J. Meteor. Res.  2018, Vol. 32 Issue (5): 693-706   PDF    
http://dx.doi.org/10.1007/s13351-018-8011-z
The Chinese Meteorological Society
0

Article Information

SHI, Xiangjun, and Xiaohong LIU, 2018.
Sensitivity Study of Anthropogenic Aerosol Indirect Forcing through Cirrus Clouds with CAM5 Using Three Ice Nucleation Parameterizations. 2018.
J. Meteor. Res., 32(5): 693-706
http://dx.doi.org/10.1007/s13351-018-8011-z

Article History

Received February 1, 2018
in final form July 17, 2018
Sensitivity Study of Anthropogenic Aerosol Indirect Forcing through Cirrus Clouds with CAM5 Using Three Ice Nucleation Parameterizations
Xiangjun SHI1, Xiaohong LIU2,3     
1. Earth System Modeling Center and School of Atmospheric Sciences, Nanjing University of Information Science & Technology, Nanjing 210044, China;
2. Department of Atmospheric Science, University of Wyoming, WY 82071, USA;
3. Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029, China
ABSTRACT: Quantifying the radiative forcing due to aerosol–cloud interactions especially through cirrus clouds remains challenging because of our limited understanding of aerosol and cloud processes. In this study, we investigate the anthropogenic aerosol indirect forcing (AIF) through cirrus clouds using the Community Atmosphere Model version 5 (CAM5) with a state-of-the-art treatment of ice nucleation. We adopt a new approach to isolate anthropogenic AIF through cirrus clouds in which ice nucleation parameterization is driven by prescribed pre-industrial (PI) and present-day (PD) aerosols, respectively. Sensitivities of anthropogenic ice AIF (i.e., anthropogenic AIF through cirrus clouds) to different ice nucleation parameterizations, homogeneous freezing occurrence, and uncertainties in the cloud microphysics scheme are investigated. Results of sensitivity experiments show that the change (PD minus PI) in global annual mean longwave cloud forcing (i.e., longwave anthropogenic ice AIF) ranges from 0.14 to 0.35 W m–2, the change in global annual mean shortwave cloud forcing (i.e., shortwave anthropogenic ice AIF) from –0.47 to –0.20 W m–2, and the change in net cloud forcing from –0.12 to 0.05 W m–2. Our results suggest that different ice nucleation parameterizations are an important factor for the large uncertainty of anthropogenic ice AIF. Furthermore, improved understanding of the spatial and temporal occurrence characteristics of homogeneous freezing events and the mean states of cirrus cloud properties are also important for constraining anthropogenic ice AIF.
Key words: aerosol indirect forcing     cirrus clouds     CAM5    
1 Introduction

The estimation of anthropogenic aerosol indirect forcing (AIF) is important for the projection of future climate change (Boucher et al., 2013). Compared with warm clouds, much less attention has been paid to anthropogenic AIF through cirrus clouds (Myhre et al., 2013; Seinfeld et al., 2016). Although increasingly more general circulation models (GCMs) are including the treatment of aerosol–cirrus interactions (Liu et al., 2007; Gettelman et al., 2010; Barahona et al., 2014), it is often hard to clearly separate the anthropogenic ice AIF (i.e., anthropogenic AIF through cirrus clouds) from other AIFs (i.e., anthropogenic AIFs through warm clouds and mixed-phase clouds) without significant modifications of these GCMs. Thus, very few published papers present the anthropogenic ice AIF (Gettelman et al., 2012). Furthermore, compared with warm clouds, there are more uncertainties in representing cirrus clouds in GCMs because of the lack of observations. Thus, it is necessary to quantify the uncertainties in quantifying anthropogenic ice AIF. This study tries to fill this gap.

The uncertainties in estimating anthropogenic AIF by GCMs stem largely from two sources. First, GCMs themselves have substantial uncertainties in representing cloud processes including aerosol–cloud interactions. These uncertainties are an inevitable outcome of the limited understanding of realistic cloud processes and the scale disparity between GCMs and realistic cloud processes in the atmosphere (Seinfeld et al., 2016). Second, the change in the aerosol state itself from pre-industrial (PI) to present-day (PD) is uncertain. This uncertainty is caused by uncertainties in aerosol emissions and atmospheric aerosol processes (Boucher et al., 2013; Carslaw et al., 2013). In this study, we focus on the first kind of uncertainties.

Here, the Community Atmosphere Model version 5 (CAM5) is used to quantify the ice AIF. To separate the ice AIF from other AIFs, we adopt a new approach in which ice nucleation parameterization is driven by prescribed PD or PI aerosols. Sensitivities of ice AIF to several uncertain processes and parameters in CAM5 are investigated. The remainder of this paper is organized as follows. We introduce the model description and modifications in Section 2. Section 3 presents the sensitivity of ice AIF to different ice nucleation parameterizations and the uncertainties in the cloud microphysics scheme. Conclusions and discussion are given in Section 4.

2 Model and experiments 2.1 CAM5

The CAM5 model (Neale et al., 2012) is used in this study. Stratiform cloud microphysics in CAM5 is based on the MG scheme, which predicts the mass and number mixing ratios of cloud liquid and cloud ice, and diagnoses the mass and number mixing ratios of rain and snow (Morrison and Gettelman, 2008). The MG scheme is a framework of all cloud microphysics processes, such as activation of cloud condensation nuclei, accretion of cloud droplets and ice by rain, and sedimentation. In the MG scheme, each cloud microphysics process is described by empirical formula or parameterization. Note that, in the original MG scheme, ice nuclei concentrations are a function of temperature and not coupled to aerosol characteristics. To enable the CAM5 model to treat the aerosol indirect effects on cirrus clouds, the physically based ice nucleation parameterization of Liu and Penner (2005; hereafter LP) was applied in the MG scheme by Gettelman et al. (2010).

Here, the aerosol processes in CAM5 are treated with a three-mode (i.e., Aitken, accumulation, and coarse modes) aerosol module (Liu et al., 2012), which is coupled to the MG scheme for aerosol–cloud interactions. In the default CAM5, sulfate aerosols in the Aitken mode is used for homogeneous freezing and dust in the coarse aerosol mode is taken as potential ice nucleating particles (INPs). It is necessary to point out that the online calculated aerosols also drive the droplet activation parameterization, and the ice nuclei concentrations in mixed-phase clouds are a function of supersaturation and not coupled to aerosol characteristics.

There are two methods to diagnose the subgrid vertical velocity (Wsub), which is used to drive the ice nucleation parameterization. One is derived from the turbulent diffusion coefficient (Morrison and Gettelman, 2008), and the other is diagnosed from the turbulence kinetic energy (Gettelman et al., 2010).

The MG scheme distinguishes cloud ice and snow as two different classes of solid-phase condensates. The crystal/snow separating diameter (Dcs) has been used as a major tuning parameter in GCMs. Simulation results show that cirrus cloud properties are very sensitive to this tunable parameter (Zhang et al., 2013b). Unfortunately, it is difficult to directly constrain the Dcs value from observations (Eidhammer et al., 2017). Different Dcs values have been used in the MG scheme. In the version of CAM5 used in this work, the default value of Dcs is 400 μm, while Dcs has been set to 200 μm in an earlier version of CAM (Morrison and Gettelman, 2008).

