1. Introduction
Cooper (1989) theoretically explored the implications of turbulent fluctuations on droplet size distributions, and recent experimental (Chandrakar et al. 2016; Prabhakaran et al. 2020; Chandrakar et al. 2020b), field (Gerber 1991; Ditas et al. 2012; Siebert and Shaw 2017; Yang et al. 2019), and numerical (Kulmala et al. 1997a,b; Vaillancourt et al. 2002; Paoli and Shariff 2009; Abade et al. 2018; Sardina et al. 2018; Li et al. 2018; Saito et al. 2019) studies have demonstrated the importance of scalar fluctuations caused by turbulence on activation, condensational growth and deactivation processes for aerosol and cloud particles, in addition to the mean supersaturation (Yau and Rogers 1996; Krueger 2020). Thus, an accurate representation of the supersaturation variability is required to capture the cloud microphysics effects (Hoffmann et al. 2019).
In modeling studies, if supersaturation is treated as a random variable at all, its probability density function (PDF) is usually treated as Gaussian (Sardina et al. 2018; Saito et al. 2019; Chandrakar et al. 2018), similar to scalars like temperature and water vapor mixing ratio. However, the supersaturation PDF is dependent on the process by which supersaturation is produced. In a parcel view of the atmospheric clouds, supersaturation can be produced by the vertical ascent of parcels (Yau and Rogers 1996) and by the isobaric mixing of parcels (Korolev and Mazin 2003). Cloud entrainment (Korolev and Isaac 2000; Pinsky and Khain 2018) and cloud-free Rayleigh–Bénard convection (Saito et al. 2019; Zhang et al. 2019; Chandrakar et al. 2020a) are examples of processes that can produce supersaturation via isobaric mixing, occurring both in nature and in the laboratory. For the current study, we focus on supersaturation generated via mixing processes and cloud-free Rayleigh–Bénard convection (RBC) is an ideal surrogate for such processes. RBC can be considered the simplest model of the subgrid-scale mixing within a typical cloud large-eddy simulation (LES), for example. It is further advantageous because it efficiently produces a statistically stationary thermodynamic state corresponding to the mixing processes. Furthermore, an atmospheric LES model can be modified to simulate cloud-free RBC to exclusively study mixing processes without any complexities and uncertainties involving cloud-supersaturation feedback interactions and boundary forcings. This model not only serves as the test bed to reveal insights into the nature of supersaturation PDF produced by mixing processes in the absence of cloud droplets, but also helps in validating a computationally inexpensive Gaussian mixing model introduced here.
In this study, we investigate the shape of the supersaturation PDF in the context of atmospheric mixing processes in the absence of cloud droplets using an atmospheric LES and a Gaussian mixing model detailed in section 3. The results are presented in section 4, and atmospheric implications are discussed further in section 5.
2. Theory
a. Scalar equations
Consider Eq. (2); the rate of change of temperature at a point depends on the temperature advected by the fluid motion, the diffusional heat transfer due to local gradients, rate of release/absorption of latent heat due to condensation/evaporation, and finally, any external forcing. Similarly in Eq. (3), for water vapor, all the terms on the right-hand side are analogous to Eq. (2) except for the latent heat effects term, which is replaced by rate of condensation/evaporation of water vapor. Note that we have used temperature (T) instead of pressure compensated potential temperature because isobaric mixing assumes the process to be local in nature. For parcel studies such as Abade et al. (2018), fT and fq represent the change in forcing of temperature and water vapor due to the entrainment of surrounding environmental air into the parcel.
From Eqs. (2) and (3) we gather that for a given flow field, the difference between appropriately normalized temperature and water vapor fields at a location can arise only from one of the following scenarios: (i) differential diffusivity of scalars, (ii) condensation/evaporation processes, and (iii) correlation between fT and fq.
