1. Introduction
Stratocumulus (Sc) clouds are one of the most common cloud types on Earth (Hahn and Warren 2007). They form under strong temperature inversions and are prevalent off the western coast of continents, on the descending side of the Hadley cell. Their impact on Earth’s energy budget is significant as they strongly reflect incoming solar radiation, with a much weaker effect on outgoing longwave radiation (Wood 2012). Accurate modeling of Sc clouds has high importance for several reasons: (i) they are one of the key sources of uncertainty in climate predictions (Bony and Dufresne 2005; Zelinka et al. 2017), (ii) they affect solar power integration into the electric grid (Yang and Kleissl 2016; Zhong et al. 2017; Wu et al. 2018), and (iii) they impact aviation by hindering the takeoff and landing of flights (Reynolds et al. 2012).
Physical processes governing the evolution of the stratocumulus-topped boundary layer (STBL)—such as cloud-top radiative cooling, entrainment, evaporative cooling, surface fluxes, wind shear, and precipitation—widely range on spatial and temporal scales, and modeling Sc clouds is quite challenging as a result (e.g., Lilly 1968; Stevens 2002; Wood 2012). Efforts through both observational campaigns (e.g., Stevens et al. 2003; Malinowski et al. 2013; Crosbie et al. 2016) and high-resolution numerical modeling (e.g., Stevens et al. 2005; Kurowski et al. 2009; Yamaguchi and Randall 2012; Chung et al. 2012; Blossey et al. 2013; de Lozar and Mellado 2015; Pedersen et al. 2016; Mellado et al. 2018; Matheou and Teixeira 2019) have significantly advanced our understanding of the physics of Sc clouds. These physical insights are important for numerical weather prediction (NWP) and general circulation models (GCMs) where grid resolution is coarse.
The picture emerging from those studies is that cloud-top radiative cooling is a critical source of STBL turbulence (Matheou and Teixeira 2019), contributing to cloud-top entrainment (Mellado 2017). The combined effect of both evaporative and radiative cooling—the former typically enhanced by wind shear (Mellado et al. 2014)—destabilizes the top of cloud layer through buoyancy reversal that leads to the formation of negatively buoyant weak downdrafts. This process is often considered responsible for the generation of cloud holes in largely unbroken Sc clouds (Gerber et al. 2005; Kurowski et al. 2009). Many small-scale phenomena (e.g., entrainment, shear, evaporative cooling, cloud microphysics) are at play in the origin of downdrafts and can strongly influence vertical mixing (Mellado 2017). Exactly how these processes interact with each other remains a research challenge.
Turbulent transport in the STBL is the main driver to the formation, maintenance, and dissipation of Sc clouds. In coarse-resolution models, turbulent transport is typically parameterized using simplified one-dimensional planetary boundary layer (PBL) schemes. Global NWP models (e.g., Teixeira 1999) and climate models tend to underestimate Sc clouds (Teixeira et al. 2011; Lin et al. 2014), although there is an improvement in the representation of the radiative properties by a newer generation of climate models (Engström et al. 2014). In terms of mesoscale models, Ghonima et al. (2017) compared three different PBL schemes in the Weather Research and Forecasting (WRF) Model and found that they all underestimate entrainment, producing too moist and cold STBLs. Huang et al. (2013) compared five different WRF PBL parameterizations and highlighted the difficulties of simulating the STBL. Recent studies supported the importance of downdrafts in transporting turbulent heat and moisture flux in the PBL (Chinita et al. 2018; Davini et al. 2017; Brient et al. 2019) through analyzing LES of STBL. Brient et al. (2019) concluded that for a more accurate parameterization of turbulence within STBL, downdrafts should be explicitly included in climate models. Downdrafts were recently implemented by Han and Bretherton (2019) in a turbulent kinetic energy (TKE)-based moist eddy-diffusivity/mass-flux (EDMF) parameterization within the GFS model, and they found more accurate liquid water and wind speed profiles for marine STBLs.
This study introduces parameterized downdrafts into NWP and aims at investigating their impact on the evolution of the STBL. To test whether convective downdrafts are necessary to properly represent Sc clouds, we implement a new downdraft parameterization in WRF based on the EDMF approach that uses Mellor–Yamada–Nakanishi–Niino (MYNN) as the ED component. This differs from Han and Bretherton (2019) where different ED and MF models were used and additional features were implemented to advance the vertical turbulence mixing parameterization for not only STBL but also other conditions. We place a special emphasis on evaluating the role of nonlocal transport in STBL with gradual changes to the model in order to separate the effects coming from convective downdrafts. The new parameterization is evaluated in two typical STBL cases, frequently used in modeling studies.
Section 2 describes the EDMF and MYNN schemes as well as the updraft and downdraft implementation in WRF. The numerical design of the LES setup, WRF single-column model (SCM), and updraft and downdraft properties are described in section 3. WRF SCM results for both STBL cases are shown in section 4. Finally, conclusions are presented in section 5.