Finally, it is necessary to point out that the shape parameter (μ) of the gamma size distribution is assumed to be 0 for ice crystals (i.e., exponential distribution) in the MG scheme, whereas observations show 0.5 < μ < 2.5 ( Heymsfield et al., 2002; Morrison and Grabowski, 2008). For ice particles with a certain mass and number, this assumption leads to overestimation of the number concentration for large ice particles, which then impacts the relevant cloud microphysical processes, such as the accretion of cloud ice by snow and the fall velocity (Khain et al, 2015).

2.2 Modifications to the standard CAM5

First, we improve the treatment of ice nucleation in a more realistic manner by including the effects of pre-existing ice crystals on ice nucleation in cirrus clouds (Shi et al., 2015). Second, a maximum distance for air parcel ascent (DHmax) is introduced to reduce the maximum ice supersaturation (Simax) that the rising cloud parcel could reach within one model time step (20–30 min), as model-diagnosed Wsub actually represents the vertical velocity fluctuation rather than the constant updraft velocity needed by ice nucleation parameterization (Shi and Liu, 2016). If Simax is less than the homogeneous freezing supersaturation threshold (Sihom), homogeneous freezing cannot occur—undoubtedly, considering DHmax can reduce occurrence frequency of homogeneous freezing (Frehom). Unfortunately, there are too few observed Frehomvalues to constrain the DHmax value (Cziczo et al., 2013). In this study, DHmax is a tunable parameter used for the restriction of Frehom. Finally, we modify the cirrus cloud formulas in the MG scheme to allow non-zero μ. The μ-relevant ice processes (i.e., number and mass weighted ice-crystal fall velocity, effective radius, ice deposition and sublimation rate, accretion of cloud ice by snow, and autoconversion of cloud ice to snow) are calculated based on the gamma size distribution (i.e., non-zero μ) instead of the default exponential distribution (i.e., zero μ).

Besides the above modifications, to investigate the sensitivity of ice AIF to different ice nucleation parameterizations, two other ice nucleation parameterizations developed by Kärcher et al. (2006, hereafter KL) and Barahona and Nenes (2009, hereafter BN) are implemented in the MG scheme. Thus, there are three options available (LP, BN, and KL) for calculating new-formed ice crystal number concentration in the MG scheme. All three ice nucleation parameterizations are developed based on the ice formation process in an adiabatically arising air parcel, in which the competition between homogeneous freezing on sulfate aerosol particles and heterogeneous freezing on INPs is considered. Although the microphysical processes are similar in the three corresponding air parcel models, model results are not quite similar in some extents because of different parcel model assumptions (e.g., nuclei spectra to represent the heterogeneous freezing) and different numerical solution methods (e.g., the homogeneous freezing competition among different sizes of sulfate aerosol particles). Furthermore, the approaches for deriving these ice nucleation parameterizations from the corresponding air parcel models are also different. For example, the LP parameterization is derived from fitting parcel model results. The BN parameterization is derived from the analytical solution of parcel model equations. The KL parameterization explicitly calculates the evolution of ice supersaturation in the rising air parcel. Therefore, it is possible for significant differences in newly formed ice number concentration (Ni) calculated from the three parameterizations.

Here we compare the three ice nucleation parameterizations (i.e., LP, BN, and KL) used in this study. In Fig. 1, Ni is shown as a function of sulfate aerosol number concentration (Nso4, Fig. 1a) and dust number concentration (Ndust, Fig. 1b), calculated from the three parameterizations. Ni calculated from the LP and BN parameterizations increases with increasing Nso4 (a tenfold increase in Nso4 roughly doubles Ni), whereas Ni calculated from the KL parameterization is not sensitive to Nso4 except at very low Nso4. The reason is that the numerical solution method of homogeneous freezing used in the KL parameterization neglects the competition among different sizes of sulfate aerosol particles, which results in very small sensitivity of Ni to Nso4 (Liu and Shi, 2018). Our parcel model using this KL approach can reproduce a nearly zero sensitivity of Ni to Nso4 (figure omitted), which is similar to the KL results in Fig. 1. Under these low Nso4 conditions all sulfate aerosols will be activated to ice crystals. This is the reason why Ni calculated from the KL parameterization increases with increasing Nso4. Although Ni calculated from the LP and BN parameterization show the same trend, Ni from LP is larger than BN under most conditions (Fig. 1a). The main reason is that the homogeneous nucleation rate is calculated using different methods in the corresponding air parcel models. It is clear that the homogeneous freezing forms most ice crystals as Ndust used for heterogeneous freezing is only 10 L–1. Compared with Ni with an updraft velocity of 0.1 m s–1, Ni with an updraft velocity of 0.5 m s–1 increases by a factor of ~10. With the aid of INPs, heterogeneous freezing tends to occur at lower Si than that required by homogeneous freezing (Sihom), preventing the homogeneous freezing from occurring or reducing the homogeneous freezing ice crystal number (Spichtinger and Gierens, 2009). Taking the BN parameterization driven by an updraft velocity of 0.1 m s–1 as an example, Ni decreases with increasing Ndust under 1 L–1 < Ndust < 10 L –1 (Fig. 1b). The reason is that ice crystals produced by heterogeneous freezing deplete the water vapor in the parcel, and then reduce the homogeneous freezing ice crystal number. At Ndust > 10 L –1, homogeneous freezing is completely shut off, and Ni from heterogeneous freezing increases with increasing Ndust. Here, the competition between homogeneous and heterogeneous freezing is considered in all three parameterizations.

Figure 1 Ice number concentration (a) as a function of sulfate number concentration and (b) as a function of dust number concentration. Results are derived from offline calculations based on the three ice nucleation parameterizations with (a) dust number concentration of 10 L–1 and (b) sulfate number concentration of 100 cm–3. The blue lines indicate LP parameterization, red lines indicate BN parameterization, and green lines indicate KL parameterization. Solid lines are for an updraft velocity of 0.1 m s–1 and dashed lines for an updraft velocity of 0.5 m s–1. Initial temperature and air pressure are 220 K and 250 hPa, respectively.
2.3 Prescribed aerosol data

Zhang et al. (2013a) introduced a new approach to CAM5, which allows the separation of the impact of aerosols on cold ice clouds from that on liquid clouds by using the prescribed aerosol capability. The prescribed PD and PI aerosol data are acquired from the baseline CAM5 simulations. In the default CAM5 model, aerosol characteristics, which are used for driving the ice nucleation parameterization and droplet activation parameterization, are online calculated from the aerosol module. To isolate the anthropogenic AIF on cirrus clouds from that on liquid-phase clouds, the ice nucleation parameterization is driven by the prescribed PD and PI aerosol data, respectively, whereas the droplet activation parameterization is still driven by the online PD calculated aerosol. In this study, each experiment has a pair of simulations (hereafter, PD and PI) with ice nucleation parameterization driven by the prescribed PD and PI aerosol data, respectively. Therefore, the difference between PD and PI simulations within one experiment only indicates the anthropogenic aerosol (i.e., the difference between PD and PI prescribed aerosol) indirect effects through cirrus clouds. Note that the anthropogenic ice AIF, which is calculated as the top-of-atmosphere radiative flux perturbation, not only comes from the changes in cirrus clouds but also from other radiation-relevant changes caused by rapid adjustments to anthropogenic aerosols (e.g., the interactions among ice, mixed-phase, and liquid clouds). Undoubtedly, this method also has shortcomings. Because the aerosol data used for driving ice nucleation is prescribed, the aerosol characteristics cannot respond to the influence of clouds and precipitation, such as wet removal of aerosols by precipitation. Also this method may inhibit the interactions among ice, mixed-phase, and liquid clouds through conversions of water vapor and cloud hydrometeors.

