1. Introduction
Clouds consisting exclusively of ice particles, so-called cirrus clouds, account for roughly one-third of the total cloud cover (e.g., Gasparini et al. 2018), yet their net radiative effect is still one major source of uncertainty in the climate system. Since the albedo effect and greenhouse effect are on the same order of magnitude for those clouds, microphysical details of ice crystals (as, e.g., shape or size; see Zhang et al. 1999; Krämer et al. 2020) may determine the net radiative effect. The microphysical properties, however, are strongly influenced by a complex interplay of nuclei composition, microscale cloud processes, and multiscale interactions with the surrounding atmosphere. All those components are poorly understood and, if at all, only crudely represented in climate models.
Cirrus clouds can be subdivided into liquid origin and in situ cirrus (e.g., Krämer et al. 2016). The former class describes clouds originating from cloud droplets, which freeze in upward motions, e.g., in mesoscale convective outflow or warm conveyor belts. In contrast, the ice crystals of in situ cirrus are formed without any preexisting cloud droplets: either by homogeneous freezing of aqueous solution droplets (short: homogeneous nucleation; see, e.g., Koop et al. 2000; Baumgartner et al. 2022), or by heterogeneous nucleation (e.g., Pruppacher and Klett 2010; Hoose and Möhler 2012; Baumgartner et al. 2022) initiated by solid aerosol particles.
Observational studies indicate that cirrus properties and life cycle can be crucially affected by gravity wave (GW) dynamics (e.g., Kärcher and Ström 2003; Kim et al. 2016; Bramberger et al. 2022). GWs are generated to a large fraction in the troposphere and often propagate over considerable horizontal and vertical distances before breaking. During their propagation those waves can generate substantial oscillations in the atmospheric fields and the GW drag, exerted in the breaking region, alters the mean atmospheric state. Because of the importance of small-scale GW dynamics some cirrus studies explicitly resolve the GWs using LES models (e.g., Joos et al. 2009; Kienast-Sjögren et al. 2013), or detailed parcel models (e.g., Haag and Kärcher 2004; Jensen and Pfister 2004; Spichtinger and Krämer 2013).
There are several classes of schemes for modeling the influence of GWs on ice clouds in climate models. In all schemes, a subgrid-scale GW vertical velocity is diagnosed and then directly used in the cirrus cloud scheme. One should keep in mind that climate models diagnose the number concentration of ice crystals in a (homogeneous) nucleation event from the vertical velocity (see, e.g., Kärcher and Lohmann 2002; Ren and Mackenzie 2005; Wang and Penner 2010). First, there are schemes without any physical constraint on GWs. These schemes mostly rely on the use of a turbulent kinetic energy (TKE) scheme in the upper troposphere. This approach is per se questionable since most TKE schemes were developed for parameterizing turbulence in the planetary boundary layer. The TKE approach is known to produce quite high vertical velocities (see, e.g., Joos et al. 2008; Zhou et al. 2016) and the pattern of enhanced vertical velocities usually do not agree with regions of enhanced GW activity (see Fig. 5 in Joos et al. 2008). Second, there are schemes using distributions of temperature fluctuations constructed from measurements. From these subgrid-scale vertical velocities are derived (Kärcher and Burkhardt 2008; Wang and Penner 2010; Podglajen et al. 2016; Kärcher and Podglajen 2019). In this approach it is not possible to isolate solely the contribution from GWs. Rather than this, one is considering the combined effect of GWs and other fluctuations (e.g., turbulence), which might modify or even mask the GW contribution. In addition, there is no direct link to the GW sources, such as mountains, convection, spontaneous imbalance, or others. One should also keep in mind that the measurements are sparse and they are largely extrapolated into other regions without any measurements. Further, it is not clear a priori that the underlying statistical description of temperature fluctuations will remain unaltered under climate change. Finally, there have been two attempts to diagnose the GW vertical velocity using linear theory for mountain waves (Dean et al. 2007; Joos et al. 2008). In both schemes there is a direct connection between the source of GWs and the cloud scheme, which is a clear advantage in comparison to schemes relying on statistical information. However, no other sources than mountain waves are represented up to now.