2. PBL scheme with downdrafts
In coarse-resolution atmospheric models, the PBL scheme determines turbulent flux profiles within the PBL as well as the overlying air, providing tendencies of temperature, moisture, and horizontal momentum due to mixing and turbulent transport for the entire atmospheric column. This section first gives an overview of the EDMF framework, then the details of ED and MF models are presented (sections 2b and 2c). The properties of updrafts and downdrafts are diagnosed using LES and presented in section 3 in order to quantify the validity of the parameterized mass-flux model.
a. The eddy-diffusivity/mass-flux (EDMF) approach
Siebesma and Teixeira (2000), Teixeira and Siebesma (2000), and Siebesma et al. (2007) introduced the eddy diffusivity/mass-flux (EDMF) approach for parameterizing turbulence in the dry convective boundary layer, and additional improvements have been made by Witek et al. (2011). The idea behind EDMF is to parameterize the turbulent fluxes as a sum of local (diffusive) transport through ED and nonlocal (convective) transport through the mass-flux contribution. The EDMF approach has been extended to represent moist convection since then (e.g., Soares et al. 2004; Neggers et al. 2009; Neggers 2009; Angevine et al. 2010, 2018; Sušelj et al. 2013; Suselj et al. 2019a,b). For the moist extention, the updrafts start out as dry and condense when the saturation conditions are met. In other words, dry updrafts can continuously change into moist updrafts, without assuming any coupling between those two layers.
b. ED scheme: The Mellor–Yamada–Nakanishi–Niino (MYNN)
c. Adding mass flux to MYNN
The MYNN Level 2.5 ED model determines turbulent mixing at each vertical level based on the gradients in scalars between immediately adjacent vertical levels [Eq. (1)]. When deep mixing due to larger eddies becomes important, the MYNN scheme has been shown to produce erroneous thermodynamic profiles (Huang et al. 2013). Nonlocal models, such as the YSU and ACM2 schemes, account for this deep mixing by using a countergradient term (Hong et al. 2006) or a transilient mass-flux matrix (Pleim 2007). Another common approach is the EDMF framework, which decomposes the subgrid vertical mixing into local mixing through ED and nonlocal [mass flux (MF)] transport through convective plumes. Traditionally, PBL schemes, such as MYNN, model the turbulence within the PBL through only turbulent diffusion. In the EDMF framework, ED is used to model the turbulent transport in the nonconvective environment, with an additional contribution from the convective mass flux.
1) Mass-flux model overview
2) Surface-driven updrafts
A version of EDMF including surface-driven updrafts (Olson et al. 2019) has been implemented as an add-on option in MYNN since WRF v3.8 and is used for NOAA’s operational Rapid Refresh (RAP; Benjamin et al. 2016) and High Resolution Rapid Refresh (HRRR) forecast systems. The original version of this dynamic multiplume mass-flux scheme in WRF v3.8 (bl_mynn_edmf = 1) followed Sušelj et al. (2013), but the version in the current WRF v4.0 contains considerable changes from the original form. We do not base our EDMF implementation (bl_mynn_edmf = 3) on what is currently available in WRF, but instead follow Sušelj et al. (2013) and Suselj et al. (2019a,b). The numerical implementation is documented in Suselj et al. (2019b) (their appendix B).
While Sušelj et al. (2013) did not include either the dynamical pressure effect [i.e.,
Since each updraft is characterized by different surface conditions and entrainment rates, the thermodynamic properties and termination heights also differ. Each plume is integrated independently in the vertical until the vertical velocity becomes negative. As mentioned above, we assume Gaussian distribution of the subgrid-scale θl and qt, which determines the grid-mean liquid water and cloud cover. We then compute the condensation in the convective plumes assuming an uniform distribution of θl and qt within each of the plumes. Therefore, the condensation within a plume occurs when qt exceeds the saturated mixing ratio (qs). The precise calculation of condensation within each plume is necessary because of its impact on plume buoyancy. Finally, there exist dry and partly moist plumes among the N updrafts, and the fate of each plume is determined by its initial conditions, dynamical pressure effect, and lateral entrainment. Since each individual updraft is integrated independently, whenever the vertical velocity becomes negative and the updraft terminates, the total updraft area is reduced. This is common in regions with strong lateral entrainment rates.
3) Cloud-top-triggered downdrafts
Several important physical processes are at play near the STBL top. Radiative and evaporative cooling produces cooled downdrafts and drives buoyant production of turbulence in the PBL. Entrainment from the free troposphere can impact downdrafts near the cloud top: warm air from the free troposphere counteracts the radiative cooling and buoyant production of turbulence. When the PBL is less turbulent, the entrainment rate decreases, indicating a negative feedback loop (Wood 2012). Surface-driven updrafts may also affect the downdrafts. As updrafts approach the inversion, they begin to diverge and can help initiate or enhance downdrafts (Kurowski et al. 2009; Davini et al. 2017). This enhances the downdraft vertical velocity and, in turn, the turbulence in the PBL. In the proposed 1D parameterization of downdrafts, those dependencies are important for the formulation of the downdraft initial conditions. Our downdraft parameterization in MYNN can be activated by specifying bl_mynn_edmf_dd = 1 in the namelist. The numerical implementation follows Suselj et al. (2019b) (see their appendix C).
Similar to the updrafts, equations for each downdraft are independently integrated in the vertical until the downdraft velocity vanishes. Condensation occurs within a downdraft if its total water mixing ratio exceeds the saturated water mixing ratio. Similarly to updrafts, there can exist dry and partly moist plumes among the M downdrafts, and the fate of each plume is determined by its initial conditions, dynamical pressure effect near the surface, and lateral entrainment. We assume that both updrafts and downdrafts interact with the mean field only and not with other updrafts or downdrafts. This assumption has been tested in Kurowski et al. (2019) for updrafts in shallow convection, and further investigation for downdrafts should be explored. Since each individual downdraft is integrated independently, whenever vertical velocity becomes zero/positive and the downdraft terminates, the total downdraft area is reduced.