To understand the role of turbulence on the mean supersaturation, one can subtract Eq. (6) from Eq. (7) and the role of turbulence is reflected by the difference. The coefficient of terms with
b. Discussions on the water vapor temperature covariance
In Eq. (14), the first term on the left-hand side is the time evolution of
To evaluate the phase change effects, consider a system with temperature and water vapor transported only by advection processes but including condensation effects. In such a case, any local condensation results in the depletion of water vapor and increase in temperature and vice versa for any local evaporation. Therefore,
The production terms
3. Analysis tools
In this paper, we use two computational approaches to explore supersaturation fluctuations in a turbulent Rayleigh–Bénard convection flow. First, we describe a detailed large-eddy simulation approach, and second, we introduce an idealized Gaussian mixing model based on observed behavior of scalar fields from measurements and numerical studies of Rayleigh–Bénard convection. The latter model also can explore the effect of differential diffusivity, forcings of temperature and water vapor and their correlations on the supersaturation PDF.
a. LES
The System for Atmospheric Modeling (SAM) (Khairoutdinov and Randall 2003) coupled with Hebrew University Spectral Bin Microphysics (Khain et al. 2000; Fan et al. 2009) is configured to simulate the Michigan Tech Pi Cloud Chamber as described in Thomas et al. (2019). We provide a brief discussion of the model for completeness. The RBC system is a 2 m × 2 m ×1 m box modeled as 64 × 64 × 32 grid points with a grid size of 3.125 cm. The convective system is initialized by imposing an unstable temperature gradient and water vapor mixing ratio gradient along the height of the chamber, keeping the top and bottom boundaries saturated. Furthermore, adiabatic conditions for temperature and water vapor mixing ratio are imposed for the sidewalls. Once initialized, the system is allowed to evolve for 2 hours of physical time and the simulation reaches a stationary state in 20 minutes. The results from the last 1 hour of the simulation are used for the analysis presented here.
b. Gaussian mixing model
Figure 1 shows the supersaturation PDFs of four cases with temperature differences of 8, 10, 14, and 18 K. The solid line shows the data obtained from LES at the midplane of the chamber, at least 12.5 cm away from the sidewalls. The dashed lines are the results from the Gaussian mixing model (GMM) with correlation coefficient between T and qυ to be 0.9994 and the constant C set to 3.378. We notice the shapes of the PDFs are qualitatively the same and the modes are shifted by 10% maximum. This level of agreement of the GMM will suffice for exploring the qualitative behavior of the supersaturation distribution under varying assumed T–qυ correlations.
Assuming the correlation coefficient between temperature and water vapor remains the same, the effects of differential diffusivity of scalars are explored by varying Pr and Sc. To understand the effect of scalar forcings, the correlation coefficient between
4. Results
The supersaturation PDFs simulated using LES are shown in Fig. 1, for temperature differences of 8, 10, 14, and 18 K with an initial mean of 283.16 K. Though the bulk temperature and bulk water vapor PDFs are Gaussian in nature, a negatively skewed supersaturation PDF is observed in the bulk of the chamber. From Table 1, it is clear that the magnitude of the skewness is larger at lower temperature differences than at higher values. For LES, the term “bulk” here refers to all the grid cells that are at least 12.5 cm away from the walls of the chamber, in order to avoid the wall effects.
Mean, mode, and skewness of supersaturation for different temperature differences. These results are obtained from the LES starting with a mean temperature of 283.16 K. Note the decrease in supersaturation skewness as ΔT increases.
To understand this negative skewness, the supersaturations obtained from the LES runs are plotted against temperature. A mixing curve obtained by mixing parcels from top and bottom plates, characterized by different temperature and saturated water vapor mixing ratios, in different proportions, is also shown. In Fig. 2 the mixing curves (dashed lines) are plotted in supersaturation and temperature coordinates. The filled circles are the LES results from the bulk of the chamber. For a RBC system without the density effects, the density-weighted mean temperature is the mean temperature between the top and the bottom plates. At low temperature differences, the peak of the mixing curve coincides with the density-weighted mean temperature; hence, the mode of supersaturation is the maximum supersaturation. As the temperature difference increases, the peak of the mixing curve shifts to lower temperature. This leftward shift of the mixing curve arises from the nonlinear nature of the Clausius–Clapeyron equation. Hence, the density-weighted mean temperature in the fluid moves from the maximum to the relatively linear region of the mixing curve. Furthermore, the region of the mixing curve sampled by the bulk increases, due to the increased variance of temperature and water vapor as a result of increase in Rayleigh number. Though density effects (non-Boussinesq effects) can counteract these effects by reducing the positive skewness of the mixing curve and reducing the mean temperature, these effects are negligible for our conditions (refer to Table 1). Please note that all of the mixing line is not populated because only the bulk is sampled, the rest of the mixing line can be sampled from the boundary layer regions near the top and bottom walls.