Evaluation of CAM5 aerosol simulations can be found in Liu et al. (2012). Here, the prescribed aerosol data used as inputs to the ice nucleation parameterization are shown in Fig. 2. Annual and zonal mean PD (for year 2000) sulfate aerosol number concentrations in the Aitken mode are higher than 100 cm–3 in the tropical upper troposphere due to the new particle formation favored at colder temperatures. Unlike the sulfate aerosol number concentrations used for homogenous freezing, coarse mode dust number concentrations used for heterogeneous freezing are lower than 100 L–1 in the upper troposphere, and even lower than 10 L–1 over the Southern Hemisphere (SH). Compared with the PI (for the year 1850) aerosol concentrations, PD sulfate aerosol number concentrations are significantly increased, with normalized changes larger than 100% over the Northern Hemisphere (NH), and larger than 200% over the NH mid- and high latitudes. In contrast to sulfate aerosol, changes in dust number concentrations are not obvious. In this work, one sensitivity experiment uses higher INP number concentrations, following the study of Penner et al. (2015) in which coarse mode dust, 10% of accumulation mode dust, and 0.1% of accumulation mode soot are allowed to act as INPs. The total dust number concentrations (coarse mode + 10% accumulation mode) used for this experiment increase noticeably compared with the coarse mode dust. The 0.1% of accumulation mode soot concentrations, which can act as INPs, are very low (< 10 L–1) in the upper troposphere. However, their changes from PI to PD are substantial.

Figure 2 Annual and zonal mean number concentrations (L–1) of present-day (PD) aerosol particles: (a) Aitken mode sulfate, (b) coarse mode dust, (c) coarse mode dust + 10% of accumulation mode dust, and (d) 0.1% of accumulation mode soot. The relative changes in number concentrations compared with the pre-industrial (PI) aerosols, namely (PD – PI) / PI (%) are shown in (e)–(h) correspondingly.
2.4 Description of model experiments

In this study, the most uncertain aspects in the representation of cirrus clouds (including ice nucleation parameterization) that impact the estimation of anthropogenicice AIF were investigated. Table 1 summarizes all the experiments performed in this study. Compared with the default CAM5, the CTL experiment takes the pre-existing ice effects into account, and restricts the occurrence of homogeneous freezing with DHmax. Note that the effects of pre-existing ice and DHmax have been implemented in the development trunk of the CAM model (not yet released). As well as the CTL experiment, we carry out eight sensitivity experiments in three groups. Because ice nucleation is the root process for ice AIF, we first investigate the uncertainties of ice nucleation parameterization in group A. The CTL, ParaBN, and ParaKL experiments show the impact of different sensitivities of homogeneous freezing-produced Ni to Nso4. Comparison between the CTL and INP experiments shows the impact of uncertainties regarding INPs (i.e., heterogeneous freezing). Besides uncertainties in ice nucleation parameterization, how to drive the ice nucleation parameterization in GCMs can also significantly impact the ice nucleation process especially for homogeneous freezing. Unfortunately, cloud-scale vertical motions are poorly resolved in GCMs (e.g., the tunable parameter DHmax). The experiments in group B investigate the uncertainties in how to use ice nucleation parameterization (i.e., sensitivity to homogeneous nucleation occurrence). Finally, there are also some uncertainties about other cloud processes in the cloud microphysics scheme due to the lack of cirrus cloud observations or the coarse-resolution of GCMs. CAM5 simulation results show that cirrus cloud properties are very sensitive to some tunable parameter in the cloud microphysics schemes (Zhang et al., 2013b). The base state of cirrus clouds might significantly impact the estimation of anthropogenic ice AIF. Group C shows the sensitivity of ice AIF to some important uncertain variables (i.e, Wsub, Dcs, and μ) in the MG scheme.

Table 1 Description of sensitivity experiments used in this study
Name Description
CTL LP ice nucleation parameterization. Only coarse mode dust is considered as INPs. Subgrid updraft velocity (Wsub) is
 derived from turbulent kinetic energy. The maximum distance for air parcel ascent (DHmax) is 450 m. Cloud
 ice/snow separating diameter (Dcs) is 400 μm. Spectra shape parameter of ice size distribution (μ) is zero.
Group A: Sensitivity to different ice nucleation parameterizations and INPs
 ParBN  Same as CTL, but with BN ice nucleation parameterization.
 ParKL  Same as CTL, but with KL ice nucleation parameterization.
 INP  Same as CTL, but with coarse mode dust, 10% of accumulation mode dust, and 0.1% of accumulation mode soot
  considered as INPs.
Group B: Sensitivity to homogeneous nucleation occurrence
 NoDH  Same as CTL, but with DHmax removed.
 DH400  Same as CTL, but with DHmax set to be 400 m.
Group C: Sensitivity to cloud microphysics scheme uncertainties
 Wsub  Same as CTL, but with subgrid updraft velocity derived from turbulent diffusion coefficient.
 Dcs200  Same as CTL, but with Dcs set to be 200 μm.
 Miu2  Same as CTL, but with μ set to be 2.

All simulations in this study were run for 11 yr at a 1.9° × 2.5° horizontal resolution with 30 vertical levels and a 30-min time step, using prescribed sea surface temperatures and PD aerosol emissions. Each experiment had a pair of simulations with ice nucleation parameterization driven by PD and PI prescribed aerosols, whereas online calculated aerosols are used to drive the droplet activation parameterization for liquid clouds. The last 10 years’ results were used in our analysis.

3 Ice AIF and its uncertainty 3.1 Sensitivity to different ice nucleation parameterizations

Table 2 gives the global annual mean values of cloud radiative forcings, ice water path (IWP), high-cloud fraction (CLDH), column cloud ice number concentration (CDNI), and 60°S–60°N averages at 197 hPa of ice effective radius (Ri197) and cirrus cloud occurrence frequency (FRE197) from the PD simulations. Table 2 gives model results from all experiments. Here in this subsection, we only discuss about the sensitivity to the CTL, ParBN, and ParKL experiments. Results from other experiments will be analyzed in the following subsections. There are no remarkable differences in those variables listed in Table 2 among the CTL, ParBN, and ParKL experiments.

