McCoy et al. two IPSL-CM5A models, that differ only in resolution, and
(2018) do not offer a validation of the relation between their difference from the third IPSL model, that has differ-
SO4 and Nd , and they are also less frequently occurring ent model physics, including convection representation.
than the large areas of weak ascent and positive corre- Noteworthy is also the agreement between the MRI mod-
lations over the Southern Ocean. Similarly, subsidence els (MRI-CGCM3, MRI-ESM1 and MRI-fixed), where the
regions (including the Sc regions) partly coincide with correction of the Nd-related error appears to lead to some-
negative correlations for both MODIS and MAC L. For what weaker R(L, Nd ), especially in subsidence regions. The
most cases of stronger subsidence (corresponding to two MIROC models on the other hand have quite different
areas in the subtropics and mid-latitudes, of which the Sc behaviours. While MIROC-ESM largely shows positive
regions are only a small part), correlations go to zero for correlations, MIROC5 almost exclusively shows negative
MODIS but are positive for MAC L. correlations, although both MIROC models have cloud
The R(L, Nd ) for MERRA-2 are overall small, suggesting schemes including a dependence of autoconversion rate on
that L-Nd correlations arising from meteorology alone, with the droplet spectrum. The IPSL models on the other hand
no indirect effects will introduce only small uncertainties in do not have a parameterized dependence of rain rate on Nd
the correlations. (2nd indirect effect), and the relatively strong correlations
at all 𝜔500-values in IPSL-CM5B-LR and at the limits of the

13
4386 F. A.-M. Bender et al.

°
IPSL-CM5A-LR °
IPSL-CM5A-MR °
IPSL-CM5B-LR
90 N 90 N 90 N

45 ° N 45 ° N 45 ° N

0° ° 0° ° 0° °
180 W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E °
180 W 135 W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E °
180 W 135 W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E

° ° °
45 S 45 S 45 S

° °
90 S 90 S 90 ° S

°
MIROC5 °
MIROC-ESM
90 N 90 N
1
° °
45 N 45 N
0.6

0° ° 0° ° 0.2
° ° ° ° ° ° ° ° ° ° ° ° ° ° ° ° R(SO ,N )
180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 4 d
-0.2
45 ° S 45 ° S
-0.6
90 ° S 90 ° S -1

MRI-CGCM3 MRI-ESM MRI-fixed


90 ° N 90 ° N 90 ° N

45 ° N 45 ° N 45 ° N

0° ° ° ° ° ° ° ° ° °
0° ° 0° °
180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 ° W °
90 W
°
45 W 0
° °
45 E
°
90 E
°
135 E 180 E
°
180 W 135 ° W °
90 W
°
45 W 0
° °
45 E
°
90 E
°
135 E 180 E
°

° ° °
45 S 45 S 45 S

90 ° S 90 ° S 90 ° S

Fig. 8  Spatial distribution of temporal correlation between SO 4 and Nd , for 8 CMIP5 models. De-seasonalized monthly means for 10 years are
used for each model

displayed 𝜔500 range for IPSL-CM5A cannot be ascribed to observations is not particularly improved, although the elim-
cloud lifetime effects. ination of Nd dampens the spuriously strong L-sensitivity
in the Sc regions in MRI-ESM1 and MRI-CGCM. For the
3.2.2 Global distribution of R(SO4 , Nd ) IPSL models, where Nd does not respond to SO4 changes,
and R(L, SO4 ) is weak.
McCoy et al. (2017) and McCoy et al. (2018) have demon-
strated the usefulness of SO4 as a predictor of Nd , in a range
of regions, and on daily to decadal time scale. The correla- 4 Discussion
tion between SO4 and Nd , combining satellite observations
and MERRA-2 reanalysis is hence positive. Among the 4.1 Model representation of Nd
CMIP5 models we also find R(SO4 , Nd ) to be predominately
positive in the ten focus regions. Figure 8 shows the global Nd is a central parameter in aerosol–cloud interactions, but
distribution of this correlation, indicating that positive corre- in global models, its representation and relation to other
lations dominate globally for the MIROC and MRI models, variables is quite varying. Although Nd is mostly positively
the most striking exception being the negative correlations correlated with SO4 (supporting McCoy et al. (2017)) in the
at 30◦ S in MIROC5. There is a clear distinction between models where Nd is a prognostic function of aerosol and
these model families, that have prognostic Nd , and the IPSL supersaturation, this is not the case for models with diag-
models with diagnostic Nd , where R(SO4 , Nd ) is weak and nostic Nd . Further, the susceptibility dlog(Nd )∕dlog(SO4 )
varying in sign. The corresponding linear regression coef- for column integral SO4 varies largely between models and
ficient of the relation dlog(Nd )∕dlog(SO4 ), as quantified in between regions, in contrast with what McCoy et al. (2017)
McCoy et al. (2017), is shown in Figure S12. showed for Nd from satellite observations and surface level
In the MIROC and MRI models, Nd further affects the SO4 mass concentration from reanalysis.
autoconversion rate implying a link between SO4 and L, and Technically, the model Nd is also problematic to compare
Fig. 9 displays the geographical distribution of R(L, SO4 ) with observations. While a few models supply temporally
that for these models can be compared to the observational evolving 3D-fields of Nd , others provide only a vertically
R(L, Nd ) in Fig. 5. The agreement between models and integrated droplet number concentration. Here we utilize