Figure 3a plots qυ versus T, comparing the Gaussian mixing model (red dotted line) and LES results (blue). Notice that they lie on a straight line joining the points corresponding to the state of the top and bottom plates. Figure 3b compares the supersaturation PDFs obtained from the LES and the GMM. The deviation of the model from the LES results is probably due to the approximation of scalar fluctuations to be Gaussian. The skewness of the temperature data from the LES reveals a slight positive skewness on the order of 0.1, compared to 0.0 for a perfect Gaussian distribution. The NOB effects drive the mean bulk temperature to slightly less than the average of top and bottom plate temperatures; hence, more positive fluctuations arise to reduce this difference.
Figure 4a illustrates the effect of differential diffusivities on supersaturation fluctuations. The case νυ = α = Le−1 = 1, shown in red is the diffusivity formulation ubiquitous across LES and most DNS. As discussed earlier, this result in a negatively skewed distribution of the supersaturation PDF. The physical diffusivities follow Le−1 = 1.16, and the role of differential diffusivity is explored with Le−1 of 0.75 and 1.33. An interesting observation is that except for when Le−1 = 1, the mixing process no longer follows an isobaric mixing line (black dotted line). From the earlier arguments based on the density-weighted mean temperature, it is easy to see that the differential diffusivities do reduce the negative skewness of supersaturation as illustrated in Fig. 4b. However, note that the PDF is still non-Gaussian and negatively skewed with Le−1 = 1.16. A detailed treatment of differential diffusivity and its role in the supersaturation PDF pertaining to RBC can be found in Chandrakar et al. (2020a).
In all the cases discussed above, the forcings of temperature and water vapor, fT and fq, respectively, have a perfect correlation. In Fig. 5a we can see a broad symmetric supersaturation PDF for uncorrelated and anticorrelated forcings of temperature and water vapor. We recall that for a cloud-free RBC system, as described in section 3, the minimum saturation ratio that is allowed is 100%. However, for lower correlation coefficients, saturation ratios are as low as 90% as show in the figure. Any decrease in forcing correlation from a perfect correlation coefficient of 1 (shown in blue) results in a change in the average slope of the distribution of points (Fig. 5a) and an increased spread of the distribution of points around the average slope. From Fig. 5b, it is observed that the spread reaches a maximum when the scalar forcings are perfectly uncorrelated (shown in red) and as they become anticorrelated the spread starts to reduce and falls on a line for correlation coefficient of −1 (shown in green). During this process, the points fall below the limit imposed by the Clausius–Clapeyron line resulting in subsaturated conditions.
Figure 6 illustrates the effect of cloud droplet growth on the supersaturation generated by mixing. The blue dots represent the mixing line in the absence of cloud droplets and red dots represent the mixing in the presence of cloud droplets at the high Damköhler (Da) number limit (Chandrakar et al. 2016). The high Da case is similar to the bulk microphysics limit for which the mixing leads to points collapsing onto the Clausius–Clapeyron line. In Fig. 6 the straight, cloud-free mixing line approaches the Clausius–Clapeyron curve as the Damköhler number increases. The slight deviation the from Clausius–Clapeyron curve can either be the result of a numerical artifact or a physical process and cannot be resolved using the current LES model. This transition requires careful investigation and will be explored in a future study.
5. Discussion and concluding remarks
The comparison and verification of the previously demonstrated numerical results with experiments is the focus of ongoing research. It depends on making measurements of the distribution of supersaturation in a turbulent flow, which is a significant experimental challenge. Very few direct measurements are available from the field (Gerber 1991; Siebert and Shaw 2017). Progress toward in situ measurement of supersaturation in cloud-free Rayleigh–Bénard convection is discussed in Anderson et al. (2021, manuscript submitted to Atmos. Meas. Tech. Discuss.). Efforts for simultaneous remote measurement of temperature and water vapor concentration at sufficiently high precision for obtaining reliable supersaturation estimates are also being made (Capek et al. 2020).