3. Design of numerical experiments
a. LES setup
Large eddy simulations are performed using the UCLA-LES model (Stevens 2010) and treated as “ground truth.” Two idealized nondrizzling marine Sc cases are chosen as baseline simulations: the DYCOMS-II RF01 (Stevens et al. 2005) and CGILS S12 Control (Blossey et al. 2013) (hereafter DYCOMS and CGILS). The experiments are set up following the respective intercomparison studies. Interactive radiation is treated differently in the two cases. Specifically, a simplified model of radiative forcing matching the δ-four stream transfer code (Stevens et al. 2005) is used in DYCOMS. As for CGILS, a full radiative transfer code is used, which utilizes Monte Carlo sampling of the spectral integration (Pincus and Stevens 2009). The DYCOMS case is run for 4 h, and the CGILS case is run for 24 h. While we focus our analysis of the updraft and downdraft properties on nocturnal quasi-steady conditions (first 4 h), the 24-h simulation of CGILS provides reference to the generalization of the parameterization during the day. In both experiments, a nonuniform vertically stretched grid is used with 5-m resolution around the inversion, and a several times coarser resolution in the horizontal. This LES setup is identical to that in Ghonima et al. (2017). A summary of the model setups is provided in Table 1.
Summary of large eddy simulation setups in UCLA-LES, including uniform horizontal grid spacing Δx, y, vertical grid spacing at the inversion Δzinv (m), horizontal domain size Lx,y, and divergence of large-scale winds D.
Determining plume properties
Simulation outputs are stored at 1 min intervals from hours 3 to 4 in order to diagnose updraft and downdraft properties. The statistics are averaged over 1 h. We use the joint normal probability density function (PDF) between vertical velocity w, total water mixing ratio (qt = qυ + ql), virtual potential temperature [θυ = θ(1 + 0.61qυ − ql)], and liquid water potential temperature [θl = θ − (Lυql)(cpπ)−1] to define LES updrafts and downdrafts. Lυ is the latent heat of vaporization, cpd is the specific heat of dry air at constant pressure, π is the Exner function, and subscripts are υ for vapor, l for liquid. We define the normalized variable to be
The mean downdraft and updraft properties are shown in Fig. 2 for DYCOMS and Fig. 3 for CGILS. Updraft and downdraft areas are comparable in the middle of the PBL (Figs. 2a and 3a), with updrafts decreasing near cloud top and downdrafts decreasing before reaching the surface. Figures 2b,c and 3b,c show partial contributions to the total heat and moisture fluxes from the environment, updrafts, and downdrafts. Similar results are found in both STBL cases: cloud-top entrainment heat flux is largely from updrafts; the peak in downdraft heat and moisture transport is slightly below the peak in updrafts (≈100 m lower); heat and moisture transport from downdrafts is stronger than updrafts in cloudy region; environmental mean of w, θl, θυ, qt, and ql is very close to the grid mean. Both cases have similar updraft and downdraft properties: downdrafts terminate before reaching the surface (Figs. 2a and 3a); updraft and downdraft vertical velocity are approximately a mirror image of each other (Figs. 2d and 3d); downdrafts become negatively buoyant (
The properties shown in these two STBL cases compare well to the case in Brient et al. (2019), where the First ISCCP Regional Experiment (FIRE) study was simulated for 24 h to study the diurnal cycle of coherent updraft and downdraft properties. Specifically, the nighttime results of Brient et al. (2019) show that the areas of updrafts and downdrafts are comparable in the middle of the PBL (around 12%) and the downdraft area decreases quickly to zero below 100 m, which corresponds well with our findings for DYCOMS. CGILS results show a slightly smaller downdraft area in the middle of the PBL (around 9%). The turbulent heat flux in Brient et al. (2019) shows that the transport of heat by updrafts is the strongest at cloud top, the peak of the downdraft heat transport is slightly below that for the updrafts (≈50 m lower), and the heat transport by updrafts in cloudy region is nearly zero when downdrafts dominate. This corresponds well with DYCOMS, while updrafts in CGILS have a slightly positive heat transport in the cloudy region. As for the turbulent moisture flux, Brient et al. (2019) shows that updrafts dominate from the surface up to slightly above cloud base, while downdrafts dominate in the cloud layer. Moisture flux is similar in DYCOMS and CGILS, but our results show a positive peak of updraft moisture flux near cloud top, making the updraft contribution to the moisture flux a dominating term around cloud top. Chinita et al. (2018) shows large differences in the contribution of updrafts and downdrafts to total flux for DYCOMS in the cloud layer. In general, they find that updrafts account for most of the organized motions near the surface, while downdrafts are more important near the boundary layer top. While the overall properties are similar, updraft and downdraft areas in Chinita et al. (2018) are 5%–10% larger.
b. WRF single-column model
DYCOMS and CGILS case are simulated using the Weather Research and Forecasting (WRF) v4.0 single-column model (SCM) and compared against LES. Initial conditions and forcing are identical to that in LES (i.e., fixed surface fluxes for DYCOMS and CGILS, large-scale subsidence as in Table 1) and was used previously in Ghonima et al. (2017). The SCM vertical domain includes 116 levels to resolve the lowest 12 km of the troposphere, which comes out to be Δz ≈ 20 m in the first 1 km. A simulation time step of 40 s is used. In section 4c, we show that results are insensitive to time step between 5 and 60 s. Three different versions of one PBL scheme are used to determine the importance of the introduced changes: 1) the original Mellor–Yamada–Nakanishi–Niino scheme (MYNN; hereafter ED) (Nakanishi and Niino 2006, 2009), 2) MYNN with updrafts (EDMFU), and 3) MYNN with updrafts and downdrafts (EDMFUD). For EDMFU and EDMFUD, the MYNN scheme is used as a parameterization of local transport in the nonconvective environment. The radiation scheme is RRTMG (Iacono et al. 2008). No microphysics or cumulus schemes are used since both cases represent nonprecipitating STBL.