Most of the current GW parameterizations in climate models rely on the single-column, steady-state approximation. Under this assumption GWs propagate only in the vertical and instantly fast up to the breaking altitude, where they deposit energy and momentum. The limitations of steady-state parameterizations were demonstrated in the study by Bölöni et al. (2016), where a transient approach was proposed. The new transient parameterization was implemented by Bölöni et al. (2021) in the weather-forecast and climate model ICON and Kim et al. (2021) showed that the resulting intermittency patterns of convectively generated GWs are similar to observations. Thus, developing a cirrus scheme to be coupled to a transient GW parameterization is a promising route for more realistic representation of ice clouds in climate models. Such development requires the systematic identification of the dominant interaction processes between GWs and cirrus and their self-consistent description. Baumgartner and Spichtinger (2019, hereafter BS19), utilized a matched-asymptotic approach [e.g., see Holmes (2013) for an introduction to asymptotics] for studying homogeneous nucleation due to constant updraft velocities. The resulting parameterization successfully reproduces the results of the classical scheme of Kärcher and Lohmann (2002). Encouraged by the results of BS19, we extend their asymptotic approach to allow for GW dynamics. We construct a self-consistent simplified model for GW–cirrus interactions and corresponding asymptotic solutions applicable for diagnosing ice crystal numbers in nucleation events forced by passing GWs. An application of our analytical approach would be a direct coupling of the transient GW parameterization (Bölöni et al. 2021; Kim et al. 2021) to our analytical model. The GW parameterization will provide information about the wave amplitudes, frequencies, and wavenumbers, which can directly be used for our approach. The detailed information on the wave spectrum allows one to predict the ice crystal number concentration more realistically than simple diagnostic relations in large-scale models (e.g., Kärcher and Lohmann 2002), which are based on constant vertical updraft motion.
This paper is organized as follows: The unified asymptotic representation of the GW and ice microphysics can be found in section 2. In section 3 we derive asymptotic solutions, modeling the dynamics during a nucleation event as well as the pre- and postnucleation dynamics. The reduced model for the GW–cirrus interactions and the corresponding asymptotic solutions are summarized in sections 3g and 3h, respectively. The numerical simulations of the full ice physics model and validation of the asymptotic solutions can be found in section 4. In section 5 the present approach is extended to take into account variations of the (ice crystal distribution) mean mass in the deposition. In section 6 the asymptotic solution is extended to the case of multiple GWs driving the ice physics. Concluding discussions are summarized in section 7.
2. Asymptotic approach for studying GW–cirrus interactions
a. Gravity wave dynamics: Governing equations and scalings
Within the framework of multiscale asymptotics, we have to specify a distinguished limit in order to define the regimes we are interested in. This is carried out in the following way:
Reference quantities for high-frequency, β = 0, and midfrequency, β = 1, gravity wave scaling. Here we choose T00 = 210 K and N = 10−2 s−1 for the troposphere and N = 2 × 10−2 s−1 for the tropopause region. Note, however, that regimes with other values of T00 and N can be considered as well.
b. Ice microphysics: Governing equations and scalings
The cirrus clouds are described by a double-moment bulk microphysics scheme assuming a unimodal ice mass distribution function. The scheme is the same as the one from BS19, except that the sedimentational sinks are included here. A more detailed description of the ice model can be found in Spichtinger and Gierens (2009) and Spreitzer et al. (2017); in this section we only briefly refer to some key properties. As in BS19 we assume spherical shape of ice crystals, which leads to a simpler description of the cloud processes.
Parameters of the ice physics scheme.
Reference quantities used for nondimensionalization of the ice physics scheme.
Distinguished limits for the nondimensional numbers in the ice scheme. In the rightmost column, an asterisk denotes an order-one constant. For the nondimensionalization, a time scale Tw, with Tw ∼ Td, was used.
c. Asymptotic expansion
d. Coupling of the GW and diffusion time scale
3. Reduced model of GW–cirrus interactions
a. GW dynamics
b. Single-parcel model approximation and single monochromatic GW
c. The different regimes in the ice dynamics
The gravity wave dynamics change the vertical velocity, pressure, and temperature fields in (43) and hence leads to variations of S and consequently of n. Time series illustrating the qualitative behavior of S and n under GW forcing are shown in Fig. 1 (see the discussion in section 4 for details). A typical situation observed is that S fluctuates until it reaches (or approaches sufficiently) the critical value Sc at time t0. At t0 the nucleation term in (41) leads to an explosive production of ice crystals. The increased number concentration n implies a reduction of S below Sc through the diffusional growth term in (43). After this reduction S continues to fluctuate due to the GW forcing and might again approach Sc. Thus, in some cases we have to consider ice nucleation in the presence of preexisting ice crystals, which might be suppressed under certain conditions.