In the current study, we use LES and a Gaussian mixing model to explore the isobaric mixing processes in an idealized turbulent cloud-free Rayleigh–Bénard convection system. In the idealized system we observe the supersaturation PDF to be non-Gaussian and negatively skewed, as shown in Fig. 1. Further, we observe the PDF to be more negatively skewed for smaller temperature differences than at higher temperature differences.
To understand the supersaturation PDF and how it may be generalized to other contexts, we explore the covariance term
A detailed understanding of the effect of condensation/evaporation on
A key point emerging out of the current study is the importance of correlation coefficient between external forcings of temperature and water vapor,
It should be noted that even though cloud-free RBC requires the forcings of temperature and water vapor to follow a relation of the form
For LES studies of the convection-cloud chamber (Chang et al. 2016) such as in Thomas et al. (2019), the boundary fluxes are modeled using Monin–Obukhov similarity theory, resulting in a perfectly correlated forcing from the boundaries. However, in the subgrid-scale model the temperature and water vapor fields are diffused with the same turbulent diffusivity. Therefore, any positive supersaturations arising due to differential diffusivity are not captured, thus impeding the cloud droplet growth. Therefore, the droplet size distributions obtained from such simulations should be at least somewhat narrower than what would arise from experiments or from a DNS accounting for differential diffusivity. In DNS studies that do not account for differential diffusivity effects, such as the cloud parcel studies by Saito et al. (2019) that have only water vapor forcing (refer to the red points in Fig. 7), a broader size distribution of cloud droplets is obtained than warranted by a physically consistent supersaturation field.
Atmospheric models (Clark 1973; Khairoutdinov and Randall 2003) typically use two separate prognostic variables to capture temperature and water vapor. Subsequently, the diagnostic variable—mean supersaturation—is calculated from temperature and water vapor in individual grid boxes ignoring any subgrid-scale variability that is important for cloud droplet activation (Prabhakaran et al. 2020) and growth (Chandrakar et al. 2016). The calculated supersaturation interacts with the microphysics scheme to produce cloud droplet numbers and the corresponding masses or higher moments depending on the scheme’s complexity. Often, DNS studies (Siewert et al. 2017; Sardina et al. 2018; Li et al. 2018) intended to understand the cloud droplet growth in a turbulent environment and treat supersaturation as a prognostic scalar disregarding the nonlinear behavior of the Clausius–Clapeyron equation. The treatment of supersaturation as a scalar is suitable in regimes where the Clausius–Clapeyron equation can be linearly approximated. However, in systems such as Rayleigh–Bénard convection, this is no longer true since the production of mixing supersaturation relies inherently on the nonlinear behavior of the Clausius–Clapeyron equation. Furthermore, there may be scenarios in which differential diffusivity needs to be accounted for, which would lead to the decorrelation of
In the larger context, the concerns about subgrid-scale variability of temperature, water vapor and subsequent microphysics interactions highlighted by Sommeria and Deardorff (1977) and Clark (1973) remains an open challenge even today, even in spite of LES studies with increasing resolution. One approach for addressing the subgrid-scale fluctuations considered by Hoffmann et al. (2019) is the use of a linear-eddy model, although this may be computationally expensive in full implementation. However, the GMM described here may provide a computationally inexpensive but efficient alternative to incorporate physically consistent subgrid-scale variability. A second part of the puzzle, involving supersaturation–cloud particle interactions still needs to be addressed. Reexamination of lateral entrainment studies with the consideration of negatively skewed supersaturation–microphysics interactions in the context of droplet activation and growth can help in answering the latter part of the puzzle.
Acknowledgments
This work was supported by National Science Foundation Grant AGS-1754244. We thank the two anonymous reviewers, Dr. Susan C. van den Heever, Dr. Mikhail Ovchinnikov, and Dr. Fan Yang for suggestions that improved the manuscript. Portage, a high-performance computing infrastructure at Michigan Technological University, was used in obtaining results presented in this publication. The data are permanently archived at MTU Digital Commons, https://digitalcommons.mtu.edu/all-datasets/4/.