4. Results
a. DYCOMS-II RF01
Figure 4 shows the mean fields of θl, qt, ql, cloud fraction, u, υ, heat flux (
Figures 6 and 7 show the vertical flux contribution from the individual components: environment (ED), updraft, and downdraft. Figure 6 is for EDMFU, which includes only ED and updraft. Note that LES transport in 6A and D includes LES environmental and downdraft transport because in the case of updrafts only, the remaining area is considered to be the environment and should therefore be modeled by ED. Updraft contribution to the heat flux matches the profile in LES well, however it is overestimated in most of PBL and the cloud-top entrainment heat flux is too strong. It is important to note that cloud-top entrainment is not fully understood even in LES. We find here that even though entrainment heat flux appears to be strong, boundary layer averaged temperature in EDMFU is still too cold compared to LES (Fig. 5b). However, EDMFU produces a warmer boundary layer compared to ED, which strongly underestimates entrainment heat flux. Updraft contribution to the moisture flux is overestimated throughout the PBL, but ED component is underestimated and the total moisture flux matches LES well. The initial updraft starting θl and qt are stronger than LES (not shown) and eventually leads to overestimation of moisture flux. This indicates that the formulation of updraft surface condition in STBL may be different from shallow convection since we retain the same updraft starting condition used in Suselj et al. (2019a). In shallow convection, surface fluxes are the main driver for updraft surface conditions. Whether other physical processes are at play in the parameterization of updraft surface conditions in STBL should be investigated in the future. We find that in the current configuration, ED compensates for the overestimation of updraft moisture flux, resulting in a good match with LES in the total moisture flux.
Based on 800 additional simulations, exploring the parameter space, with different lateral entrainment rates and dynamical effects (varying L0 and cent in Eq. (10) from and 10 to 100 m 0.5 to 5 m−1, as well as varying z00 in Eq. (8) from 50 to 200 m; not shown), we observe that the most important impact of the updraft is the transport near cloud top because ED models an insufficient heat and moisture transport in this location, causing a cold and moist bias. Additionally, ED does not accurately represent a well-mixed layer, while EDMFU has a better well-mixed profile in both θl and qt. The final configuration was chosen to have the best match in the mean field of θl, qt, and total heat and moisture transport with LES.
For EDMFUD, Fig. 7 shows partial contributions to the total transport from ED, updrafts, and downdrafts. Comparing Figs. 6 and 7, we argue that the downdraft transport is implicitly included in the ED contribution in EDMFU (Figs. 6a,d) as the sum of heat and moisture transport for EDMFU versus EDMFUD is similar. Averaged plume properties from EDMFUD are shown in Fig. 8. For downdraft contribution to total fluxes, EDMFUD underestimates the strength in heat and moisture flux. More spefically, downdraft heat transport decreases too quickly before reaching the surface (Fig. 7c). For moisture tansport, downdraft qt also decrease quickly, and the starting downdraft qt is underestimated (Fig. 8c). Updraft contribution to heat transport (Fig. 7b) is similar to that in EDMFU, and they both slightly overestimate compared to LES in terms. This can be seen in the overestimation of updraft area and vertical velocity (Figs. 8a,b), and is a result of the positive bias in updraft starting surface conditions, espicially updraft starting vertical velocity. For updraft moisture transport, updrafts in EDMFUD do not overestimate as strongly as EDMFU. This is likely due to downdrafts transporting dry and warm air in the PBL and causing updrafts to mix differently. On top of that, the mean fields of θl and qt are different in EDMFU and EDMFUD. Note that since the definition of updrafts and downdrafts in LES is somewhat arbitrary, the total transport should be the main indicator of success for a parameterization. Nevertheless, the definition of updrafts and downdrafts as in section 3 is a reference point for bench-marking updraft and downdraft parameterizations. Overall, general agreement of plume properties are found between the SCMs and LES. For DYCOMS, downdraft transport decreases too quickly for both heat and moisture. We find that modeling downdraft transport in the upper part of the boundary layer correctly is more important than retaining downdraft throughout the PBL. The mean fields respond more to changes in turbulent transport in the upper part of the PBL. Indeed, the qt profile is most well mixed in EDMFUD, signaling the importance of downdraft moisture transport. This is consistent with the hypothesis in Sušelj et al. (2013), suggesting that the inclusion of downdrafts could increase vertical mixing in the upper part of the boundary layer. In STBL, mixing from the surface provides moisture and entrainment from the free troposphere dries the boundary layer. However, in the heat profile, both the surface and entrainment from the free troposphere heats the boundary layer. We find here that downdrafts help provide stronger moisture mixing near cloud top and keep the bias in total moisture low. In addition, EDMFUD has the least bias in boundary layer averged θl, as downdrafts also contribute to transporting warm air in the PBL.
Downdraft model coefficients and final lateral entrainment configuration are chosen to have the best match against LES in the mean field of θl, qt, u, and υ. EDMFU and EDMFUD have the same updraft lateral entrainment configuration.
Comparing EDMFU with SCM results from Sušelj et al. (2013), a resemblance of the updraft transport of heat and moisture is found. The formulations of updrafts are identical except for the added entrainment and dynamical pressure effect near cloud top in EDMFU. It is no surprise that some differences are seen, given the different assumptions made in ED. Specifically, the vertical transport in the middle of the boundary layer is different in the two models. While EDMFU shows positive transport from updraft in the cloudy region for heat, the updraft model in Sušelj et al. (2013) shows a negative heat transport. For moisture, EDMFU produces stronger transport. This is likely due to the added entrainment dynamic effect in our updraft model, different subgrid cloud assumption, and different ED model for the nonconvective environment. In the end, the total heat and moisture transport is similar between the two models as ED compensates for the difference, and they both match LES well.