Time evolution of number concentration n and saturation ratio S for two different initial conditions: (a) n(0) = 0 and (b) n(0) = 2 × 106 kg−1. Tw = 500 s and initial phase of the wave
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
Following the matched asymptotic approach of BS19, three different regimes are considered here. First, the prenucleation regime with S < Sc, where the dynamics takes place on the GW time scale. This is followed by a nucleation regime, centered around time t0 with S(t0) = Sc and dynamics on the much faster nucleation time scale. After the nucleation event the postnucleation regime is entered with S < Sc, characterized again by dynamics on the GW time scale.
We observe that due to the assumption
d. Pre- and postnucleation regime
e. Nucleation regime
f. Matching
g. Summary of the reduced model
h. Summary of the asymptotic solution
4. Numerical experiments and discussion of the asymptotic solution
In this section the asymptotic model is validated against the full ice microphysics model and the reduced model for realistic parameter values taken from BS19 and summarized in Table 2. The full model solves (37)–(39) omitting only sedimentation, the details of it can be found in appendix E. All models are forced with single monochromatic GW from section 3b. The GW vertical velocity amplitude is set to 1 ms−1 corresponding to 0.7Wc, where Wc is the critical vertical velocity amplitude for breaking due to static instability. The GW frequency is
The results for two different initial number concentrations are summarized in Fig. 1. Figure 1a shows a situation where the asymptotic solution reproduces with a high accuracy the time evolution of the ice crystal number concentration and saturation ratio. Figure 1b depicts a case where the nucleation time t0 from the full model is slightly missed by the asymptotic and the reduced models. Although the overall evolution of S is reproduced well, the nucleated ice crystal number is underestimated by roughly 15%. From the asymptotic theory, we expect that the discrepancy will vanish in the limit ε → 0. This asymptotic limit is verified numerically by considering smaller values of ε in the full and in the reduced model, this corresponds to increasing the time scale separation between the different processes in the models [see Dolaptchiev et al. (2013) for another example of this procedure]. The results are summarized in Fig. 2 for ε = 10−1 and ε = 10−2; note that in Fig. 1 ε = 1 implying no increased time scale separation. Figure 2 shows that the models converge quickly to the asymptotic limit already for moderately small values of ε.
As in Fig. 1b, but for (a) ε = 10−1 and (b) ε = 10−2; in Fig. 1b ε = 1.
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
Since the phase of the GW is typically unknown in coarse models, we study the sensitivity of the results with respect to this parameter. For that purpose, we vary the GW phase at the initial time t = 0 and determine the nucleated ice crystals within one wave period for an initial condition n(0) = 0. From Fig. 3 it is visible that the asymptotic solution captures, for all GW phases, the number of nucleated ice crystals in the full model. In addition, the values of n are limited from above by the asymptotic estimate
Nucleated number concentration n as a function of the initial GW phase. Gray horizontal lines denote the initial condition n(0) and the asymptotic estimates
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
In Fig. 4 the normalized vertical velocity at the nucleation time is displayed. It suggests that nucleation takes place at sufficiently high updrafts but not necessary at the maximal.
Normalized GW vertical velocity at t0 as a function of the initial GW phase for the simulations from Fig. 3.
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
Figure 5 summarizes the dependence of the nucleated ice crystals on the initial number concentration. The asymptotic solution reproduces nearly exactly n from the reduced model. Both models are very close to the full model for n(0) well below
As in Fig. 3, but for initial number concentration (a) n(0) = 106 kg−1 and (b) n(0) = 2 × 106 kg−1. In (b), the initial n is the same as the one used in Figs. 1b and 2; the dashed gray vertical line marks the initial phase used for the simulation in Figs. 1b and 2.
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
5. The effect of variable ice crystal mean mass in the deposition
Time evolution of number concentration n, saturation ratio S, and mean ice mass m computed for different models: the reduced model with constant mass, (93)–(95); the reduced model with variable mass, (104)–(106); the asymptotic solution for constant mass (see section 3h); and the asymptotic solution with variable mass correction, (117). Note that for the latter model, only the pre- and postnucleation values are plotted, resulting in a jump at t0. The mean mass in the reduced models is diagnosed using m = q/n. In the figure’s legend, m0 and m(t) denote models with constant mass and variable mass, respectively.