13
Aerosol-cloud-radiation correlations 4387

IPSL-CM5A-LR IPSL-CM5A-MR IPSL-CM5B-LR


90 ° N 90 ° N 90 ° N

° ° °
45 N 45 N 45 N

0° ° ° ° ° ° ° ° ° °
0° ° ° ° ° ° ° ° ° °
0° ° ° ° ° ° ° ° ° °
180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E

° ° °
45 S 45 S 45 S

° ° °
90 S 90 S 90 S

°
MIROC5 °
MIROC-ESM
90 N 90 N
1
45 ° N 45 ° N
a


0° 0.2
180 ° W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E 180 ° W 135 ° W 90 ° W 45 ° W 0° 45 ° E 90 ° E 135 ° E 180 ° E R(L,SO4)
-0.2
° °
45 S 45 S
-0.6

90 ° S 90 ° S -1

MRI-CGCM3 MRI-ESM MRI-fixed


90 ° N 90 ° N 90 ° N

45 ° N 45 ° N 45 ° N

0° ° ° ° ° ° ° ° ° °
0° ° ° ° ° ° ° ° ° °
0° ° ° ° ° ° ° ° ° °
180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E 180 W 135 W 90 W 45 W 0 45 E 90 E 135 E 180 E

° ° °
45 S 45 S 45 S

° °
90 S 90 S 90 ° S

Fig. 9  Spatial distribution of temporal correlation between SO 4 and L, for 8 CMIP5 models. De-seasonalized monthly means for 10 years are
used for each model

the droplet number concentration at cloud top, considering supported negative correlations seen in MRI and largely in
liquid clouds only, which is the measure that we consider MIROC models.
most directly comparable to observations, and that is avail- Based on different observational estimates, Quaas et al.
able for the largest number of models, which is, however, (2009) show positive coefficients between Nd and 𝜏a on
limited to seven. Given the importance of Nd , its representa- global scale, but we find that 𝜏a is not positively correlated
tion in models as well as its requirement as an output field in with Nd in all focus regions in MODIS observations. Sig-
model intercomparison projects should be considered more nificant positive correlations are seen for all the volcanic
carefully. and anthropogenic regions except Vanuatu, but for the Sc
Nd has specific issues in the two MRI models MRI- regions, no correlations are significant. On the other hand, in
CGCM3 and MRI-ESM1. In the AMIP simulation with MERRA-2, 𝜏a is consistently positively correlated with SO4,
corrected Nd -parameterization (MRI-fixed) the pattern of which can arguably be used as an Nd proxy. In the models,
R(L, Nd ) is similar to that of the MRI-CGCM3 and MRI- R(𝜏a , SO4 ) is typically positive whereas R(𝜏a , Nd ) varies more
ESM1 but with weaker positive correlations in the sub- in sign in the different regions.
tropics. By-passing the Nd -parameterization, i.e. analyz- Ekman (2014) showed that models with an explicit
ing R(L, SO4 ) instead of R(L, Nd ) has a similar effect in the parameterization of Nd , as a function of aerosol as well
MRI models. In general, however, discrepancies between as supersaturation, better reproduce observed temperature
models and observations in R(L, Nd ) cannot be solely attrib- trends, while the inclusion of aerosol effects on precipita-
uted to mis-representation of the relation between SO4 and tion formation was a less useful determiner of model skill.
Nd ; replacing R(L, Nd ) with R(L, SO4 ) does not particularly We see that the MRI and MIROC models produce a more
improve the agreement with observations. realistic SO4-Nd relation, assuming that the positive correla-
Models typically determine re as a function of Nd and tion between Nd and SO4 described by McCoy et al. (2017)
L, see e.g. Peng and Lohmann (2003). This is the case for is realistic, but otherwise no clear distinction in model skill
the MRI and MIROC models, but in the IPSL models re is can be easily made based on the level of sophistication of
dependent on aerosol concentration, yielding positive cor- aerosol–cloud interactions.
relations between Nd and re, rather than the observationally

13
4388 F. A.-M. Bender et al.