Comparing EDMFUD with SCM results from Han and Bretherton (2019), we found contrary conclusions for the effect of the downdraft parameterization. While Han and Bretherton (2019) found a slight overprediction for θl and overmixing for qt in their DYCOMS experiment, we found slight underprediction for θl and undermixing for qt.
b. CGILS S12 control
Figure 9 shows the mean fields of θl, qt, ql, cloud fraction, u, υ, heat flux (
During hours 3 to 4, EDMFU and EDMFUD show small bias in heat and moisture profile, whereas ED is too cold and too moist. This causes the overestimation of LWP in ED. The cloud-top height in EDMFUD is one grid point above ED, likely due to the stronger entrainment flux near cloud top from mass flux. EDMFUD overestimates u and underestimates υ in the PBL. ED shows similar results as DYCOMS, where the horizontal wind does not have a strong transition between the PBL and the free troposphere. EDMFU shows a very good match in total heat and moisture transport, while EDMFUD has a slightly stronger moisture transport near cloud top. Similar to DYCOMS, ED does not capture cloud-top entrainment flux. Figures 11 and 12 show the vertical flux contribution from each individual component: environment (ED), updrafts, and downdrafts. In both EDMFU and EDMFUD, updraft heat and moisture transport are overestimated. However, in the presence of downdrafts, updraft moisture transport decreases more strongly in-cloud. Downdrafts in EDMFUD partially compensate for these changes, resulting in a similar total transport. Averaged plume properties from EDMFUD are shown in Fig. 13. In CGILS, a good agreement of plume properties is found between the SCMs and LES. Again, we find that simulation results are more sensitive to the modeling of downdraft transport in the upper part of the PBL. In the end, we select model parameters that result in realistic mean profiles of θl, qt, u, and υ for both DYCOMS and CGILS. While the parameterized downdrafts terminate too quickly in DYCOMS, we find that they mostly reach the surface in CGILS.
In the present study, we develop our updraft and downdraft parameterization using their nocturnal properties. The 24-h simulation of CGILS suggests that updrafts and downdrafts may play different roles during the daytime. This is also observed in the study done by Brient et al. (2019). Parameterization of updrafts and downdrafts during the day should be investigated in the future.
c. Simulation time step and run time
To test the numerical stability of the scheme and the convergence of the results, we also run simulations with different time steps: 5, 10, 20, 30, and 40, and 60 s as shown in Fig. 14. Note that the figures shown in this study use a time step of 40 s. The obtained results confirm that both EDMFU and EDMFUD are not sensitive to the imposed time step changes, proving high robustness of the scheme. The LWP, and the boundary layer-averaged heat and moisture amounts all converge to the same values at the end of the simulation. Additionally, we record simulation run times normalized by the ED-simulation run time for different time steps (Table 2). On average, including the updrafts slows the simulation down by approximately 5%, while including both updrafts and downdrafts slows it down by 7%. This indicates that EDMF is a numerically inexpensive scheme.
EDMFU and EDMFUD run time normalized by ED using different time steps.
5. Summary and conclusions
In this study, we investigated the role of nonlocal transport on the development and maintenance of the STBL in coarse-resolution atmospheric models. A special emphasis has been put on the evaluation of downdraft contribution, recently suggested as an important missing element of convection/turbulence parameterizations (Chinita et al. 2018; Davini et al. 2017; Brient et al. 2019). A new parameterization of cloud-top-triggered downdrafts has been proposed along with a complementary parameterization of surface-driven updrafts. The parameterization was validated against large-eddy simulations of two marine stratocumulus cases: DYCOMS and CGILS. The applied nonlocal mass-flux scheme is part of the stochastic multiplume EDMF approach decomposing the turbulent transport into the local and nonlocal contributions. The local transport in the boundary layer is represented by the MYNN scheme. The EDMF scheme has been implemented and tested in the WRF single-column modeling framework.
In the new parameterization, the thermodynamic and dynamic properties of downdrafts are controlled by stochastic lateral entrainment affecting their dilution along the vertical development. The number of downdrafts is fixed to 10, and all downdrafts are assumed to start randomly in the upper half of cloud layer, with the total starting area of approximately 15%, similarly to updrafts. The strength of the downdraft vertical velocity is formulated as a combined effect of the intensity of the surface-driven updrafts and cloud-top radiative cooling. The starting downdraft thermodynamic properties are proportional to the entrainment flux at the STBL top, which is determined by the jump values of heat or moisture across the inversion.
To evaluate the importance of the updraft and downdraft contributions, we run three different SCM simulations for the tested STBLs: without mass flux (ED), with updrafts only (EDMFU), and with both updrafts and downdrafts (EDMFUD). When there is no mass-flux (neither updraft nor downdraft), ED underestimates the cloud-top entrainment flux, yielding a cold and moist bias that leads to a strong overestimation of LWP. The inclusion of updrafts increases the cloud-top entrainment flux and keeps the mean STBL profiles more well-mixed and reduces the temperature and moisture biases. We find that including downdrafts increases vertical mixing in the upper part of the boundary layer especially for qt, and it results in a warmer and drier STBL than for EDMFU. Overall, the proposed parameterization reproduces the LES profiles because of the addition of downdraft heat and moisture transport in the WRF SCM. However, we find that differences in EDMFU and EDMFUD are not significant.
Based on the results from the two STBL cases, we conclude that, for the tested version of the WRF Model, it is necessary to include updrafts as part of the nonlocal mass-flux as ED does not represent correctly the cloud-top entrainment flux. An addition of downdrafts shows some improvements in these two cases. However, further investigations are needed to determine whether downdrafts play a greater role in different meteorological conditions. We hypothesize that ED would have a better match with LES when there is less cloud-top entrainment (e.g., when the PBL is less turbulent), and that the inclusion of downdrafts would be necessary when surface fluxes are small. A recent study by Matheou and Teixeira (2019) compared various LESs of STBL for different physical processes included and concluded that surface fluxes, surface shear, and cloud-top radiative cooling all contribute substantially to the turbulence in STBL. Whether the EDMF parameterization responds similarly in such conditions will be investigated in the future.