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
In Fig. 6 an increase of the ice crystal mass is observed before the nucleation event. Taking this into account with the model (104)–(106) results in a larger number of nucleated ice particles, as compared to the constant mass model. The asymptotic solution for the constant mass case (denoted with “asym. m0” in the figure) reproduces the behavior of the constant mass reduced model and underestimates n, as well. In the following, we extend the asymptotic approach to allow for variable mean mass effects.
a. Prenucleation regime
The fully coupled system (104)–(106) involves an additional fast time scale in the q equation as compared to the constant mean mass case discussed in section 3. The new time scale will induce in general nontrivial dynamics on longer time scales; however, the corresponding rigorous asymptotic analysis is out of the scope of this paper. As we will show here, the asymptotic prenucleation solution for the constant mass case can still be used for variety of configurations if appropriate corrections are introduced.
b. Nucleation regime
c. Numerical results
Equation (117) is used to find an asymptotic approximation of the nucleated number concentration. The comparison with the numerical results is shown in Fig. 6. The current procedure produces a larger number of nucleated ice crystals as compared with the constant mass model; the magnitude of n is close to the one of the variable mean mass model.
The performance of the current approach is systematically evaluated by varying the initial GW phase. The corresponding results are summarized in Fig. 7. The figure suggests that the proposed procedure captures the nucleated number of ice crystals for various initial GW phases.
A note of caution should be added on the relevance of the variable mean mass model presented in this section. The use of the prenucleation S from (66) in (107) might be invalid on longer time scales: see the first paragraph of section 5a. Further, Fig. 6 suggests that there might be situations with considerable growth of m before nucleation takes place. Obviously, for large m the sedimentation term will provide a sink for the mean mass: see (42). The correct incorporation of the sedimentation effects will be the subject of a future study.
6. Ice physics forced by superposition of gravity waves
The results are summarized in Fig. 8a for the case where ωj is drawn from a narrow range around the frequency
Nucleated number concentration n for 103 different realizations of a superposition of GWs with frequencies: (a)
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
Probability density function (PDF) of n from the realizations presented in Fig. 8. The PDF was constructed using kernel density estimation with Gaussian kernel.
Citation: Journal of the Atmospheric Sciences 80, 12; 10.1175/JAS-D-22-0234.1
Next, we consider the full frequency range of GWs. The corresponding results are summarized in Fig. 8b. It has to be stressed that our asymptotic analysis is only valid for the frequency range around
7. Conclusions
We present an asymptotic approach allowing us to identify a reduced model for the self-consistent description of ice physics forced by a superposition of GWs including the effect of diffusional growth and homogeneous nucleation of ice crystals. Furthermore, using matched asymptotic techniques analytical solutions are constructed, involving a novel parameterization (90) for the ice crystal number concentration n. The latter has as input parameters the wave amplitudes and phases, and the time of the nucleation event. It allows the derivation of an upper bound for the nucleated n, as well as a threshold for the initial n that would inhibit nucleation. The numerical simulations with a Lagrangian parcel model show that the parameterization reproduces nucleation events triggered by a monochromatic GW for a variety of initial conditions. Furthermore, in the case of superposition of GWs within the midfrequency range the parameterization generates a distribution of n matching the one of the full model. By extending the parameterization to high-frequency GWs, it is shown that the asymptotic solution produces distribution similar to the one of the full model even if the complete GW frequency spectrum is used as forcing. The results presented here demonstrate the potential of our approach for constructing improved cirrus schemes in climate models with realistic GW variability as simulated with transient GW parameterizations (Bölöni et al. 2021; Kim et al. 2021).