4.2 
R(L, Nd ) as indicator of the 2nd indirect effect schemes, finding high L-sensitivity to the scheme used, but
also pointing at other common factors, including too large
A striking result is that in the Sc regions observed L is rather dependence on autoconversion in diagnostic rain schemes,
consistently negatively correlated with Nd and positively as causing biases. Neubauer et al. (2017) show, by com-
with re , both spatially and temporally. This is contrary to paring a prognostic and a diagnostic precipitation scheme,
what may be expected from a dominance of the 2nd indirect that overestimation of Nd-dependent autoconversion is not
effect (Albrecht 1989) where more, smaller droplets would the primary reason for overestimated L-susceptibility in
prevent rainfall and lead to greater L. their model. Based on the argument that cloud top entrain-
Wood (2007) points at the competition between effects ment competes with suppressed precipitation (Wood 2007),
on precipitation and entrainment of dry air from above in MIROC5 seems to better capture this entrainment process
determining the L-response to aerosol changes in stratiform than MIROC-ESM. Noting that MIROC5 uses fewer verti-
clouds. Stevens and Feingold (2009) have suggested that cal levels than MIROC-ESM (see Table 2) this is not due to
counteracting effects of enhanced cloud-top entrainment better vertical resolution.
and reduced sub-cloud mixing in response to suppressed Even models that do not parameterize the 2nd indirect
precipitation may lead to a net cloud thinning. Negative cor- effect by relating rain formation to droplet number concen-
relations in the given Sc regions may also be related to the tration, in particular IPSL-CM5B, show positive correla-
fact that these clouds typically give rise to little precipita- tions R(LWP, Nd ) in several areas. Hence positive correla-
tion, limiting the effects of precipitation suppression. Nega- tions R(L, Nd ) may have other origins than increased cloud
tive susceptibility of L to aerosol in non-raining regions has lifetime as prescribed by the 2nd indirect effect, related to
been suggested to be related to enhanced entrainment due meteorology or other aerosol–cloud interactions. For exam-
to decreased droplet sedimentation when cloud droplets are ple, Wilcox (2010), Wilcox et al. (2016) propose dynamical
smaller (Bretherton et al. 2007; Small et al. 2009; Chen et al. pathways for increased L at higher loading of absorbing aer-
2015; Neubauer et al. 2017). Neubauer et al. (2017) also osol, through suppressed turbulence in the boundary-layer
suggest that improved representation of cloud-top entrain- and reduced entrainment of dry air from above in response
ment could reduce the L-susceptibility in unstable and/or to the aerosol-induced changes in stability.
dry regions in models. For MERRA-2 there is no parameterized aerosol–cloud
Compared to the observations, the models show more interaction and no Nd estimate. R(L, SO4 ) is used as a proxy
variability and more positive values for the correlation for R(L, Nd ), resulting in correlations that are neither as neg-
between L and Nd , particularly in the Sc regions. The model ative as seen for observations in the Sc regions, or as positive
that best reproduces the observed negative correlations as seen in many cases in the CMIP5 models. This indicates
R(L, Nd ) in Sc regions is MIROC5. This model has the same that in reality microphysical or small-scale dynamical cou-
parameterized dependence of autoconversion rate on drop- plings not included in the reanalysis play a role, but that in
let number concentrations as MIROC-ESM, based on Berry models these processes are not correctly represented.
(1968), but unlike MIROC-ESM, MIROC5 still produces Investigating the dependence of the 2nd indirect effect on
negative correlations in three Sc regions, and also in the ambient meteorological conditions, Neubauer et al. (2017)
three volcanic regions. found that their model is less sensitive to environmental
Quaas et al. (2009) previously questioned the implemen- regimes than what is suggested by observations. Michibata
tation of the lifetime effect only through autoconversion, et al. (2016) found that their model produced positive sus-
as did Michibata et al. (2016). Wang et al. (2012) rightly ceptibility of L to Nd regardless of environmental regime,
emphasize that many cloud dynamical feedbacks that are but this appears to be related to the dependence of both L
not included in climate models affect cloud fraction, water and vertically integrated Nd on cloud geometrical thickness,
content, and lifetime. From our analysis, although the dif- rather than a microphysical coupling between Nd and L. We
ferences in sensitivity between models, observations and find that R(L, Nd ) varies with larger amplitude with varying
reanalysis suggest that models have too strong L– Nd links upper tropospheric vertical velocity in models than in obser-
for the specific studied regions, it is not clear that parameter- vations, but that reanalysis has yet weaker dependence on
ized dependence of autoconversion on Nd directly leads to dynamical regime, again using SO4 to approximate Nd vari-
exaggerated positive R(L, Nd ). ations, and again indicating that the lack of aerosol–cloud
In MIROC5 and MIROC-ESM that share the same aero- coupling in the reanalysis can underestimate the observed
sol scheme and the same autoconversion scheme, quite dif- correlations between L and Nd.
ferent patterns of R(L, Nd ) appear. Hence other differences
between their cloud schemes are more decisive in this rela-
tion than the autoconversion parameterization. Michibata
and Takemura (2015) compare different autoconversion