Table 2 Global annual mean results from all experiments. Shown are net cloud forcing (CF; W m–2) as well as the longwave (LWCF; W m–2) and shortwave (SWCF; W m–2) components compared with ERBE data (Kiehl and Trenberth, 1997) and CERES data (Loeb et al., 2009), ice water path (IWP; g m–2) compared with CloudSat data (Li et al., 2012), high-cloud fraction (CLDH; %) compared with ISCCP data (Rossow and Schiffer, 1999), and column cloud ice number concentration (CDNI; 106 m–2). Also shown are 60°S–60°N averages at 197 hPa of ice effective radius (Ri197; μm) and cirrus cloud occurrence frequency (FRE197; %)
Name CF LWCF SWCF IWP CLDH CDNI Ri197 FRE197
OBS 27–31 –46 to –53 25.8 21–33
CTL –27.19 26.65 –53.84 19.67 37.92 185.68 37.08 16.98
ParBN –27.08 27.07 –54.12 19.45 37.63 191.01 35.65 17.14
ParKL –27.41 26.27 –53.69 19.02 38.46 184.09 36.80 16.51
INP –27.31 26.58 –53.89 19.68 37.87 182.97 37.33 16.92
NoDH –26.92 29.24 –56.15 21.07 39.70 245.62 35.24 18.69
DH400 –27.56 24.49 –52.04 18.18 36.59 129.50 38.91 15.56
Wsub –28.09 36.60 –64.69 25.29 45.29 799.26 30.18 22.08
Dcs200 –30.77 24.65 –55.42 10.35 35.54 219.58 26.11 14.76
Miu2 –26.49 35.78 –62.27 38.06 43.71 286.28 36.50 22.17

Figure 3 compares the variation of modeled Ni as a function of temperature against a compilation of field campaign data (Krämer et al., 2009). It should be noted that observed Ni might be overestimated due to the shattering of large ice particles, especially for relatively warm cirrus clouds (> 230 K) (Field et al., 2006). Here, we discuss modeled Ni from the CTL, ParBN, and ParKL experiments from the view of three temperature ranges, > 230 K, 205–230 K, and < 205 K. At temperatures above 230 K, we should keep in mind that the observed Ni might be overestimated and thus it is difficult to evaluate modeled Ni at these temperatures. In the temperature range of 205–230 K, modeled Ni from the CTL, ParBN, and ParKL experiments are close to the observations. There are no large differences among these three experiments. At temperatures below 205 K, modeled Ni from all experiments is obviously higher than observations. It is likely that some important mechanisms are missing in current GCMs. For example, aerosols rich with organic matter may become glassy below the glassy-transition temperature (approximately 205 K), thereby preventing homogeneous freezing from occurring (Jensen et al., 2010; Murray et al., 2010).

Figure 3 In-cloud ice crystal number concentration (Ni; L–1) versus temperature for experiments in (a) group A, (b) group B, and (c) group C. Model results are sampled every three hours over 30°S–75°N regions including the observation locations reported in Krämer et al. (2009). The solid lines indicate the 50th percentile. The vertical bars overloading on the solid lines indicate the ranges between 25th and 75th percentiles. The gray color indicates observations between 25th and 75th percentiles.

Figure 4 compares the probability distribution frequency (PDF) of modeled Ni against aircraft measurements from the Small Particles in Cirrus (SPARTICUS) campaign, which was carried out over the Southern Great Plains (SGP) site (36.6°N, 97.5°W). We sample instantaneous Ni over the field campaign site every three hours. Although the observation date has taken the shattering of ice crystals into account by using new probes and improved algorithms (Lawson, 2011), it is unclear to what extent very high Ni (> 1000 L–1) is influenced by the ice shattering. Furthermore, it is important to point out that the observed Ni is from in situ aircraft measurements, while the modeled Ni represents the average of cloud fraction within a model grid cell (about 100 km). Therefore, the occurrence frequency of very high Ni from modelresults should be lower than observations. This point should be kept in mind when comparing modeled Ni with in-situ observations. Except for very high Ni, PDFs of modeled Ni from the CTL, ParBN, and ParKL experiments are close to each other, and agree well with observations.

Figure 4 Probability distribution frequency (PDF) of in-cloud ice number concentration (Ni; L–1) for experiments in (a) groups A, (b) group B, and (c) group C. Black lines refer to observations. Model results are sampled over the field campaign site (36.6°N, 97.5°W) every three hours.

Figure 5 shows the changes (PD minus PI) in the annual zonal mean in-cloud ice crystal number concentration (ΔNi) from the CTL, ParBN, and ParKL experiments. Note that, in this work, we use “Δ” to indicate a change between PD and PI. As LP and BN parameterizations show moderate sensitivities of Ni to Nso4 (Fig. 1) and Nso4 is significantly increased from PI to PD over the NH (Fig. 2), ΔNi from the CTL and ParBN experiments are statistically significant in the NH upper troposphere. In contrast, ΔNi from the ParKL experiment is generally not statistically significant except in limited regions of the NH upper troposphere.

Figure 5 Changes between present-day (PD) and pre-industrial (PI) times in annual zonal mean in-cloud ice crystal number concentration [Ni (PD–PI); L–1] from (a) the CTL, (b) ParBN, and (c) ParKL experiments. Black dots represent values reaching the 5% significance level of Student’s t-test. Only shown are model grids where the annual zonal mean ice nucleation occurrence frequency is greater than 0.001.

Table 3 gives the changes (PD minus PI) in global annual mean values for the variables listed in Table 2. The change in column cloud ice number concentration (ΔCDNI) from the ParKL experiment is 7.58 × 106 m–2, lower than that from the CTL (11.00 × 106 m–2) and ParBN (10.84 × 106 m–2) experiments. Note that homogeneous freezing usually occurs under high Wsub conditions because both pre-existing ice effects and DHmax are considered. If Nso4 is relatively low, most sulfate aerosols will be nucleated to become ice crystals under these high Wsub conditions. This is the reason why ΔCDNI from the ParKL experiment is not much lower than those from CTL and ParBN experiments. The change in ice water path (ΔIWP) from the ParKL experiment is 0.09 g m–2, which is also lower than CTL (0.21 g m–2) and ParBN (0.25 g m–2) experiments. Note that the ice cloud fraction is diagnosed based on total water (vapor + ice) in the CAM5 model. Therefore, an increasing IWP might lead to positive changes in the high-cloud fraction (ΔCLDH). ΔCLDH from the CTL, ParBN, and ParKL experiments are 0.15, 0.26, and 0.20%, respectively. All experiments show that the averages of Ri197 decrease from PI to PD. ΔRi197 from the CTL, ParBN, and ParKL experiments are –0.20, –0.32 and –0.19 μm, respectively. To analyze the changes in cloud top height, we calculated cirrus cloud occurrence frequency (FRE) at 121, 168, and 197 hPa (only FRE197 shown in Tables 2 and 3). All experiments showed that the averaged FRE at 121, 168, and 197 hPa are increased from PI to PD. These results indicate that cloud top height increase from PI to PD. ΔFRE197 from the CTL, ParBN, and ParKL experiments are 0.15, 0.24, and 0.08%, respectively. All three experiments show significant changes in longwave cloud forcing (ΔLWCF) and shortwave cloud forcing (ΔSWCF). The change in longwave cloud forcing (ΔLWCF) from the ParKL experiment (0.21 W m–2) is weaker than those from the CTL (0.27 W m–2) and ParBN (0.33 W m–2) experiments. One possible reason is that ΔIWP from the ParKL experiment (0.09 g m–2) is significantly lower than the CTL (0.21 g m–2) and ParBN (0.25 g m–2) experiments. ΔSWCF from the CTL, ParBN, and ParKL experiments are –0.24, –0.31, and –0.26 W m–2, respectively.