When comparing the treatment of the ice physics in our approach with the one from BS19, we observe different scaling in the nucleation term: in the latter work J ∼ B ∼ ε−1 is used, whereas here we apply J ∼ ε−1, B ∼ ε−2. Nevertheless, our parameterization (90) is equivalent to the closure of BS19 for constant updraft velocity if the velocity there is replaced by the GW vertical velocity at the nucleation time t0. This is not surprising, since the GW nearly does not vary on the fast nucleation time scale. The correspondence of the two parameterizations becomes more clear if one takes into account that in BS19 ε = O(10−2) and here ε = O(10−1), implying the same magnitude of the nucleation exponent B under the different scalings. As shown by Spichtinger et al. (2023) the exact value of the nucleation rate J is not crucial as long as it is sufficiently large. Still, we have to stress that the present approach generalizes the framework of BS19 to include wave dynamics and the consistency between the two parameterizations only supports our results. In addition, we derive a novel parameterization for the variable mean mass model, see (117), a threshold for nucleation inhibition,
The present asymptotic solutions are applicable mainly to the midfrequency GW in the troposphere and tropopause region, as well as to high-frequency GWs in the troposphere. For high-frequency GWs in the tropopause region, the ratio between the time scale of the diffusional growth and of the wave is given by Td/Tw ∼ ε−1. The simulations from section 6 show larger values of n if the high-frequency GWs are included. This suggests a new regime dominated by the GW forcing term and we propose some asymptotic corrections to account for it. We expect that this regime corresponds to the temperature-limit events studied in Dinh et al. (2016). For low-frequency GWs, the scaling Td/Tw ∼ ε is appropriate. In this case the GW forcing term becomes weaker by a factor of ε, when compared to the depositional growth term. This regime is relevant for low updraft velocities and will be considered in an upcoming study.
Our asymptotic analysis assumes a reference number concentration nc. However, the results from section 6 suggest that the resulting asymptotic model is valid for a wider range of n. If regimes with other values of nc are of interest, the present asymptotic framework can be adapted for the systematic investigation of these, too.
The models presented here predict for the particular GW forcings investigated, values of n which are within or at the upper range of observations; e.g., see Fig. 8 from Krämer et al. (2020). However, a direct comparison with observational data is hampered for two main reasons. First, most of the measurements lack information on the wave properties, e.g., wave amplitude and frequency, so the GW forcing cannot be determined. Second, nucleation takes place at a very fast time scale and within a confined spatial region. Therefore, the vast majority of measurements of ice crystals are taken probably after the nucleation event happened. However, at later stages of the ice cloud life cycle, other processes such as sedimentation determine the microphysical properties. The latter lead to smearing of the clear nucleation signature and to significantly smaller n values (see, e.g., Spichtinger and Gierens 2009); this effect is enhanced if ice crystals fall into subsaturated air and thus evaporate. This might explain why high number densities are quite rarely observed (see, e.g., Krämer et al. 2009, 2020).
In the present regime the magnitude of the sedimentation effects is determined by the sedimentation time scale
In the present study, only cirrus formed by homogeneous nucleation are considered, since this is the dominant formation mechanism in the cold temperature regime with strong updraft velocities (e.g., Heymsfield and Miloshevich 1993). Still, heterogeneous nucleation can considerably alter cirrus formation (see, e.g., Gierens 2003; Spichtinger and Cziczo 2010); however, the important feature is, also in case of competing nucleation pathways, the occurrence of preexisting ice crystals, as in our investigations. In addition, turbulence due to GW breaking is another source of GW-generated variability omitted in the present study (e.g., Atlas and Bretherton 2023).
Acknowledgments.
UA and PS thank the German Research Foundation (DFG) for partial support through the research unit “Multiscale Dynamics of Gravity Waves” (MS-GWaves; Grants AC 71/8-2, AC 71/9-2, AC 71/12-2, and SP 1163/5-2) and CRC 301 “TPChange” (Project-ID 428312742, Projects B06 “Impact of small-scale dynamics on UTLS transport and mixing,” B07 “Impact of cirrus clouds on tropopause structure,” and Z03 “Joint model development”). UA acknowledges DFG for partial support through CRC 181 “Energy transfers in Atmosphere an Ocean” (Project Number 274762653, Projects W01 “Gravity-wave parameterization for the atmosphere” and S02 “Improved Parameterizations and Numerics in Climate Models”).
Data availability statement.
The Python script used to generate all figures in the paper is available upon request.
APPENDIX A
Time Evolution of S
APPENDIX B
GW Dispersion Relation and Polarization Relations
APPENDIX C
Evolution Equation for n in the Nucleation Regime
APPENDIX D
Composite Solution
APPENDIX E
Description Lagrangian Parcel Model
From (E1) with initial condition z = z00 and x = x00 the parcel position is found. With this the wave fluctuations of Exner pressure π′ and of potential temperature θ′ are determined from the corresponding polarization relations. To those fluctuations one has to add the stationary contributions
REFERENCES
Achatz, U., 2022: Atmospheric Dynamics. Springer, 554 pp., https://doi.org/10.1007/978-3-662-63941-2.