13
Aerosol-cloud-radiation correlations 4389

5 Concluding remarks R(𝛼cloud , Nd ), which may be seen as an indicator for the


1st indirect effect, is weak in observations, indicating that
We study relations between several macro- and micro- droplet number is not the primary driver of the variation
physical cloud and aerosol properties, in a range of regions in cloud albedo. Given the numerous dynamical and mete-
with different characteristics in terms of meteorology and orological processes acting simultaneously, the 1st indirect
aerosol loading. An ensemble of coupled climate models effect can hardly be expected to be seen as acting in isola-
is compared to satellite observations and reanalysis, where tion. In models, however, R(𝛼cloud , Nd ) is in many cases posi-
the latter assimilates aerosol optical depth as well as mete- tive, consistent with the parameterization of the 1st indirect
orological fields but does not parameterize a microphysical effect. Still, there are also cases of weak and negative cor-
coupling between aerosols and clouds. Hence models can relations, even in the Sc regions where previous studies have
be evaluated against the observed state and a reference case found that the cloud brightening in models is strong (Bender
with no prescribed aerosol–cloud interactions. et al. 2016), and we emphasize that the monthly resolved
Some general differences between categories of regions 𝛼cloud estimate used in the present study must be interpreted
can be seen; e.g. the US and China regions, representing with discretion.
dominance of anthropogenic aerosol, show decreasing trends R(L, Nd ) in observations is positive for large areas of the
in 𝜏a and Nd over the period 2003–2015, and also larger Nd globe, but conspicuously negative for Sc regions, contrary to
and smaller re than the Iceland, Vanuatu and Hawaii regions, the expectation from the 2nd indirect effect, and this is not
representing influence from volcanic aerosols. The Iceland well captured by models. Counteracting effects of entrain-
region is different from the two other volcanic regions as ment in response to aerosol variations may explain the
degassing does not provide a strong source of SO2 , and negative correlations, and meteorologically driven co-vari-
the region is not as remote, and hence more influenced by ation between aerosol and cloud properties always remains
anthropogenic aerosol than Vanuatu and Hawaii. a possible cause of the relations seen (see e.g. Toll et al.
Our observational data analysis confirms previous find- 2017), but still model-produced relations in many ways dif-
ings of decreases in re in association with an eruptive vol- fer from those observed. The global geographical distribu-
canic event in Iceland (McCoy and Hartmann 2015; Mala- tion of R(L, Nd ) is in poor agreement with observations for
velle et al. 2017). At the time of the Holohraun eruption in all models, and the variation in R(L, Nd ) with dynamical
2014–2015, satellite observations also indicate large local regime is much stronger in models than in observations and
variations in L, but their spatial and temporal extents are reanalysis. Hence we can make a more general comparison
not in agreement with the peak emission time and transport of sensitivity of cloud water to aerosol amount, in models,
direction from the eruption, and neither MODIS nor MAC observations and reanalysis. Both geographical distribution
indicate a regional mean L-signal co-incident with the emis- of R(L, Nd ) and R(L, Nd ) as a function of vertical velocity
sion of volcanic aerosol. The re-anomalies too are spatially give a picture of reanalysis (no parameterized aerosol–cloud
and temporally confined, and the cloud property changes interaction) showing weak correlations, models (parameter-
do not result in discernible signals in 𝛼cloud or 𝛼 . Temporal ized aerosol–cloud interaction) strong correlations, and sat-
correlations in the volcanically influenced regions are found ellite observations somewhere in-between. In other words, it
to be insensitive to the inclusion or exclusion of specific appears that microphysical aerosol–cloud interactions con-
periods affected by volcanic eruptions. tribute to the relations seen on larger scale, but that in mod-
In observations, several correlation combinations of els these relations may be too dominant. We do not, however,
macrophysical properties confirm expected relations among find support for this being caused directly by a too strong
cloud, aerosol and radiation quantities, with which models link between L and Nd through autoconversion, as suggested
can be compared. For instance, observations show that fc by e.g. Quaas et al. (2009) and Michibata et al. (2016). Mod-
variability is closely related to variation in 𝛼 , and that L co- els without autoconversion dependence on Nd (three IPSL
varies with 𝛼cloud . Models typically agree with observations models) create positive correlations R(L, Nd ) and models
in these cases, yielding positive R(fc , 𝛼) and R(L, 𝛼cloud ) with with the same parameterization of that link (two MIROC
few exceptions. models) create different patterns of R(L, Nd ). Along these
Observational fc and in-cloud L are positively correlated lines, Michibata and Takemura (2015) also suggest that not
in all regions, whereas the calculated in-cloud L in reanalysis only microphysics parameterizations need to be improved
and models is in several cases negatively correlated with fc . for models to better represent the relations between clouds,
Correlations are generally not sensitive to the choice of L precipitation and radiation.
(MAC or MODIS), particularly in the Sc regions where both Our results with consistent non-zero correlations R(L, Nd )
can robustly perform retrievals, and the two L data sets are for the Sc regions also challenge the view that L should not
consistently positively correlated in all regions. respond to aerosol at all (Malavelle et al. 2017). Co-varia-
tions are clearly region-dependent, and more work remains