Acknowledgments
Parts of this research were carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration, and were supported by the U.S. Department of Energy, Office of Biological and Environmental Research, Earth System Modeling; and the NASA MAP Program. E. Wu acknowledges the support provided by the JPL Education Office. We thank Mónica Zamora Zapata and Thijs Heus for constructive comments. We also thank Minghua Ong for proofreading the manuscript. The WRF Model is freely available at https://github.com/wrf-model/WRF, the modifications made in this paper can be found at https://github.com/elynnwu/EDMF_JPL.
REFERENCES
Angevine, W. M., H. Jiang, and T. Mauritsen, 2010: Performance of an eddy diffusivity–mass flux scheme for shallow cumulus boundary layers. Mon. Wea. Rev., 138, 2895–2912, https://doi.org/10.1175/2010MWR3142.1.
Angevine, W. M., J. Olson, J. Kenyon, W. I. Gustafson, S. Endo, K. Suselj, and D. D. Turner, 2018: Shallow cumulus in WRF parameterizations evaluated against LASSO large-eddy simulations. Mon. Wea. Rev., 146, 4303–4322, https://doi.org/10.1175/MWR-D-18-0115.1.
Benjamin, S. G., and Coauthors, 2016: A North American hourly assimilation and model forecast cycle: The Rapid Refresh. Mon. Wea. Rev., 144, 1669–1694, https://doi.org/10.1175/MWR-D-15-0242.1.
Blossey, P. N., and Coauthors, 2013: Marine low cloud sensitivity to an idealized climate change: The CGILS LES intercomparison. J. Adv. Model. Earth Syst., 5, 234–258, https://doi.org/10.1002/jame.20025.
Bony, S., and J.-L. Dufresne, 2005: Marine boundary layer clouds at the heart of tropical cloud feedback uncertainties in climate models. Geophys. Res. Lett., 32, L20806, https://doi.org/10.1029/2005GL023851.
Brient, F., F. Couvreux, N. Villefranque, C. Rio, and R. Honnert, 2019: Object-oriented identification of coherent structures in large eddy simulations: Importance of downdrafts in stratocumulus. Geophys. Res. Lett., 46, 2854–2864, https://doi.org/10.1029/2018GL081499.
Chinita, M. J., G. Matheou, and J. Teixeira, 2018: A joint probability density–based decomposition of turbulence in the atmospheric boundary layer. Mon. Wea. Rev., 146, 503–523, https://doi.org/10.1175/MWR-D-17-0166.1.
Chung, D., G. Matheou, and J. Teixeira, 2012: Steady-state large-eddy simulations to study the stratocumulus to shallow cumulus cloud transition. J. Atmos. Sci., 69, 3264–3276, https://doi.org/10.1175/JAS-D-11-0256.1.
Crosbie, E., and Coauthors, 2016: Stratocumulus cloud clearings and notable thermodynamic and aerosol contrasts across the clear–cloudy interface. J. Atmos. Sci., 73, 1083–1099, https://doi.org/10.1175/JAS-D-15-0137.1.
Davini, P., F. D’Andrea, S.-B. Park, and P. Gentine, 2017: Coherent structures in large-eddy simulations of a nonprecipitating stratocumulus-topped boundary layer. J. Atmos. Sci., 74, 4117–4137, https://doi.org/10.1175/JAS-D-17-0050.1.
de Lozar, A., and J. P. Mellado, 2015: Mixing driven by radiative and evaporative cooling at the stratocumulus top. J. Atmos. Sci., 72, 4681–4700, https://doi.org/10.1175/JAS-D-15-0087.1.
de Roode, S. R., A. P. Siebesma, H. J. J. Jonker, and Y. de Voogd, 2012: Parameterization of the vertical velocity equation for shallow cumulus clouds. Mon. Wea. Rev., 140, 2424–2436, https://doi.org/10.1175/MWR-D-11-00277.1.
Engström, A., F. A.-M. Bender, and J. Karlsson, 2014: Improved representation of marine stratocumulus cloud shortwave radiative properties in the CMIP5 climate models. J. Climate, 27, 6175–6188, https://doi.org/10.1175/JCLI-D-13-00755.1.
Gerber, H., G. Frick, S. P. Malinowski, J.-L. Brenguier, and F. Burnet, 2005: Holes and entrainment in stratocumulus. J. Atmos. Sci., 62, 443–459, https://doi.org/10.1175/JAS-3399.1.
Ghonima, M. S., H. Yang, C. K. Kim, T. Heus, and J. Kleissl, 2017: Evaluation of WRF SCM simulations of stratocumulus-topped marine and coastal boundary layers and improvements to turbulence and entrainment parameterizations. J. Adv. Model. Earth Syst., 9, 2635–2653, https://doi.org/10.1002/2017MS001092.
Hahn, C., and S. Warren, 2007: A gridded climatology of clouds over land (1971–1996) and ocean (1954–1997) from surface observations worldwide. Numeric Data Package NDP-026E ORNL/CDIAC-153, CDIAC, Department of Energy, Oak Ridge, TN, 71 pp., https://doi.org/10.3334/CDIAC/CLI.NDP026E.
Han, J., and C. S. Bretherton, 2019: TKE-based moist Eddy-Diffusivity Mass-Flux (EDMF) parameterization for vertical turbulent mixing. Wea. Forecasting, 34, 869–886, https://doi.org/10.1175/WAF-D-18-0146.1.
Hong, S.-Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an explicit treatment of entrainment processes. Mon. Wea. Rev., 134, 2318–2341, https://doi.org/10.1175/MWR3199.1.