Achatz, U., R. Klein, and F. Senf, 2010: Gravity waves, scale asymptotics and the pseudo-incompressible equations. J. Fluid Mech., 663, 120–147, https://doi.org/10.1017/S0022112010003411.
Achatz, U., B. Ribstein, F. Senf, and R. Klein, 2017: The interaction between synoptic-scale balanced flow and a finite-amplitude mesoscale wave field throughout all atmospheric layers: Weak and moderately strong stratification. Quart. J. Roy. Meteor. Soc., 143, 342–361, https://doi.org/10.1002/qj.2926.
Atlas, R., and C. S. Bretherton, 2023: Aircraft observations of gravity wave activity and turbulence in the tropical tropopause layer: Prevalence, influence on cirrus clouds, and comparison with global storm-resolving models. Atmos. Chem. Phys., 23, 4009–4030, https://doi.org/10.5194/acp-23-4009-2023.
Baumgartner, M., and P. Spichtinger, 2019: Homogeneous nucleation from an asymptotic point of view. Theor. Comput. Fluid Dyn., 33, 83–106, https://doi.org/10.1007/s00162-019-00484-0.
Baumgartner, M., R. Weigel, A. H. Harvey, F. Plöger, U. Achatz, and P. Spichtinger, 2020: Reappraising the appropriate calculation of a common meteorological quantity: Potential temperature. Atmos. Chem. Phys., 20, 15 585–15 616, https://doi.org/10.5194/acp-20-15585-2020.
Baumgartner, M., C. Rolf, J.-U. Grooß, J. Schneider, T. Schorr, O. Möhler, P. Spichtinger, and M. Krämer, 2022: New investigations on homogeneous ice nucleation: The effects of water activity and water saturation formulations. Atmos. Chem. Phys., 22, 65–91, https://doi.org/10.5194/acp-22-65-2022.
Bölöni, G., B. Ribstein, J. Muraschko, C. Sgoff, J. Wei, and U. Achatz, 2016: The interaction between atmospheric gravity waves and large-scale flows: An efficient description beyond the nonacceleration paradigm. J. Atmos. Sci., 73, 4833–4852, https://doi.org/10.1175/JAS-D-16-0069.1.
Bölöni, G., Y.-H. Kim, S. Borchert, and U. Achatz, 2021: Toward transient subgrid-scale gravity wave representation in atmospheric models. Part I: Propagation model including nondissipative wave–mean-flow interactions. J. Atmos. Sci., 78, 1317–1338, https://doi.org/10.1175/JAS-D-20-0065.1.
Bramberger, M., and Coauthors, 2022: First super-pressure balloon-borne fine-vertical-scale profiles in the upper TTL: Impacts of atmospheric waves on cirrus clouds and the QBO. Geophys. Res. Lett., 49, e2021GL097596, https://doi.org/10.1029/2021GL097596.
Corcos, M., A. Hertzog, R. Plougonven, and A. Podglajen, 2021: Observation of gravity waves at the tropical tropopause using superpressure balloons. J. Geophys. Res. Atmos., 126, e2021JD035165, https://doi.org/10.1029/2021JD035165.
Corcos, M., A. Hertzog, R. Plougonven, and A. Podglajen, 2023: A simple model to assess the impact of gravity waves on ice-crystal populations in the tropical tropopause layer. Atmos. Chem. Phys., 23, 6923–6936, https://doi.org/10.5194/acp-23-6923-2023.
Dean, S. M., J. Flowerdew, B. N. Lawrence, and S. D. Eckermann, 2007: Parameterisation of orographic cloud dynamics in a GCM. Climate Dyn., 28, 581–597, https://doi.org/10.1007/s00382-006-0202-0.
Dinh, T., A. Podglajen, A. Hertzog, B. Legras, and R. Plougonven, 2016: Effect of gravity wave temperature fluctuations on homogeneous ice nucleation in the tropical tropopause layer. Atmos. Chem. Phys., 16, 35–46, https://doi.org/10.5194/acp-16-35-2016.