13
4390 F. A.-M. Bender et al.

to be done to understand the counteracting processes on Berry EX (1968) Modification of the warm rain process. In: Paper
water content, and to what extent they dampen the micro- presented at 1st National Conf. on Weather Modification, April
28 May 1. Am. Meteorol. Soc., Albany, New York. pp 81–85
physical response to aerosol, in different aerosol and mete- Boucher O, Lohmann U (1995) The sulfate-CCN-cloud albedo effect:
orological regimes. A sensitivity study with two general circulation models. Tellus
Although representation of large scale effects of aero- 47B:281–300
sol–cloud interactions in agreement with observations is a Boucher O et al (2013) Clouds and aerosols. In: Stocker TF, Qin D,
Plattnes G-K, Tignos M, Allen SK, Boshung J, Naules A, Xia
necessary condition for a trustworthy model, it is not suf- Y, Bex V, Midgley PM (eds) Climate change 2013: the physical
ficient for constraining model estimates of aerosol forcing, science basis. Contribution of working group I to the fifth assess-
not least due to the vast number of possible model configura- ment report of the intergovernmental panel on climate change.
tions that can pass the test of agreeing sufficiently well with Cambridge University Press, Cambridge
Brenguier J-L et al (2000) Radiative properties of boundary layer
PD observations (Lee et al. 2016; Johnson et al. 2018). The clouds: droplet effective radius versus number concentration. J
presented correlation-based approach is a method for model Atmos Sci 57:803–821
evaluation and points out variations in performance among Bretherton CS, Blossey PN, Uchida J (2007) Cloud droplet sedimenta-
the tested models. tion, entrainment efficiency, and subtropical stratocumulus albedo.
Geophys Res Lett 34:L03813
It is clear from our results, as well as from previous stud- Carlsaw KS et al (2013) Large contribution of natural aerosols to
ies, that individual aerosol–cloud interaction pathways do uncertainty in indirect forcing. Nature 507:67–71
not dominate the relations between variables across regional Carn SA, Fioletov VE, McLinden CA, li C, Krotkov NA (2017) A
and temporal scales (Stevens and Feingold 2009; Rosen- decade of global volcanic SO2 emissions measured from space.
Sci Rep 7:44095
feld et al. 2014; Bender et al. 2016). Our contribution here Cess RD (1976) Climate change: An appraisal of atmospheric feed-
is an evaluation of emerging correlations, microphysics, back mechanisms employing zonal climatology. J Atmos Sci
macrophysics and meteorology taken together, in satellite 33:1831–1843
observations, in reanalysis with no modelled aerosol–cloud Chen Y-C, Christensen MW, Stephens GL, Seinfeld JH (2015) Satel-
lite-based estimate of global aerosol-cloud radiaitve forcing by
interactions, and in global models, and thereby a way of marine warm clouds. Nat Geosci 7:643–646
testing models and pointing at ways of improving their Chin M et al (2002) Tropospheric aerosol optical thickness from the
performance. GOCART model and comparisons with satellite and sun photom-
eter measurements. J Atmos Sci 59:461–483
Cho H-M et al (2015) Frequency and causes of failed MODIS cloud
property retrievals for liquid phase clouds over global ocean. J
Open Access This article is distributed under the terms of the Crea- Geophys Res 120(9):4132–4154
tive Commons Attribution 4.0 International License (http://creat​iveco​ Christensen MW et al (2017) Unveiling aerosol-cloud interactions
mmons​.org/licen​ses/by/4.0/), which permits unrestricted use, distribu- part 1: cloud contamination in satellite products enhances the
tion, and reproduction in any medium, provided you give appropriate aerosol indirect forcing estimate. Chem Phys Discuss Atmos.
credit to the original author(s) and the source, provide a link to the https​://doi.org/10.5194/acp-2017-450
Creative Commons license, and indicate if changes were made. Colarco PR, da Silva A, Chin M, Diehl T (2010) On-line simulations
of global aerosol distributions in the NASA-GEOS-4 model and
comparisons to satellite and ground-based aerosol optical depth.
J Geophys Res 115:D14207
References Dufresne J-L et al (2013) Climate change projections using the IPSL-
CM5 Earth System Model: from CMIP3 to CMIP5. Clim Dyn
Ackerman AS, Kirkpatrick MP, Stevens DE, Toon OB (2004) The 40(9–10):2123–2165
impact of humidity above stratiform clouds on indirect aerosol Ekman AML (2014) Do sophisticated parameterizations of aerosol-
climate forcing. Nature 432:1014–1017 cloud interactions in CMIP5 models improve the representation
Albrecht BA (1989) Aerosols, cloud microphysics and fractional cloud- of recent observed temperature trends? J Geophys Res 119:1–13
iness. Science 245:1227–1230 Elsaesser GS et al (2016) Multisensor climatology mean liquid water
Barnes WL, Pagano TS, Salomonson VV (1998) Prelaunch charac- path L3 monthly 1 degree x 1 degree V1, Greenbelt, MD, USA,
teristics of the MODerate resolution Imaging Spectroradiom- Goddard Earth Sciences Data and Information Services Center
eter (MODIS) on EOS-AM1. IEEE Trans Geosci Remote Sens (GES DISC). https​: //doi.org/10.5067/MEASU​R ES/MACLW​
36(4):1088–1100 PM. Accessed 2017–09
Bender FA-M, Charlson RJ, Ekman AM-L, Leahy L (2011) Quantifica- Elsaesser GS et al (2017) The multi-sensor advanced climatology of
tion of monthly mean regional scale albedo of marine stratiform liquid water path (MAC-LWP). J Clim 30:10193–10210
clouds in satellite observations and GCMs. J Appl Meteorol Clim Engström A, Bender FA-M, Charlson RJ, Wood R (2015) The
50:2139–2148 nonlinear relationship between albedo and cloud fraction
Bender FA-M, Engström A, Karlsson J (2016) Factors controlling on near-global, monthly mean scale in observations and in
cloud albedo in marine stratocumulus regions in climate models the CMIP5 model ensemble. Res Lett Geophys. https​: //doi.
and satellite observations. J Clim 29:10 org/10.1002/2015G​L0662​75
Bennartz R, Fan J, Rausch J, Leung LR, Heidinger AK (2011) Pol- Engström A, Ekman AML (2010) Impact of meteorological factors
lution from China increases cloud droplet number, suppresses on the correlation between aerosol optical depth and cloud frac-
rain over the East China Sea. Res Lett Geophys. https​://doi. tion. Gepophys Res Lett 37:18
org/10.1029/2011G​L0472​35 Feingold G, McComiskey A, Yamaguchi T, Johnson J, Carslaw K,
Schmidt KS (2016) New approaches to quantifying aerosol