Table 3 Same as Table 2, but for the differences “Δ” between present-day and pre-industrial. Corresponding standard deviations are also given in brackets. The last two lines show ensemble statistics. Ensem9 is calculated from all the nine experiments. Ensem6 only includes CTL and group A and B experiments
Name CF LWCF SWCF IWP CLDH CDNI Ri197 FRE197
ΔCTL 0.04 (0.15) 0.27 (0.07) –0.24 (0.13) 0.21 (0.11) 0.15 (0.21) 11.00 (0.89) –0.20 (0.08) 0.15 (0.11)
ΔParBN 0.02 (0.21) 0.33 (0.06) –0.31 (0.18) 0.25 (0.11) 0.26 (0.14) 10.84 (1.28) –0.32 (0.20) 0.24 (0.11)
ΔParKL –0.05 (0.21) 0.21 (0.10) –0.26 (0.25) 0.09 (0.15) 0.20 (0.17) 7.58 (1.20) –0.19 (0.07) 0.08 (0.10)
ΔINP –0.10 (0.13) 0.14 (0.10) –0.24 (0.12) 0.16 (0.14) 0.01 (0.14) 9.10 (1.31) –0.14 (0.13) 0.08 (0.13)
ΔNoDH 0.05 (0.23) 0.32 (0.10) –0.27 (0.20) 0.21 (0.09) 0.19 (0.24) 14.34 (1.93) –0.29 (0.19) 0.21 (0.13)
ΔDH400 –0.11 (0.15) 0.17 (0.05) –0.29 (0.16) 0.08 (0.09) 0.18 (0.13) 5.86 (0.66) –0.12 (0.11) 0.13 (0.11)
ΔWsub –0.01 (0.30) 0.19 (0.17) –0.20 (0.26) 0.36 (0.17) 0.08 (0.25) 34.61 (5.27) –0.02 (0.19) 0.03 (0.18)
ΔDcs200 –0.09 (0.16) 0.23 (0.08) –0.32 (0.21) 0.08 (0.09) 0.14 (0.18) 15.23 (1.03) –0.11 (0.10) 0.20 (0.11)
ΔMiu2 –0.12 (0.15) 0.35 (0.19) –0.47 (0.14) 0.42 (0.28) 0.12 (0.25) 22.59 (1.89) –0.49 (0.14) 0.15 (0.27)
Ensem9 –0.04 (0.06) 0.25 (0.07) –0.29 (0.07) 0.21 (0.11) 0.15 (0.07) 14.57 (8.48) –0.21 (0.13) 0.14 (0.07)
Ensem6 –0.03 (0.07) 0.24 (0.07) –0.27 (0.04) 0.17 (0.06) 0.17 (0.08) 9.79 (2.71) –0.21 (0.07) 0.15 (0.06)
3.2 Sensitivity to uncertainty of INPs

Here we discuss the impacts of uncertainty regarding INPs. Compared with the CTL experiment, 10% of accumulation mode dust and 0.1% of accumulation mode soot are also considered to be INPs in the INP experiment. The impacts of a higher number of INPs on ice nucleation are shown in Fig. 6. The occurrence frequency of homogeneous freezing (Frehom) from the INP experiment is reduced compared with the CTL experiment, due to the higher competition between homogeneous and heterogeneous freezing. In this work, Frehom indicates the occurrence frequency of homogeneous freezing under the condition of ice nucleation events. Both CTL and INP experiments show that Frehom is less than 10% in most regions except in the tropical upper troposphere and over the South Pole. This is consistent with the observation analysis that heterogeneous freezing is the dominant formation mechanism of cirrus clouds (Cziczo et al., 2013). The contribution of ice crystals from heterogeneous freezing (Nihet) to total Ni is significantly increased over the NH in the INP experiment compared with the CTL experiment. Higher Nihet from the INP experiment (not shown) could inhibit the production of ice crystals from homogeneous freezing (Nihom). Thus, there are no significant differences in Ni between the INP and CTL experiments, but the relative contribution from homogeneous versus heterogeneous freezing changes. Furthermore, both the variations of Ni as a function of temperature and the PDFs of Ni over the SGP site are similar in the INP and CTL experiments (Figs. 3 and 4). The globalmean CDNI from the INP experiment (182.97 × 106 m–2) is close to that from the CTL experiment (185.68 × 106 m–2, Table 2). Unlike CDNI, ΔCDNI (Table 3) (i.e., difference in CDNI between PD and PI) from the INP experiment (9.10 × 106 m–2) is obviously less than that from the CTL experiment (11.00 × 106 m–2), owing to the more important role of heterogeneous freezing at the PI in the INP experiment. Correspondingly, ΔLWCF (0.14 W m–2) from the INP experiment is much lower compared with the CTL experiment (0.27 W m–2). However, ΔSWCF from the two experiments (–0.24 W m–2) are same. Therefore, ΔCF from the INP experiment (–0.10 W m–2) is much smaller than that from the CTL experiment (0.04 W m–2). Note that the standard deviations of ΔCF from the INP and CTL experiments are 0.15 and 0.13 W m–2, respectively. ΔCF in these two experiments are within their standard deviations.

Figure 6 (a) Annual zonal mean in-cloud ice crystal number concentration (Ni; L–1), (b) occurrence frequency of homogeneous freezing (Frehom; %), and (c) percentage contribution from heterogeneous freezing-produced Ni (Nihet) to total Ni when ice freezing occurs (Nihet/Ni; %) from (a–c) the CTL experiment, and those from (d–f) the INP experiment. Only shown are model grids where the annual zonal mean ice nucleation occurrence frequency is greater than 0.001.
3.3 Sensitivity to homogeneous nucleation occurrence (DHmax)

Now we look at changes in the simulations resulting from different DHmax values. Figure 7 shows Frehom and Ni at 197 hPa from CTL, NoDH and DH400 experiments. The NoDH experiment shows that Frehom is larger than 10% except near dust source regions (e.g., Saharan Desert, Arabian Desert, and Gobi Desert) where the high dust INPs concentrations could prevent the homogeneous freezing from occurring in most conditions. After considering DHmax, Frehom from the CTL (DHmax = 450 m) and DH400 (DHmax = 400 m) experiments are less than 10% in most regions. The global mean Frehom is reduced from 15.16% in the NoDH to 4.24% and 3.50% in the CTL and DH400 experiments, respectively. As a result, the global mean Ni is reduced from 493.4 L–1 in the NoDH to 356.9 and 318.1 L–1 in the CTL and DH400 experiments, respectively. Ni usually becomes very high (> 1000 L–1) once homogeneous freezing occurs, and the high Ni could persist for dozens of model time steps (Shi et al., 2015). Thus, the global mean Ni from the DH400 experiment is still high (> 300 L–1, much larger than the INPs concentration) even if Frehom is reduced to 3.50%. Figure 3 shows that Ni from the NoDH experiment is higher than the CTL and DH400 experiments. Accordingly, Figure 4 shows that the PDF of Ni > 100 L –1 from the NoDH experiment is higher than that from the CTL and DH400 experiments.

Figure 7 Annual mean (a–c) occurrence frequency of homogeneous freezing (Frehom; %) and (d–f) number concentration of in-cloud ice crystals (Ni; L–1) at 197 hPa from (a, d) CTL, (b, e) NoDH, and (c, f) DH400 experiments. Only shown are model grids where the annual mean ice nucleation occurrence frequency is greater than 0.001. Global mean values are shown in the upper right corner.