Dolaptchiev, S. I., U. Achatz, and I. Timofeyev, 2013: Stochastic closure for local averages in the finite-difference discretization of the forced Burgers equation. Theor. Comput. Fluid Dyn., 27, 297–317, https://doi.org/10.1007/s00162-012-0270-1.
Durran, D. R., 1989: Improving the anelastic approximation. J. Atmos. Sci., 46, 1453–1461, https://doi.org/10.1175/1520-0469(1989)046<1453:ITAA>2.0.CO;2.
Gasparini, B., A. Meyer, D. Neubauer, S. Münch, and U. Lohmann, 2018: Cirrus cloud properties as seen by the CALIPSO satellite and ECHAM-HAM global climate model. J. Climate, 31, 1983–2003, https://doi.org/10.1175/JCLI-D-16-0608.1.
Gierens, K., 2003: On the transition between heterogeneous and homogeneous freezing. Atmos. Chem. Phys., 3, 437–446, https://doi.org/10.5194/acp-3-437-2003.
Haag, W., and B. Kärcher, 2004: The impact of aerosols and gravity waves on cirrus clouds at midlatitudes. J. Geophys. Res., 109, D12202, https://doi.org/10.1029/2004JD004579.
Hertzog, A., M. J. Alexander, and R. Plougonven, 2012: On the intermittency of gravity wave momentum flux in the stratosphere. J. Atmos. Sci., 69, 3433–3448, https://doi.org/10.1175/JAS-D-12-09.1.
Heymsfield, A. J., and L. M. Miloshevich, 1993: Homogeneous ice nucleation and supercooled liquid water in orographic wave clouds. J. Atmos. Sci., 50, 2335–2353, https://doi.org/10.1175/1520-0469(1993)050<2335:HINASL>2.0.CO;2.
Holmes, M. H., 2013: Introduction to Perturbation Methods. Texts in Applied Mathematics, Vol. 20, Springer, 438 pp.
Hoose, C., and O. Möhler, 2012: Heterogeneous ice nucleation on atmospheric aerosols: A review of results from laboratory experiments. Atmos. Chem. Phys., 12, 9817–9854, https://doi.org/10.5194/acp-12-9817-2012.
Jensen, E., and L. Pfister, 2004: Transport and freeze-drying in the tropical tropopause layer. J. Geophys. Res., 109, D02207, https://doi.org/10.1029/2003JD004022.
Joos, H., P. Spichtinger, U. Lohmann, J.-F. Gayet, and A. Minikin, 2008: Orographic cirrus in the global climate model ECHAM5. J. Geophys. Res., 113, D18205, https://doi.org/10.1029/2007JD009605.
Joos, H., P. Spichtinger, and U. Lohmann, 2009: Orographic cirrus in a future climate. Atmos. Chem. Phys., 9, 7825–7845, https://doi.org/10.5194/acp-9-7825-2009.
Kärcher, B., and U. Lohmann, 2002: A parameterization of cirrus cloud formation: Homogeneous freezing of supercooled aerosols. J. Geophys. Res., 107, 4010, https://doi.org/10.1029/2001JD000470.
Kärcher, B., and J. Ström, 2003: The roles of dynamical variability and aerosols in cirrus cloud formation. Atmos. Chem. Phys., 3, 823–838, https://doi.org/10.5194/acp-3-823-2003.
Kärcher, B., and U. Burkhardt, 2008: A cirrus cloud scheme for general circulation models. Quart. J. Roy. Meteor. Soc., 134, 1439–1461, https://doi.org/10.1002/qj.301.
Kärcher, B., and A. Podglajen, 2019: A stochastic representation of temperature fluctuations induced by mesoscale gravity waves. J. Geophys. Res. Atmos., 124, 11 506–11 529, https://doi.org/10.1029/2019JD030680.
Kienast-Sjögren, E., P. Spichtinger, and K. Gierens, 2013: Formulation and test of an ice aggregation scheme for two-moment bulk microphysics schemes. Atmos. Chem. Phys., 13, 9021–9037, https://doi.org/10.5194/acp-13-9021-2013.
Kim, J.-E., and Coauthors, 2016: Ubiquitous influence of waves on tropical high cirrus cloud. Geophys. Res. Lett., 43, 5895–5901, https://doi.org/10.1002/2016GL069293.