APPENDIX
List of Symbols
qυ | Water vapor mixing ratio |
qsat(T) | Saturation vapor mixing ratio at temperature T |
qυ/qsat(T) | Saturation ratio |
Velocity vector of the fluid | |
Rate of condensation/evaporation of water vapor | |
Lυ | Latent heat of vaporization of water |
Cp | Specific heat of air at constant pressure |
fT, fq | External forces on T and qυ |
ρt/b | Density of air at top (t) and bottom (b) |
Tt/b | Temperature at top (t) and bottom (b) |
qυt/b | Water vapor mixing ratio at top (t) and bottom (b) |
νυ | Water vapor diffusivity |
α | Thermal diffusivity |
Le | Lewis number (α/νυ) |
∆T | Temperature difference between top and bottom plate |
∆qυ | Water vapor mixing ratio difference between top and bottom plate |
σT | Standard deviation of temperature T |
Standard deviation of water vapor mixing ratio qυ | |
Sc | Schmidt number (ν/νυ) |
Pr | Prandtl number (ν/α) |
Le | Lewis number (α/νυ) |
ν | Momentum diffusivity |
νυ | Water vapor diffusivity |
α | Thermal diffusivity |
C1 | |
z | Vertical location in the chamber, assumed to be 0.5H |
H | Height of the chamber |
Ram | Moist Rayleigh number |
g | Acceleration due to gravity |
β | Thermal expansion coefficient |
ε | Ratio of gas constants of air and water vapor (≈0.622) |
REFERENCES
Abade, G. C., W. W. Grabowski, and H. Pawlowska, 2018: Broadening of cloud droplet spectra through eddy hopping: Turbulent entraining parcel simulations. J. Atmos. Sci., 75, 3365–3379, https://doi.org/10.1175/JAS-D-18-0078.1.
Capek, T., J. Borysow, C. Mazzoleni, and M. Moraldi, 2020: Toward non-invasive measurement of atmospheric temperature using vibro-rotational Raman spectra of diatomic gases. Remote Sens., 12, 4129, https://doi.org/10.3390/rs12244129.
Chandrakar, K. K., W. Cantrell, K. Chang, D. Ciochetto, D. Niedermeier, M. Ovchinnikov, R. A. Shaw, and F. Yang, 2016: Aerosol indirect effect from turbulence-induced broadening of cloud-droplet size distributions. Proc. Natl. Acad. Sci. USA, 113, 14 243–14 248, https://doi.org/10.1073/pnas.1612686113.
Chandrakar, K. K., W. Cantrell, and R. A. Shaw, 2018: Influence of turbulent fluctuations on cloud droplet size dispersion and aerosol indirect effects. J. Atmos. Sci., 75, 3191–3209, https://doi.org/10.1175/JAS-D-18-0006.1.
Chandrakar, K. K., W. Cantrell, S. Krueger, R. A. Shaw, and S. Wunsch, 2020a: Supersaturation fluctuations in moist turbulent Rayleigh–Bénard convection: A two-scalar transport problem. J. Fluid Mech., 884, A19, https://doi.org/10.1017/jfm.2019.895.
Chandrakar, K. K., I. Saito, F. Yang, W. Cantrell, T. Gotoh, and R. A. Shaw, 2020b: Droplet size distributions in turbulent clouds: Experimental evaluation of theoretical distributions. Quart. J. Roy. Meteor. Soc., 146, 483–504, https://doi.org/10.1002/qj.3692.
Chang, K., and Coauthors, 2016: A laboratory facility to study gas–aerosol–cloud interactions in a turbulent environment: The π chamber. Bull. Amer. Meteor. Soc., 97, 2343–2358, https://doi.org/10.1175/BAMS-D-15-00203.1.
Clark, T. L., 1973: Numerical modeling of the dynamics and microphysics of warm cumulus convection. J. Atmos. Sci., 30, 857–878, https://doi.org/10.1175/1520-0469(1973)030<0857:NMOTDA>2.0.CO;2.
Cooper, W. A., 1989: Effects of variable droplet growth histories on droplet size distributions. Part I: Theory. J. Atmos. Sci., 46, 1301–1311, https://doi.org/10.1175/1520-0469(1989)046<1301:EOVDGH>2.0.CO;2.