13
Aerosol-cloud-radiation correlations 4391

influence on the cloud radiative effect. Proc Natl Acad Sci USA associated with the simulation of cloud optical properties. Clim
113:5812–5819 Dyn 5:175–187
Frey L, Bender FA-M, Svensson G (2017) Cloud albedo changes Lee LA, Reddington CL, Carslaw KS (2016) On the relationship
in response to anthropogenic sulfate and non-sulfate aerosol between aerosol model uncertainty and radiative forcing uncer-
forcings in CMIP5 models. Atmos Chem Phys 17:9145–9162 tainty. Proc Natl Acad Sci USA 113:5820–5827
Gassó S (2008) Satellite observations of the impact of weak vol- Loeb NG, Kato S, Loukachine K, Smith NM (2005) Angular distribu-
canic activity on marine clouds. J Geophys Res. https​: //doi. tion models for top-of-atmosphere radiative flux estimation from
org/10.1029/2007J​D0091​06 the Clouds and the Earth’s Radiant Energy System instrument on
Ghan S et al (2016) Challenges in constraining anthropogenic aerosol the Terra satellite. Part I: Methodology. J Atmos Ocean Technol
effects on cloud radiative forcing using present-day spatiotem- 22:338–351
poral variability. Proc Natl Acad Sci USA 113(21):5804–5811 Loeb NG, Wielicki BA, Rose FG, Doelling DR (2007) Variability in
George RC, Wood R (2010) Subseasonal variability of low cloud global top-of-atmosphere shortwave radiation between 2000 and
radiative properties over the southeast Pacific Ocean. Atmos 2005 Geophys. Res Lett 34:L03704
Chem Phys 10:4047–4063 Mace GG, Abernathy AC (2016) Observational evidence for aerosol
Grandey BS, Stier P (2010) A critical look at spatial scale choices invigoration in shallow cumulus downstream of Mount Kilauea.
in satellite-based aerosol indirect effect studies. Atmos Chem Geophys Res Lett 43(6):2981–2988
Phys 10:11459–11470 Malavelle F et al (2017) Strong constraints on aerosol-cloud interac-
Grosvenor DP, Wood R (2014) The effect of solar zenith angle tions from volcanic eruptions. Nature 546:485–491
on MODIS cloud optical and microphysical retrievals within McComiskey AM, Feingold G (2012) The scale problem in quantifying
marine liquid water clouds. Atmos Chem Phys 14:7291–7321. aerosol indirect effects. Atmos Chem Phys 12:1031–1049
https​://doi.org/10.5194/acp-14-7291-2014 McCoy DT, Hartmann DL (2015) Observations of a substantial
Grosvenor DP, Field PR, Hill AA, Shipway BJ (2017) The rela- cloud-aerosol indirect effect during the 2014–2015 Bárðarbunga-
tive importance of macrophysical and cloud albedo changes for Veiðivötn fissure eruption in Iceland. Res Lett Geophys. https​://
aerosol-induced radiative effects in closed-cell stratocumulus: doi.org/10.1002/2015G​L0670​70
insight from the modelling of a case study. Atmos Chem Phys McCoy DT, Bender FAM, Mohrmann JK, Hartmann DL, Wood R,
17:5155–5183 Grosvenor DP (2017) The global aerosol-cloud first indirect effect
Gryspeerdt E, Quaas J, Bellouin N (2016) Constraining the aerosol estimated using MODIS MERRA and AeroCom. J Res Geophys.
influence on cloud fraction. J Geophys Res 121(7):3566–3583 https​://doi.org/10.1002/2016J​D0261​41
Gryspeerdt E et al (2017) Constraining the instantaneous aero- McCoy DT, Bender FA-M, Grosvenor DP, Mohrmann JK, Hartmann
sol influence on cloud albedo. Proc Natl Acad Sci USA DL, Wood R, Field PR (2018) Predicting decadal trends in cloud
114:4899–4904 droplet number concentration using reanalysis and satellite data.
Hourdin F et al (2013) LMDZ5B: the atmospheric component of the Atmos Chem Phys 18:2035–2047
IPSL climate model with revisited parameterizations for clouds Michibata T, Takemura T (2015) Evaluation of autoconversion schemes