Kim, Y.-H., G. Bölöni, S. Borchert, H.-Y. Chun, and U. Achatz, 2021: Toward transient subgrid-scale gravity wave representation in atmospheric models. Part II: Wave intermittency simulated with convective sources. J. Atmos. Sci., 78, 1339–1357, https://doi.org/10.1175/JAS-D-20-0066.1.
Koop, T., B. Luo, A. Tsias, and T. Peter, 2000: Water activity as the determinant for homogeneous ice nucleation in aqueous solutions. Nature, 406, 611–614, https://doi.org/10.1038/35020537.
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.
Krämer, M., and Coauthors, 2009: Ice supersaturations and cirrus cloud crystal numbers. Atmos. Chem. Phys., 9, 3505–3522, https://doi.org/10.5194/acp-9-3505-2009.
Krämer, M., and Coauthors, 2016: A microphysics guide to cirrus clouds—Part 1: Cirrus types. Atmos. Chem. Phys., 16, 3463–3483, https://doi.org/10.5194/acp-16-3463-2016.
Krämer, M., and Coauthors, 2020: A microphysics guide to cirrus—Part 2: Climatologies of clouds and humidity from observations. Atmos. Chem. Phys., 20, 12 569–12 608, https://doi.org/10.5194/acp-20-12569-2020.
Podglajen, A., A. Hertzog, R. Plougonven, and B. Legras, 2016: Lagrangian temperature and vertical velocity fluctuations due to gravity waves in the lower stratosphere. Geophys. Res. Lett., 43, 3543–3553, https://doi.org/10.1002/2016GL068148.
Podglajen, A., R. Plougonven, A. Hertzog, and E. Jensen, 2018: Impact of gravity waves on the motion and distribution of atmospheric ice particles. Atmos. Chem. Phys., 18, 10 799–10 823, https://doi.org/10.5194/acp-18-10799-2018.
Pruppacher, H. R., and J. D. Klett, 2010: Microphysics of Clouds and Precipitation. Springer, 954 pp., https://doi.org/10.1007/978-0-306-48100-0.
Ren, C., and A. R. Mackenzie, 2005: Cirrus parametrization and the role of ice nuclei. Quart. J. Roy. Meteor. Soc., 131, 1585–1605, https://doi.org/10.1256/qj.04.126.
Spichtinger, P., and K. M. Gierens, 2009: Modelling of cirrus clouds—Part 1a: Model description and validation. Atmos. Chem. Phys., 9, 685–706, https://doi.org/10.5194/acp-9-685-2009.
Spichtinger, P., and D. J. Cziczo, 2010: Impact of heterogeneous ice nuclei on homogeneous freezing events. J. Geophys. Res., 115, D14208, https://doi.org/10.1029/2009JD012168.
Spichtinger, P., and M. Krämer, 2013: Tropical tropopause ice clouds: A dynamic approach to the mystery of low crystal numbers. Atmos. Chem. Phys., 13, 9801–9818, https://doi.org/10.5194/acp-13-9801-2013.
Spichtinger, P., P. Marschalik, and M. Baumgartner, 2023: Impact of formulations of the homogeneous nucleation rate on ice nucleation events in cirrus. Atmos. Chem. Phys., 23, 2035–2060, https://doi.org/10.5194/acp-23-2035-2023.
Spreitzer, E. J., M. P. Marschalik, and P. Spichtinger, 2017: Subvisible cirrus clouds—A dynamical system approach. Nonlinear Processes Geophys., 24, 307–328, https://doi.org/10.5194/npg-24-307-2017.
Wang, M., and J. E. Penner, 2010: Cirrus clouds in a global climate model with a statistical cirrus cloud scheme. Atmos. Chem. Phys., 10, 5449–5474, https://doi.org/10.5194/acp-10-5449-2010.
Zhang, Y., A. Macke, and F. Albers, 1999: Effect of crystal size spectrum and crystal shape on stratiform cirrus radiative forcing. Atmos. Res., 52, 59–75, https://doi.org/10.1016/S0169-8095(99)00026-5.
Zhou, C., J. E. Penner, G. Lin, X. Liu, and M. Wang, 2016: What controls the low ice number concentration in the upper troposphere? Atmos. Chem. Phys., 16, 12 411–12 424, https://doi.org/10.5194/acp-16-12411-2016.