The global mean CDNI from the NoDH experiment is 245.62 × 106 m–2, which is significantly higher than that from the CTL experiment (185.68 × 106 m–2, Table 2). As a result, the global mean LWCF (29.24 W m–2) and SWCF (–56.15 W m–2) from the NoDH experiment are stronger compared with the CTL experiment (LWCF of 26.65 W m–2; SWCF of –53.84 W m–2). The global mean ΔLWCF (0.32 W m–2) and ΔSWCF (–0.27 W m–2) from the NoDH experiment are also stronger than those from the CTL experiment (ΔLWCF of 0.27 W m–2; ΔSWCF of –0.24 W m–2, Table 3). The global mean CDNI (129.50 × 106 m–2) from the DH400 experiment, by contrast, is much lower than that from the CTL experiment (Table 2). As a result, the global mean LWCF (24.49 W m–2) and SWCF (–52.04 W m–2) from the DH400 experiment are weaker. ΔLWCF (0.17 W m–2) from the DH400 experiment is also smaller compared with the CTL experiment (Table 3).

3.4 Sensitivity to cloud scheme uncertainties (Wsub, Dcs, and μ)

Wsub from the Wsub experiment is much larger than that from the CTL experiment (not shown). From Fig. 8, we can see that CDNI from the Wsub experiment is several times higher than the CTL experiment, as Ni is very sensitive to updraft velocity (Fig. 1). Compared with the CTL experiment, IWP from the Wsub experiment is increased, and both LWCF and SWCF become stronger (Fig. 8). Figure 3 also shows that Ni from the Wsub experiment is significantly higher than the CTL experiments. In Fig. 4, it is demonstrated that the PDF of Ni > 200 L –1 from the Wsub experiment is higher than that from others experiments. The global mean CDNI from the Wsub experiment (799.26 × 106 m–2) is three times larger than that from the CTL experiment (185.68 × 106 m–2, Table 2). The global mean LWCF and SWCF from the Wsub experiment are enhanced by 9.95 and –10.84 W m–2, respectively. Although the global mean ΔCDNI from the Wsub experiment (34.61 × 106 m–2) is two times larger than that from the CTL experiment (11.0 × 106 m–2), the relative change (ΔCDNI/CDNI) from the Wsub experiment (4.3%) is less than that from the CTL experiment (5.9%). This might be the main reason why global mean ΔLWCF (0.19 W m–2) and ΔSWCF (–0.20 W m–2) from the Wsub experiment are weaker than those from the CTL experiment (Table 3). This also indicates that the sensitivity of SWCF and LWCF to Ni become weakerunder high Ni conditions. Figure 9 shows the changes in annual and zonal mean ΔLWCF, ΔSWCF, ΔCDNI, and ΔIWP. ΔCDNI from the Wsub experiment is larger than 50 × 106 m–2 between 40° and 70°N, which is several times larger than that from the CTL experiment. However, ΔLWCF from the Wsub experiment is close to that from the CTL experiment between 40° and 70°N. Compared with ΔLWCF, the fluctuation of ΔSWCF is more complicated because changes in warm clouds can also impact SWCF. ΔSWCF from the CTL experiment between 50° and 70°N (positive) has the opposite sign to those from the other experiments (negative). However, these ΔSWCF are within the ranges of 2 standard deviations.

Figure 9 Changes in the annual zonal mean (a) longwave cloud forcing (LWCF), (b) shortwave cloud forcing (SWCF), (c) column cloud ice number concentration (CDNI), and (d) ice water path (IWP), due to anthropogenic aerosol indirect effect. The vertical bars indicate the ranges of two standard deviations calculated from the difference of each year.

The threshold diameter Dcs that separates cloud ice and snow can significantly impact the autoconversion from cloud ice to snow in the MG scheme, and consequently exert a strong impact on the simulated IWP and LWCF (Gettelman et al., 2010; Zhang et al., 2013b). Compared with the CTL experiment, IWP from the Dcs200 experiment is significantly reduced due to a higher autoconversion rate from cloud ice to snow (Fig. 8). The global mean IWP from the Dcs200 experiment (10.35 g m–2) is half of that from the CTL experiment (19.67 g m–2), and much lower than the observation (25.8 g m–2, Table 2). The global mean LWCF from the Dcs200 experiment is decreased by 2.0 W m–2 compared with the CTL experiment (26.65 W m–2). The global mean CDNI from the Dcs200 experiment (219.58 × 106 m–2) is increased compared with the CTL experiment (185.68 × 106 m–2). One possible reason is that the number-weighted ice-crystal fall velocity from the Dcs200 experiment is much smaller than that from the CTL experiment (not shown). The autoconversion rate from cloud ice to snow (i.e., Dcs) has significant impacts not only on cirrus clouds but also on warm clouds through the interactions among ice, mixed-phase, and liquid clouds. The Dcs200 experiment produces a higher liquid water path and higher column cloud droplets number concentration (not shown). This might be the main reasonwhy the global mean SWCF from the Dcs200 experiment (–55.42 W m–2) is stronger than that from the CTL experiment (–53.83 W m–2). The annual zonal mean distributions of ΔCDNI in the Dcs200 and CTL experiments are similar (Fig. 9). Compared with the CTL experiment, the Dcs200 experiment gives a weaker global mean ΔLWCF (0.23 W m–2) and a stronger ΔSWCF (–0.32 W m–2, Table 3). Thus, ΔCF from the Dcs200 experiment (–0.09 W m–2) is less than that from the CTL experiment (0.04 W m–2). Therefore, a smaller Dcs leads to a weaker ΔLWCF and a stronger ΔSWCF, and a lower ΔCF.

In the absence of any changes in ice number concentration or ice mass mixing ratio, an increase of μ results in a lower number concentration of larger ice crystals (not shown), and thus reduces the autoconversion rate from cloud ice to snow. This is a plausible explanation for the larger IWP in the Miu2 experiment (Fig. 8). Actually, the impacts of μ are complex because μ affects some other cloud microphysics (e.g., number and mass weighted ice-crystal fall velocities, and ice deposition and sublimation rate). The global mean IWP from the Miu2 experiment is increased to 38.06 g m–2, much larger than the observation (25.8 g m–2, Table 2). The Miu2 experiment produces much stronger LWCF (35.78 W m–2) and SWCF (–62.27 W m–2) compared with the CTL experiment. Because IWP from the Miu2 experiment between 40° and 70°S is high (> 40 g m–2, Fig. 8), fluctuations of ΔIWP from the Miu2 experiment between 40° and 70°S are also large (Fig. 9). For the Miu2 experiment, the global mean ΔLWCF (0.35 W m–2) and ΔSWCF (–0.47 W m–2) are stronger than those from the CTL experiment (ΔLWCF of 0.27 W m–2; ΔSWCF of –0.24 W m–2); the global mean ΔCF (–0.12 W m–2) is much smaller than that from the CTL experiment (0.04 W m–2, Table 3). To sum up, μcan significantly impact cirrus cloud properties and anthropogenic ice AIF.