and convection. Clim Dyn 40(9–10):2193–3333 in a single model framework with satellite observations. J Geo-
Jiang H, Feingold G (2006) Effect of aerosol on warm convective phys Res Atmos 120:9570–9590
clouds: aerosol–cloud-surface flux feedbacks in a new coupled Michibata T, Suzuki K, Sato Y, Takemura T (2016) The sources of
large eddy model. J Geophys Res 111:D01202 discrepancies in aerosol-cloud-precipitation interactions between
Jiang JH, Su H, Zhai C, Massie ST, Schoeberl MR, Colarco PR, Plat- GCM and A-train retrievals. Atmos Chem Phys 16:15413–15424
nick S, Gu Y, Liou K-N (2011) Influence of convection and aero- Minnis PS et al (2011) CERES Edition-2 cloud property retrievals
sol pollution on ice cloud particle effective radius. Atmos Chem using TRMM VIRS and Terra and Aqua MODIS data, Part I:
Phys 11:457–463 Algorithms. IEEE Trans Geosci Remote Sens 49:4374–4400
Jiang JH et al (2012) Evaluation of cloud water vapor simulations in Minnis PS et al (2009) Cloud detection in nonpolar regions for CERES
CMIP5 climate models using NASA “A-Train” satellite observa- using TRMM VIRS and Terrra and Aqua MODIS data. IEEE
tions. J Geophys Res 117:D14 Trans Geosci Remote Sens 46(11):3857–3884
Johnson JS et al (2018) The importance of comprehensive parame- Molod A et al (2015) Development of the GEOS-5 atmospheric general
ter sampling and multiple observations for robust constraint of circulation model: evolution from MERRA to MERRA2. Geosci
aerosol radiative forcing. Chem Phys Discuss Atmos. https​://doi. Model Dev 8:1339–1356
org/10.5194/acp-2018-174 Neubauer D, Christensen MW, Poulsen C, Lohmann U (2017) Unveil-
Kawai H, Yukimoto S, Koshiro T, Oshima N, Tanaka T, Yoshimura H ing aerosol-cloud interactions part 2: minimixing the effects of
(2017) Improved representation of clouds in climate model MRI- aerosol swelling and wet scavengint in ECHAM6-HAM2 for com-
ESM2. CAS/JSC WGNE research activities in atmospheric and parison to satellite data. Chem Phys Discuss Atmos. https​://doi.
oceanic modelling/WMO, 47, 7.07-7.08. [tech report] org/10.5194/acp-2017-449
Kirkevåg A et al (2013) Aerosol–climate interactions in the Norwegian Penner JE, Xu L, Wang M (2011) Satellite methods underestimate
earth system model—NorESM1-M. Geosci Model Dev 6:207–244 indirect climate forcing by aerosols. Proc Natl Acad Sci USA
Klein S, Hartmann DL (1993) The seasonal cycle of low stratiform 108:13404–13408
clouds. J Clim 6:1587–1606 Peng Y, Lohmann U (2003) Sensitivity study of the spectral dispersion
Krotkov NA et al (2016) Aura OMI observations of regional SO 2 and of the cloud dorplet size distribution on the indirect aerosol effect.
NO 2 pollution changes from 2005 to 2015. Atmos Chem Phys Geophys Res Lett 20:10
16:4605–4629. https​://doi.org/10.5194/acp-16-4605-2016 Pincus R, Baker M (1994) Effect of precipitation on the albedo
Konsta D, Dufresne J-L, Chepfer H, Idelkadi A, Cesana G (2016) susceptibility of clouds in the marine boundary layer. Nature
Use of A-train satellite observations (CALIPSO-PARASOL) to 372:250–252
evaluate tropical cloud properties in the LMDZ5 GCM. Clim Dyn Possner A, Zubler EM, Lohmann U, Schär C (2016) The resolution
47:1263–1384 dependence of cloud effects and ship-induced aerosol-cloud
Le Treut H, Li Z-X (1991) Sensitivity of an atmospheric general interactions in marine stratocumulus. J Geophys Res Atmos
circulation model to prescribed SST changes: feedback effects 121:4810–4829

13
4392 F. A.-M. Bender et al.

