1. Introduction
Realistic and computationally affordable multiscale simulations of atmospheric flows involve coupling of mesoscale and large-eddy simulation (LES) models. Such coupling requires efficient generation of three-dimensional turbulence at the inflow boundaries of the nested LES domains, which are forced with “smooth” mesoscale model fields. Mesoscale inflow does not contain any resolved eddies at the scales of atmospheric boundary layer (ABL) turbulence that the LES model can explicitly resolve with typical grid spacings of
The cell perturbation (CP) method applies stochastic perturbations to the potential temperature field that, through local buoyancy effects, break the two-dimensionality of the incoming flow and accelerate the development of three-dimensional turbulence. Perturbations are arranged as three horizontal cells of 8 × 8 grid points in the region near the inflow boundaries of the LES domain. This is done in order to effectively affect fully resolved modes in the Kolmogorov’s inertial range without being rapidly damped by the dissipative nature of the discretized advection term at higher wavenumbers. The CP method was generalized for neutrally stratified ABLs by Muñoz-Esparza et al. (2015), with all the required parameters related to available mesoscale information. The generalization for neutral ABL conditions uses the proposed perturbation Eckert number
Inflow generation in neutrally stratified conditions has historically received primary attention (e.g., Lund et al. 1998; di Mare et al. 2006; Xie and Castro 2008; Golaz et al. 2009; Nakayama et al. 2012). In the case of the CP method, the optimum
Herein, we first extend and then evaluate CP method to nonneutral ABLs. We focus on understanding the mechanisms that are responsible for turbulence transition and development in coupled mesoscale–microscale simulations under different stability conditions. The remainder of the paper is structured as follows. The numerical model, simulation setup, and cases of study are described in section 2. A modified Richardson scaling is proposed and tested in five stable ABLs in section 3, including analysis of turbulence statistics (variance, turbulent fluxes, and higher-order moments) and energy spectra and cospectra. In addition, an explanation for the transition in stably stratified conditions is proposed, which is then used to further refine the proposed scaling. A similar analysis for four characteristic convective ABLs is presented in section 4, proposing an alternative scaling using a thermal variance relationship. This relationship is further optimized based on the nature of the identified transition mechanisms. Finally, section 5 is devoted to summarizing the main findings of this work.
2. Methodology
a. Large-eddy simulation model and simulation setup
We perform simulations of idealized ABLs using the Weather Research and Forecasting (WRF) Model v3.8.1 (Skamarock and Klemp 2008; Skamarock et al. 2008). WRF is a compressible, nonhydrostatic code that solves the Navier–Stokes and energy equations for high Reynolds numbers (no viscous term) with filtering of acoustic modes. Equations are discretized over a staggered Arakawa C grid, with a terrain-following hydrostatic-pressure vertical coordinate. Time integration utilizes a third-order Runge–Kutta scheme, while the advection term is discretized using the hybrid even–odd-order numerical scheme from Kosović et al. (2016). This hybrid scheme improves effective spectral resolution (
To provide homogeneous inflow conditions to the LES domain, we utilize two domains with flat terrain that communicate via one-way nesting. The parent domain uses a one-dimensional PBL parameterization (Nakanishi and Niino 2006), and lateral boundaries are set to periodic in both directions. In the presence of the horizontally homogeneous forcing prescribed as a vertical sounding profile, this setup results in a uniform flow solution in the xy plane across the domain. The nested LES domain receives Dirichlet boundary conditions for all the prognostic equations provided by the mesoscale parent domain. A similar configuration has been used in previous studies targeting the development of turbulence generation methods with the WRF Model (e.g., Mirocha et al. 2014; Muñoz-Esparza et al. 2014).
Implicit filtering of the governing equations is used, which assumes that the attenuation of high wavenumbers present in the differencing schemes acts as a low-pass filter, and thus, the grid size acts implicitly as a filter. For the parameterization of subgrid-scale (SGS) effects on the momentum equations in the LES domain, we use the nonlinear backscatter and anisotropy (NBA) model from Kosović (1997), which is implemented in the open release of the WRF Model (Mirocha et al. 2010). From the two variants of the NBA scheme, we use the one that employs an additional prognostic equation for subgrid-scale turbulence kinetic energy (TKE) (Lilly 1966, 1967), following Deardorff (1980) and Moeng (1984). SGS thermal effects are parameterized using a countergradient approach, and the subgrid length scale for both momentum and heat is corrected to take into account possible length scale reductions due to local stable stratification (Deardorff 1980). Monin–Obukhov similarity theory (MOST) is applied at the first vertical grid point to provide surface boundary conditions for horizontal velocity components (vertical velocity is set to zero) and temperature (Jiménez et al. 2012). No cloud, microphysics, radiation, or land surface processes are considered.
b. Description of the simulated cases
The number of grid points used on the LES domain for both sets of simulations is 1800 × 900 × 150, in the streamwise, spanwise, and vertical directions, respectively. For the stable ABL cases, the LES domain extent is
List of cases, forcing conditions, and mean boundary layer characteristics for the stably stratified ABL simulations. All the cases used uniform grid spacings of