Ditas, F., R. Shaw, H. Siebert, M. Simmel, B. Wehner, A. Wiedensohler, and M. Krämer, 2012: Aerosols-cloud microphysics-thermodynamics-turbulence: Evaluating supersaturation in a marine stratocumulus cloud. Atmos. Chem. Phys., 12, 2459–2468, https://doi.org/10.5194/acp-12-2459-2012.
Fan, J., M. Ovtchinnikov, J. M. Comstock, S. A. McFarlane, and A. Khain, 2009: Ice formation in Arctic mixed-phase clouds: Insights from a 3-D cloud-resolving model with size-resolved aerosol and cloud microphysics. J. Geophys. Res., 114, D04205, https://doi.org/10.1029/2008JD010782.
Gerber, H., 1991: Supersaturation and droplet spectral evolution in fog. J. Atmos. Sci., 48, 2569–2588, https://doi.org/10.1175/1520-0469(1991)048<2569:SADSEI>2.0.CO;2.
Hoffmann, F., T. Yamaguchi, and G. Feingold, 2019: Inhomogeneous mixing in Lagrangian cloud models: Effects on the production of precipitation embryos. J. Atmos. Sci., 76, 113–133, https://doi.org/10.1175/JAS-D-18-0087.1.
Katzwinkel, J., H. Siebert, T. Heus, and R. A. Shaw, 2014: Measurements of turbulent mixing and subsiding shells in trade wind cumuli. J. Atmos. Sci., 71, 2810–2822, https://doi.org/10.1175/JAS-D-13-0222.1.
Khain, A., M. Ovtchinnikov, M. Pinsky, A. Pokrovsky, and H. Krugliak, 2000: Notes on the state-of-the-art numerical modeling of cloud microphysics. Atmos. Res., 55, 159–224, https://doi.org/10.1016/S0169-8095(00)00064-8.
Khairoutdinov, M. F., and D. A. Randall, 2003: Cloud resolving modeling of the ARM summer 1997 IOP: Model formulation, results, uncertainties, and sensitivities. J. Atmos. Sci., 60, 607–625, https://doi.org/10.1175/1520-0469(2003)060<0607:CRMOTA>2.0.CO;2.
Korolev, A. V., and G. A. Isaac, 2000: Drop growth due to high supersaturation caused by isobaric mixing. J. Atmos. Sci., 57, 1675–1685, https://doi.org/10.1175/1520-0469(2000)057<1675:DGDTHS>2.0.CO;2.
Korolev, A. V., and I. P. Mazin, 2003: Supersaturation of water vapor in clouds. J. Atmos. Sci., 60, 2957–2974, https://doi.org/10.1175/1520-0469(2003)060<2957:SOWVIC>2.0.CO;2.
Kostinski, A., 2009: Simple approximations for condensational growth. Environ. Res. Lett., 4, 015005, http://doi.org/10.1088/1748-9326/4/1/015005.
Krueger, S. K., 2020: Equilibrium droplet size distributions in a turbulent cloud chamber with uniform supersaturation. Atmos. Chem. Phys., 20, 7895–7909, https://doi.org/10.5194/acp-20-7895-2020.
Kulmala, M., A. Laaksonen, R. J. Charlson, and P. Korhonen, 1997a: Clouds without supersaturation. Nature, 388, 336–337, https://doi.org/10.1038/41000.
Kulmala, M., Ü. Rannik, E. L. Zapadinsky, and C. F. Clement, 1997b: The effect of saturation fluctuations on droplet growth. J. Aerosol Sci., 28, 1395–1409, https://doi.org/10.1016/S0021-8502(97)00015-3.
Li, X.-Y., G. Svensson, A. Brandenburg, and N. E. Haugen, 2018: Cloud-droplet growth due to supersaturation fluctuations in stratiform clouds. Atmos. Chem. Phys., 19, 639–648, https://doi.org/10.5194/acp-19-639-2019.
Mellado, J. P., 2017: Cloud-top entrainment in stratocumulus clouds. Annu. Rev. Fluid Mech., 49, 145–169, https://doi.org/10.1146/annurev-fluid-010816-060231.