Quaas J et al (2009) Aerosol indirect effects—general circulation model Twomey SA (1977) The influence of pollution on the shortwave albedo
intercomparison and evaluation with satellite data. Atmos Chem of clouds. J Atmos Sci 34:1149–1152
Phys 9:8697–8717 Wang M et al (2012) Constraining cloud lifetime effects of aerosols
Randles C et al (2016) The MERRA-2 aerosol assimilation. Technical using A-Train satellite observations. Geophys Res Lett 39:L15709.
report series on global modeling and data assimilation 45 https​://doi.org/10.1029/2012G​L0522​04
Remer LA et al (2005) The MODIS aerosol algorithm, products and Watanabe M et al (2010) Improved climate simulation by MIROC5:
validation. J Atmos Sci 62:947–973 mean states, variability, and climate sensitivity. J Clim
Remer LA et al (2008) Global aerosol climatology from the MODIS 23:6312–6335
satellite sensors. J Geophys Res 113:D14S07 Watanabe S et al (2011) MIROC-ESM 2010: model description and
Rienecker MM et al (2011) MERRA: NASA’s Modern-Era Ret- basic results of CMIP5-20c3m experiments. Geosci Model Dev
rospecitve analysis for Research and Applications. J Clim 4:845–872
24:3624–3648 Wielicki BA et al (1996) Clouds and the earth’s radiant energy system
Schutgens NAJ, Gryspeerdt E, Weigum N, Tsyro S, Goto D, Schulz (CERES): an earth observing system experiment. Bull Am Mete-
M, Stier P (2016) Will a perfect model agree with perfect obser- orol Soc 77:853–868
vations? The impact of spatial sampling. Atmos Chem Phys Wilcox EM (2010) Stratocumulus cloud thickening beneath layers of
16:6335–6353 absorbing smoke aerosol. Atmos Chem Phys 10:11769–11777
Rosenfeld D, Sherwood S, Wood R, Donner L (2014) Climate effects Wilcox EM et al (2016) Black carbon solar absorption suppresses tur-
of aerosol-cloud interactions. Science 343:379–380 bulence in the atmospheric boundary layer. Proc Natl Acad Sci
Schmidt A et al (2015) Satellite detection, long-range transport, and USA 113:11794–11799
air quality impacts of volcanic sulfur dioxide from the 2014–2015 Wilson DR, Ballard SP (1999) A microphysically based precipitation
flood lava eruption at Bárðarbunga (Iceland). J Geophys Res scheme for the UK Meteorological Office Unified Model. Q J R
Atmos 120:9739–9757 Meteorol Soc 125:1607–1636
Seethala C, Hovath A (2010) Global assessment of AMSR-E and Wood R (2007) Cancellation of aerosol indirect effects in marine stra-
MODIS cloud liquid water path retrievals in warm oceanic clouds. tocumulus through cloud thinning. J Atmos Sci 64:2657–2669
J Geophys Res 115:D13 Yuan T, Remer LA, Yu H (2012) Microphysical, macrophysical and
Small JD, Chuang PY, Feingold G, Jiang H (2009) Can aerosol radiative signatures of volcanic aerosols in trade wind cumulus
decrease cloud lifetime? Geophys Res Lett 36:L16806 observed by the A-train. Atmos Chem Phys 11:7119–7132
Stevens B, Feingold G (2009) Untangling aerosol effects on clouds and Yukimoto S et al (2012) A new global climate model of the meteoro-
precipitation in a buffered system. Nature 461:607–613 logical research institute: MRI-CGCM3—model description and
Takemura T, Nozawa T, Emori S, Nakajima T, Nakajima T (2005) Sim- basic performance. J Meterol Soc Jpn 90A:23–64
ulation of climate response to aerosol direct and indirect effects Zhang S et al (2016) On the characteristics of aerosol indirect effect
with aerosol transport-radiation model. J Geophys Res 110:2 based on dynamic regimes in global climate models. Atmos Chem
Taylor KE, Stouffer RJ, Meehl GA (1996) An overview of CMIP5 and Phys 16:2765–2783
the experiment design. Bull Am Meteorol Soc 93:485–498 Zhao B et al (2017) Decadal-scale trends in regional aerosol particle
Toll V, Christensen M, Gassó S, Bellouin N (2017) Volcano and ship properties and their linkage to emission changes. Environ Res
tracks indicate excessive aerosol-induced cloud water increases in Lett 12:5
a climate model. Geophys Res Lett 44:12,492–412,500
Twomey SA (1974) Pollution and the planetary albedo. Atmos Environ
8(1251):1256

13

You might also like