Huang, H.-Y., A. Hall, and J. Teixeira, 2013: Evaluation of the WRF PBL parameterizations for marine boundary layer clouds: Cumulus and stratocumulus. Mon. Wea. Rev., 141, 2265–2271, https://doi.org/10.1175/MWR-D-12-00292.1.
Iacono, M. J., J. S. Delamere, E. J. Mlawer, M. W. Shephard, S. A. Clough, and W. D. Collins, 2008: Radiative forcing by long-lived greenhouse gases: Calculations with the AER radiative transfer models. J. Geophys. Res., 113, D13103, https://doi.org/10.1029/2008JD009944.
Köhler, M., 2006: An integral approach to modeling PBL transports and clouds: EDMF@ECMWF. Proc. Workshop on Parametrization of Clouds in Large-Scale Models, Reading, United Kingdom, ECMWF, 93–101.
Kurowski, M. J., S. P. Malinowski, and W. W. Grabowski, 2009: A numerical investigation of entrainment and transport within a stratocumulus-topped boundary layer. Quart. J. Roy. Meteor. Soc., 135, 77–92, https://doi.org/10.1002/qj.354.
Kurowski, M. J., K. Suselj, and W. W. Grabowski, 2019: Is shallow convection sensitive to environmental heterogeneities? Geophys. Res. Lett., 46, 1785–1793, https://doi.org/10.1029/2018GL080847.
Kuwano-Yoshida, A., T. Enomoto, and W. Ohfuchi, 2010: An improved PDF cloud scheme for climate simulations. Quart. J. Roy. Meteor. Soc., 136, 1583–1597, https://doi.org/10.1002/qj.660.
Lilly, D. K., 1968: Models of cloud-topped mixed layers under a strong inversion. Quart. J. Roy. Meteor. Soc., 94, 292–309, https://doi.org/10.1002/qj.49709440106.
Lin, J.-L., T. Qian, and T. Shinoda, 2014: Stratocumulus clouds in southeastern Pacific simulated by eight CMIP5–CFMIP global climate models. J. Climate, 27, 3000–3022, https://doi.org/10.1175/JCLI-D-13-00376.1.
Lock, A. P., and M. K. Macvean, 1999: The parametrization of entrainment driven by surface heating and cloud-top cooling. Quart. J. Roy. Meteor. Soc., 125, 271–299, https://doi.org/10.1002/qj.49712555315.
Malinowski, S. P., and Coauthors, 2013: Physics of Stratocumulus Top (POST): Turbulent mixing across capping inversion. Atmos. Chem. Phys., 13, 12 171–12 186, https://doi.org/10.5194/acp-13-12171-2013.
Matheou, G., and J. Teixeira, 2019: Sensitivity to physical and numerical aspects of large-eddy simulation of stratocumulus. Mon. Wea. Rev., 147, 2621–2639, https://doi.org/10.1175/MWR-D-18-0294.1.
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.
Mellado, J. P., B. Stevens, and H. Schmidt, 2014: Wind shear and buoyancy reversal at the top of stratocumulus. J. Atmos. Sci., 71, 1040–1057, https://doi.org/10.1175/JAS-D-13-0189.1.
Mellado, J. P., C. S. Bretherton, B. Stevens, and M. C. Wyant, 2018: DNS and LES for simulating stratocumulus: Better together. J. Adv. Model. Earth Syst., 10, 1421–1438, https://doi.org/10.1029/2018MS001312.
Mellor, G. L., and T. Yamada, 1982: Development of a turbulence closure model for geophysical fluid problems. Rev. Geophys. Space Phys., 20, 851–875, https://doi.org/10.1029/RG020i004p00851.
Nakanishi, M., and H. Niino, 2006: An improved Mellor–Yamada level-3 model: Its numerical stability and application to a regional prediction of advection fog. Bound.-Layer Meteor., 119, 397–407, https://doi.org/10.1007/s10546-005-9030-8.
Nakanishi, M., and H. Niino, 2009: Development of an improved turbulence closure model for the atmospheric boundary layer. J. Meteor. Soc. Japan, 87, 895–912, https://doi.org/10.2151/jmsj.87.895.
Neggers, R. A. J., 2009: A dual mass flux framework for boundary layer convection. Part II: Clouds. J. Atmos. Sci., 66, 1489–1506, https://doi.org/10.1175/2008JAS2636.1.
Neggers, R. A. J., M. Kohler, and A. C. M. Beljaars, 2009: A dual mass flux framework for boundary layer convection. Part I: Transport. J. Atmos. Sci., 66, 1465–1487, https://doi.org/10.1175/2008JAS2635.1.
Olson, J. B., J. S. Kenyon, W. A. Angevine, J. M. Brown, M. Pagowski, and K. Sušelj, 2019: A description of the MYNN-EDMF scheme and the coupling to other components in WRF–ARW. NOAA Tech. Memo. OAR GSD-61, 42 pp., https://doi.org/10.25923/n9wm-be49.
Pedersen, J. G., S. P. Malinowski, and W. W. Grabowski, 2016: Resolution and domain-size sensitivity in implicit large-eddy simulation of the stratocumulus-topped boundary layer. J. Adv. Model. Earth Syst., 8, 885–903, https://doi.org/10.1002/2015MS000572.
Pincus, R., and B. Stevens, 2009: Monte Carlo spectral integration: A consistent approximation for radiative transfer in large eddy simulations. J. Adv. Model. Earth Syst., 1 (2), https://doi.org/10.3894/JAMES.2009.1.1.
Pleim, J. E., 2007: A combined local and nonlocal closure model for the atmospheric boundary layer. Part I: Model description and testing. J. Appl. Meteor. Climatol., 46, 1383–1395, https://doi.org/10.1175/JAM2539.1.