Figure 8 Observed and simulated annual zonal mean (a) longwave cloud forcing (LWCF), (b) shortwave cloud forcing (SWCF), (c) column cloud ice number concentration (CDNI), and (d) ice water path (IWP). Black solid line refers to observation from CERES data (Wielicki et al., 1996).
4 Conclusions and discussion

This study quantifies the anthropogenic ice AIF using CAM5 with a state-of-the-art treatment of ice nucleation. To separate the anthropogenic ice AIF, ice nucleation parameterization is driven by PD and PI prescribed aerosols, respectively, whereas online calculated aerosols with PD aerosol and precursor emissions are used to drive the droplet activation parameterization in both simulations. Compared with the fixed droplet number methodused in Gettelman et al. (2012), this method ensures that each experiment uses the same PD and PI aerosols to drive the ice nucleation process, thus focusing on the anthropogenic ice AIF.

One motivation for the development of cirrus clouds schemes in GCMs is to estimate ice AIF. However, there are same uncertainties in the cirrus cloud microphysics scheme (including ice nucleation parameterization) due to the lack of observation. In this study, we also try to determine the importance of each uncertainty to the estimation of anthropogenic ice AIF.

Sensitivity experiments of different ice nucleation parameterizations (i.e., CTL, ParBN, and ParKL experiments) show that longwave anthropogenic ice AIF (i.e., ΔLWCF) ranges from 0.21 (ParKL) to 0.27 (CTL) and 0.33 W m–2 (ParBN), while shortwave ice anthropogenic AIF (i.e., ΔSWCF) ranges from –0.31 (ParBN) to –0.26 (ParKL) and –0.24 W m–2 (CTL). Although the sensitivity of Nihom to Nso4 in the KL parameterization is much weaker than those in the LP and BN parameterizatons, ΔCDNI from the ParKL experiment (7.58 × 106 m–2) is not much lower than those from the CTL (11.00 × 106 m–2) and ParBN (10.84 × 106 m–2) experiments. One possible reason is that Nihom calcuated from the KL parameterization is sensitive to Nso4 under higher updraft velocity and lower Nso4 conditions, and many homogeneous freezing events occur under these conditions. The global mean ice anthropogenic AIF (i.e., ΔCF) from the ParKL experiment (–0.05 W m–2) is less than the ParBN (0.02 W m–2) and CTL (0.04 W m–2) experiments.

Ice crystals can form heterogeneously on INPs at lower Si, which can prevent the homogeneous freezing from occurring. Although Frehom is decreased due to a higher number of INPs (i.e., INP experiment), the global mean CDNI (182.97 × 106 m–2) is close to the CTL experiment (185.68 × 106 m–2) due to higher Nihet. Furthermore, ΔCDNI (9.10 × 106 m–2) from the INP experiment is not much lower than that from the CTL experiment (11.00 × 106 m–2). One possible reason is that soot number concentration is significantly increased from PI to PD. ΔLWCF (0.14 W m–2) from the INP experiment is significantly less than that from the CTL experiment (0.27 W m–2), whereas ΔSWCF (–0.24 W m–2) from the INP experiment is same as the CTL experiment. Thus, the INP experiment produces a much lower ΔCF (–0.10 W m–2).

Although physically based ice nucleation parameterizations have considered the competition between homogeneous and heterogeneous freezing, to represent ice nucleation in a more realistic manner, it is still necessary for GCMs to restrict homogeneous freezing occurrence (Cziczo et al., 2013). The DHmax method used in this study was introduced in Shi and Liu (2016). DHmax can significantly reduce CDNI and ΔCDNI. ΔLWCF and ΔCF also become weaker after introducing DHmax.

We also investigate the sensitivity of anthropogenic ice AIF to uncertainties in the cloud microphysics scheme. Wsub derived from the turbulent diffusion coefficient can produce extremely high CDNI and ΔCDNI. However, ΔLWCF and ΔSWCF do not become stronger. Reducing Dcs leads to a weaker ΔLWCF and a stronger ΔSWCF, and a much weaker ΔCF. Non-zero μ can significantly impact cirrus cloud properties and enhance ΔLWCF and ΔSWCF. However, ΔCF is significantly decreased due to the non-zero μ. This group of experiments indicates that ice anthropogenic AIF depends on the base state of cirrus clouds.

The last two lines of Table 3 show ensemble statistics. Based on the statistical analysis of all experiments, the global mean ΔLWCF is estimated at 0.25 ± 0.07 W m–2 (one standard deviation uncertainty), ΔSWCF is estimated at –0.29 ± 0.07 W m–2, and total ΔCF is estimated at –0.04 ± 0.06 W m–2. Although both longwave and shortwave ice AIFs are significant, the net ice AIF is rather weak. Note that the impacts of experiments in groups A and B on cloud properties (e.g., Ni, IWP, LWCF, and SWCF) are moderate compared with the experiments in group C (Figs. 3, 4 and Table 2). Including experiments of CTL, groups A and B (i.e., sensitivity experiments in the uncertainties on ice nucleation process), ΔLWCF is estimated at 0.24 ± 0.07 W m–2, ΔSWCF at –0.27 ± 0.04 W m–2, and the net ΔCF at –0.03 ± 0.07 W m–2. These values are similar to those of all experiments.

This study focuses on uncertainties from the cloud microphysics scheme, especially for the treatment of ice nucleation. Sensitivity experiments suggest that improved understandings of the sensitivity of Ni to Nso4, the role of INPs, the spatial and temporal distribution characteristics of homogeneous freezing events, as well as the mean state of cirrus clouds properties, are necessary for constraining the ice AIF. It is worth noting that the uncertainty in aerosols is not considered in this work, which might also play an important role in estimating anthropogenic ice AIF.

Acknowledgments. We thank Kai Zhang for his assistance with prescribed aerosol code.