Niemela, J., L. Skrbek, K. Sreenivasan, and R. Donnelly, 2000: Turbulent convection at very high Rayleigh numbers. Nature, 404, 837–840, https://doi.org/10.1038/35009036.
Paoli, R., and K. Shariff, 2009: Turbulent condensation of droplets: Direct simulation and a stochastic model. J. Atmos. Sci., 66, 723–740, https://doi.org/10.1175/2008JAS2734.1.
Pinsky, M., and A. Khain, 2018: Theoretical analysis of the entrainment–mixing process at cloud boundaries. Part I: Droplet size distributions and humidity within the interface zone. J. Atmos. Sci., 75, 2049–2064, https://doi.org/10.1175/JAS-D-17-0308.1.
Prabhakaran, P., A. S. M. Shawon, G. Kinney, S. Thomas, W. Cantrell, and R. A. Shaw, 2020: The role of turbulent fluctuations in aerosol activation and cloud formation. Proc. Natl. Acad. Sci. USA, 117, 16 831–16 838, https://doi.org/10.1073/pnas.2006426117.
Saito, I., T. Gotoh, and T. Watanabe, 2019: Broadening of cloud droplet size distributions by condensation in turbulence. J. Meteor. Soc. Japan, 97, 867–891, https://doi.org/10.2151/jmsj.2019-049.
Sardina, G., S. Poulain, L. Brandt, and R. Caballero, 2018: Broadening of cloud droplet size spectra by stochastic condensation: Effects of mean updraft velocity and CCN activation. J. Atmos. Sci., 75, 451–467, https://doi.org/10.1175/JAS-D-17-0241.1.
Siebert, H., and R. A. Shaw, 2017: Supersaturation fluctuations during the early stage of cumulus formation. J. Atmos. Sci., 74, 975–988, https://doi.org/10.1175/JAS-D-16-0115.1.
Siewert, C., J. Bec, and G. Krstulovic, 2017: Statistical steady state in turbulent droplet condensation. J. Fluid Mech., 810, 254–280, https://doi.org/10.1017/jfm.2016.712.
Sommeria, G., and J. Deardorff, 1977: Subgrid-scale condensation in models of nonprecipitating clouds. J. Atmos. Sci., 34, 344–355, https://doi.org/10.1175/1520-0469(1977)034<0344:SSCIMO>2.0.CO;2.
Sreenivasan, K. R., 2019: Turbulent mixing: A perspective. Proc. Natl. Acad. Sci. USA, 116, 18 175–18 183, https://doi.org/10.1073/pnas.1800463115.
Thomas, S., M. Ovchinnikov, F. Yang, D. van der Voort, W. Cantrell, S. K. Krueger, and R. A. Shaw, 2019: Scaling of an atmospheric model to simulate turbulence and cloud microphysics in the Pi chamber. J. Adv. Model. Earth Syst., 11, 1981–1994, https://doi.org/10.1029/2019MS001670.
Vaillancourt, P., M. Yau, P. Bartello, and W. W. Grabowski, 2002: Microscopic approach to cloud droplet growth by condensation. Part II: Turbulence, clustering, and condensational growth. J. Atmos. Sci., 59, 3421–3435, https://doi.org/10.1175/1520-0469(2002)059<3421:MATCDG>2.0.CO;2.
Wyngaard, J. C., 2010: Turbulence in the Atmosphere. Cambridge University Press, 406 pp.
Yang, F., R. McGraw, E. P. Luke, D. Zhang, P. Kollias, and A. M. Vogelmann, 2019: A new approach to estimate supersaturation fluctuations in stratocumulus cloud using ground-based remote-sensing measurements. Atmos. Meas. Tech., 12, 5817–5828, https://doi.org/10.5194/amt-12-5817-2019.
Yau, M. K., and R. R. Rogers, 1996: A Short Course in Cloud Physics. Elsevier, 304 pp.
Zhang, L., K. L. Chong, and K.-Q. Xia, 2019: Moisture transfer by turbulent natural convection. J. Fluid Mech., 874, 1041–1056, https://doi.org/10.1017/jfm.2019.463.