Reynolds, D. W., D. A. Clark, F. W. Wilson, and L. Cook, 2012: Forecast-based decision support for San Francisco International Airport: A NextGen prototype system that improves operations during summer stratus season. Bull. Amer. Meteor. Soc., 93, 1503–1518, https://doi.org/10.1175/BAMS-D-11-00038.1.
Siebesma, A. P., and J. Teixeira, 2000: An advection-diffusion scheme for the convective boundary layer: Description and 1d-results. 14th Symp. on Boundary Layer and Turbulence, Aspen, CO, Amer. Meteor. Soc., 4.16, https://ams.confex.com/ams/AugAspen/techprogram/paper_14840.htm.
Siebesma, A. P., P. M. M. Soares, and J. Teixeira, 2007: A combined eddy-diffusivity mass-flux approach for the convective boundary layer. J. Atmos. Sci., 64, 1230–1248, https://doi.org/10.1175/JAS3888.1.
Soares, P. M. M., P. M. A. Miranda, A. P. Siebesma, and J. Teixeira, 2004: An eddy-diffusivity/mass-flux parametrization for dry and shallow cumulus convection. Quart. J. Roy. Meteor. Soc., 130, 3365–3383, https://doi.org/10.1256/qj.03.223.
Stevens, B., 2002: Entrainment in stratocumulus-topped mixed layers. Quart. J. Roy. Meteor. Soc., 128, 2663–2690, https://doi.org/10.1256/qj.01.202.
Stevens, B., 2010: Introduction to UCLA-LES. 20 pp., https://www.mpimet.mpg.de/fileadmin/atmosphaere/herz/les_doc.pdf.
Stevens, B., and Coauthors, 2003: Dynamics and Chemistry of Marine Stratocumulus—DYCOMS-II. Bull. Amer. Meteor. Soc., 84, 579–594, https://doi.org/10.1175/BAMS-84-5-579.
Stevens, B., and Coauthors, 2005: Evaluation of large-eddy simulations via observations of nocturnal marine stratocumulus. Mon. Wea. Rev., 133, 1443–1462, https://doi.org/10.1175/MWR2930.1.
Sušelj, K., J. Teixeira, and D. Chung, 2013: A unified model for moist convective boundary layers based on a stochastic eddy-diffusivity/mass-flux parameterization. J. Atmos. Sci., 70, 1929–1953, https://doi.org/10.1175/JAS-D-12-0106.1.
Sušelj, K., M. J. Kurowski, and J. Teixeira, 2019a: On the factors controlling the development of shallow convection in eddy-diffusivity/mass-flux models. J. Atmos. Sci., 76, 433–456, https://doi.org/10.1175/JAS-D-18-0121.1.
Sušelj, K., M. J. Kurowski, and J. Teixeira, 2019b: A unified eddy-diffusivity/mass-flux approach for modeling atmospheric convection. J. Atmos. Sci., 76, 2505–2537, https://doi.org/10.1175/JAS-D-18-0239.1.
Teixeira, J., 1999: Simulation of fog with the ECMWF prognostic cloud scheme. Quart. J. Roy. Meteor. Soc., 125, 529–552, https://doi.org/10.1002/qj.49712555409.
Teixeira, J., and A. P. Siebesma, 2000: A mass-flux/K-diffusion approach to the parameterization of the convective boundary layer: Global model results. 14th Symp. on Boundary Layer and Turbulence, Aspen, CO, Amer. Meteor. Soc., P4.12, https://ams.confex.com/ams/AugAspen/techprogram/paper_14915.htm.
Teixeira, J., and Coauthors, 2011: Tropical and subtropical cloud transitions in weather and climate prediction models: The GCSS/WGNE Pacific Cross-Section Intercomparison (GPCI). J. Climate, 24, 5223–5256, https://doi.org/10.1175/2011JCLI3672.1.
Witek, M. L., J. Teixeira, and G. Matheou, 2011: An integrated TKE-based eddy diffusivity/mass flux boundary layer closure for the dry convective boundary layer. J. Atmos. Sci., 68, 1526–1540, https://doi.org/10.1175/2011JAS3548.1.
Wood, R., 2012: Stratocumulus clouds. Mon. Wea. Rev., 140, 2373–2423, https://doi.org/10.1175/MWR-D-11-00121.1.
Wood, R., and C. S. Bretherton, 2006: On the relationship between stratiform low cloud cover and lower-tropospheric stability. J. Climate, 19, 6425–6432, https://doi.org/10.1175/JCLI3988.1.
Wu, E., R. E. S. Clemesha, and J. Kleissl, 2018: Coastal stratocumulus cloud edge forecasts. Sol. Energy, 164, 355–369, https://doi.org/10.1016/j.solener.2018.02.072.
Yamaguchi, T., and D. A. Randall, 2012: Cooling of entrained parcels in a large-eddy simulation. J. Atmos. Sci., 69, 1118–1136, https://doi.org/10.1175/JAS-D-11-080.1.
Yang, H., and J. Kleissl, 2016: Preprocessing WRF initial conditions for coastal stratocumulus forecasting. Sol. Energy, 133, 180–193, https://doi.org/10.1016/j.solener.2016.04.003.
Zelinka, M. D., D. A. Randall, M. J. Webb, and S. A. Klein, 2017: Clearing clouds of uncertainty. Nat. Climate Change, 7, 674–678, https://doi.org/10.1038/nclimate3402.
Zhong, X., D. K. Sahu, and J. Kleissl, 2017: WRF inversion base height ensembles for simulating marine boundary layer stratocumulus. Sol. Energy, 146, 50–64, https://doi.org/10.1016/j.solener.2017.02.021.