References
Barahona, D., and A. Nenes, 2009: Parameterizing the competition between homogeneous and heterogeneous freezing in cirrus cloud formation – monodisperse ice nuclei. Atmos. Chem. Phys., 9, 369–381. DOI:10.5194/acp-9-369-2009
Barahona, D., A. Molod, J. Bacmeister, et al., 2014: Development of two-moment cloud microphysics for liquid and ice within the NASA Goddard Earth Observing System Model (GEOS-5). Geosci. Model Dev., 7, 1733–1766. DOI:10.5194/gmd-7-1733-2014
Boucher, O., D. Randall, P. Artaxo, et al., 2013: Clouds and aerosols. Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. T. F. Stocker, D. Qin, G. K. Plattner, et al., Eds., Cambridge University Press, Cambridge, United Kingdom, New York, NY, USA, 571–658.
Carslaw, K. S., L. A. Lee, C. L. Reddington, et al., 2013: Large contribution of natural aerosols to uncertainty in indirect forcing. Nature, 503, 67–71. DOI:10.1038/nature12674
Cziczo, D. J., K. D. Froyd, C. Hoose, et al., 2013: Clarifying the dominant sources and mechanisms of cirrus cloud formation. Science, 340, 1320–1324. DOI:10.1126/science.1234145
Eidhammer, T., H. Morrison, D. Mitchell, et al., 2017: Improvements in global climate model microphysics using a consistent representation of ice particle properties. J. Climate, 30, 609–629. DOI:10.1175/jcli-d-16-0050.1
Field, P. R., A. J. Heymsfield, and A. Bansemer, 2006: Shattering and particle interarrival times measured by optical array probes in ice clouds. J. Atmos. Oceanic Technol., 23, 1357–1371. DOI:10.1175/JTECH1922.1
Gettelman, A., X. Liu, S. J. Ghan, et al., 2010: Global simulations of ice nucleation and ice supersaturation with an improved cloud scheme in the Community Atmosphere Model. J. Geophys. Res., 115, D18216. DOI:10.1029/2009jd013797
Gettelman, A., X. Liu, D. Barahona, et al., 2012: Climate impacts of ice nucleation. J. Geophys. Res. Atmos., 117, D20201. DOI:10.1029/2012jd017950
Heymsfield, A. J., A. Bansemer, P. R. Field, et al., 2002: Observations and parameterizations of particle size distributions in deep tropical cirrus and stratiform precipitating clouds: Results from in situ observations in TRMM field campaigns. J. Atmos. Sci., 59, 3457–3491. DOI:10.1175/1520-0469(2002)059<3457:oapops>2.0.co;2
Jensen, E. J., L. Pfister, T. P. Bui, et al., 2010: Ice nucleation and cloud microphysical properties in tropical tropopause layer cirrus. Atmos. Chem. Phys., 10, 1369–1384. DOI:10.5194/acp-10-1369-2010
Kärcher, B., J. Hendricks, and U. Lohmann, 2006: Physically based parameterization of cirrus cloud formation for use in global atmospheric models. J. Geophys. Res. Atmos., 111, D01205. DOI:10.1029/2005JD006219
Khain, A. P., K. D. Beheng, A. Heymsfield, et al., 2015: Representation of microphysical processes in cloud-resolving models: Spectral (bin) microphysics versus bulk parameterization. Rev. Geophys., 53, 247–322. DOI:10.1002/2014RG000468
Kiehl, J. T., and K. E. Trenberth, 1997: Earth’s annual global mean energy budget. Bull. Amer. Meteor. Soc., 78, 197–208. DOI:10.1175/1520-0477(1997)078<0197:EAGMEB>2.0.CO;2
Krämer, M., C. Schiller, A. Afchine, et al., 2009: Ice supersaturations and cirrus cloud crystal numbers. Atmos. Chem. Phys., 9, 3505–3522. DOI:10.5194/acp-9-3505-2009
Lawson, R. P., 2011: Effects of ice particles shattering on optical cloud particle probes. Atmos. Meas. Tech. Discuss., 4, 939–968. DOI:10.5194/amtd-4-939-2011
Li, J.-L. F., D. E. Waliser, W. T. Chen, et al., 2012: An observationally based evaluation of cloud ice water in CMIP3 and CMIP5 GCMs and contemporary reanalyses using contemporary satellite data. J. Geophys. Res. Atmos., 117, D16105. DOI:10.1029/2012JD017640
Liu, X., and J. E. Penner, 2005: Ice nucleation parameterization for global models. Meteor. Z., 14, 499–514. DOI:10.1127/0941-2948/2005/0059
Liu, X., and X. Shi, 2018: Sensitivity of homogeneous ice nucleation to aerosol perturbations and its implications for aerosol indirect effects through cirrus clouds. Geophys. Res. Lett., 45, 1684–1691. DOI:10.1002/2017GL076721
Liu, X., J. E. Penner, S. J. Ghan, et al., 2007: Inclusion of ice microphysics in the NCAR community atmospheric model version 3 (CAM3). J. Climate, 20, 4526–4547. DOI:10.1175/Jcli4264.1
Liu, X., R. C. Easter, S. J. Ghan, et al., 2012: Toward a minimal representation of aerosols in climate models: Description and evaluation in the Community Atmosphere Model CAM5. Geosci. Model Dev., 5, 709–739. DOI:10.5194/gmd-5-709-2012
Loeb, N. G., B. A. Wielicki, D. R. Doelling, et al., 2009: Toward optimal closure of the earth’s top-of-atmosphere radiation budget. J. Climate, 22, 748–766. DOI:10.1175/2008jcli2637.1
Morrison, H., and W. W. Grabowski, 2008: A novel approach for representing ice microphysics in models: Description and tests using a kinematic framework. J. Atmos. Sci., 65, 1528–1548. DOI:10.1175/2007jas2491.1
Morrison, H., and A. Gettelman, 2008: A new two-moment bulk stratiform cloud microphysics scheme in the community atmosphere model, version 3 (CAM3). Part I: Description and numerical tests. J. Climate, 21, 3642–3659. DOI:10.1175/2008jcli2105.1
Murray, B. J., T. W. Wilson, S. Dobbie, et al., 2010: Heterogeneous nucleation of ice particles on glassy aerosols under cirrus conditions. Nat. Geosci., 3, 233–237. DOI:10.1038/ngeo817
Myhre, G., B. H. Samset, M. Schulz, et al., 2013: Radiative forcing of the direct aerosol effect from AeroCom Phase II simulations. Atmos. Chem. Phys., 13, 1853–1877. DOI:10.5194/acp-13-1853-2013
Neale, R. B., A. Gettelman, S. Park, et al., 2012: Description of the NCAR Community Atmosphere Model (CAM 5.0). Climate and Global Dynamics Division, National Center for Atmospheric Research, Boulder, Colorado, USA.
Penner, J. E., C. Zhou, and X. H. Liu, 2015: Can cirrus cloud seeding be used for geoengineering? Geophys. Res. Lett., 42, 8775–8782. DOI:10.1002/2015gl065992
Rossow, W. B., and R. A. Schiffer, 1999: Advances in understanding clouds from ISCCP. Bull. Amer. Meteor. Soc., 80, 2261–2288. DOI:10.1175/1520-0477(1999)080<2261:aiucfi>2.0.co;2
Seinfeld, J. H., C. Bretherton, K. S. Carslaw, et al., 2016: Improving our fundamental understanding of the role of aerosol–cloud interactions in the climate system. Proc. Natl. Acad. Sci. USA, 113, 5781–5790. DOI:10.1073/pnas.1514043113
Shi, X., and X. Liu, 2016: Effect of cloud-scale vertical velocity on the contribution of homogeneous nucleation to cirrus formation and radiative forcing. Geophys. Res. Lett., 43, 6588–6595. DOI:10.1002/2016gl069531
Shi, X., X. Liu, and K. Zhang, 2015: Effects of pre-existing ice crystals on cirrus clouds and comparison between different ice nucleation parameterizations with the Community Atmosphere Model (CAM5). Atmos. Chem. Phys., 15, 1503–1520. DOI:10.5194/acp-15-1503-2015
Spichtinger, P., and K. M. Gierens, 2009: Modelling of cirrus clouds—Part 2: Competition of different nucleation mechanisms. Atmos. Chem. Phys., 9, 2319–2334. DOI:10.5194/acp-9-2319-2009
Wielicki, B. A., B. R. Barkstrom, E. F. Harrison, et al., 1996: Clouds and the earth’s radiant energy system (CERES): An earth observing system experiment. Bull. Amer. Meteor. Soc., 77, 853–868. DOI:10.1175/1520-0477(1996)077<0853:CATERE>2.0.CO;2
Zhang, K., X. H. Liu, J. H. Yoon, et al., 2013a: Assessing aerosol indirect effect through ice clouds in CAM5. AIP Conf. Proc., 1527, 751. DOI:10.1063/1.4803379
Zhang, K., X. Liu, M. Wang, et al., 2013b: Evaluating and constraining ice cloud parameterizations in CAM5 using aircraft measurements from the SPARTICUS campaign. Atmos. Chem. Phys., 13, 4963–4982. DOI:10.5194/acp-13-4963-2013