The resulting profiles of potential temperature and wind speed after 10 h of physical time integration of the mesoscale domain are shown in Fig. 1. As the cooling rate is increased from 0.25 to 0.75 K h−1, the boundary layer height progressively decreases, leading to wider low-level wind maxima with slightly enhanced velocities at the center of the jet and larger potential temperature gradients. As a result, a variety of local wind shear and temperature gradient scenarios is produced, which serves as a reasonable database to evaluate the CP method. To judge the adequacy of the grid size utilized, we calculate the Ozmidov length scale

Vertical profiles of (a) temperature θ and (b) ageostrophic wind component
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
While some recent LES studies of the SBL have utilized
For convective conditions, four different ABLs were considered. To set up these cases, two geostrophic wind conditions (
List of cases, forcing conditions, and mean boundary layer characteristics for the convective ABL simulations. All the cases used uniform grid spacings of

3. Turbulence generation in stably stratified ABLs
a. Perturbation amplitude based on a modified Richardson number
One aspect of the perturbation Eckert number that complicates its application to mesoscale–LES downscaling of stable boundary layers in real conditions is the use of the geostrophic wind in its definition. While idealized cases are typically forced by a geostrophic wind constant with height (e.g., Beare et al. 2006), SBLs often develop under the presence of baroclinicity. These horizontal temperature gradients originating from fronts or differential heating lead to the development of geostrophic wind shear (i.e., geostrophic forcing varying with height). Moreover, low-level jets occurring during stable conditions can present a strong ageostrophic component when induced by inertial oscillation mechanism (e.g., Vanderwende et al. 2015). This requires estimation of large-scale pressure gradient from mesoscale results, which further complicates the derivation of an accurate yet simple parameterization of the potential temperature perturbation amplitude that works under realistic forcing conditions.







In order for the thermal perturbations to result in an efficient destabilizing effect, the modified Richardson number should be
To have an initial qualitative assessment of the performance of the modified Richardson number scaling, the five mesoscale-to-LES nested SBL cases were simulated using

Instantaneous contours of vertical velocity w (m s−1) at a vertical height of
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
A series of simulations with the CP method were carried out to span a range of

Spatial development of vertical velocity variance
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
To assess the performance of the different
For completeness, the results from
Streamwise evolution of resolved turbulent kinetic energy

Spatial development within the nested LES domain of turbulence kinetic energy q, momentum flux
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
Spatial development of skewness and kurtosis in Fig. 4 reveals the signature of the above-described mechanism. Higher-order moments appear to be more sensitive to the perturbations, as seen in the evolution of
b. Energy spectra, turbulence length scale, and transition mechanism in the absence of inflow perturbations
To analyze the quality of the generated turbulence with the modified Richardson number-based perturbations, streamwise evolution of compensated spectra for the vertical velocity

Development of spanwise compensated velocity spectra of w and θ and
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
The

Development of spanwise compensated velocity spectra of w in the streamwise direction at
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
A closer inspection at the turbulent modes is provided by the quasi-equilibrium (x = 12.5 km,

Quasi-equilibrium compensated velocity spectra of w in the nested LES domain (
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
To gain further insight into the turbulence transition process in the absence of perturbations, vertical cross sections of potential temperature at

Instantaneous contours of potential temperature θ (K) at a spanwise location of
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
Finally, an explanation for the variability of optimum modified Richardson number from

(a) Correlation between the optimum modified Richardson number
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
4. Turbulence generation in convective ABLs
a. Perturbation amplitude based on a thermal variance scaling


From Eq. (2), and knowing

Spatial development of vertical velocity variance
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
b. Transition mechanisms in the absence of inflow perturbations
To elucidate the physical mechanisms responsible for the observed differences in turbulence generation and its equilibration, the temporal evolution of the potential temperature field is analyzed. Figure 11 displays potential temperature contours in a vertical plane at

Instantaneous contours of potential temperature θ (K) at a spanwise location of
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
In the U15H2 case, notable differences are observed (see Fig. 12) with respect to the U5H3 case. In spite of the onset of thermal instability originating at the surface at a close distance to lateral boundary of the nested LES domain, the growth of the instability is now delayed in space and remains in the near-surface region, t = 18 min. Organized updraft motions start to develop, progressively losing coherence and giving birth to smaller-scale eddies (t = 20, 22, 24 min). By the time the flow has reached its quasi-equilibrium state, t = 30 min, it can be clearly seen that the thermal instability at the surface does not lead to the onset of turbulence until

Instantaneous contours of potential temperature θ (K) at a spanwise location of
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
Analysis from the CBL cases clearly indicates a competition between the thermal instability that attempts to propagate vertically and the horizontal wind that tries to advect the instability in the downstream direction. Therefore, we propose the ratio

Spatial development within the nested LES domain of turbulence kinetic energy q, momentum flux
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
It is worth mentioning that the perturbation Eckert number proposed by Muñoz-Esparza et al. (2015) performs very well in all convective ABL scenarios analyzed herein, with only minor differences with respect to the
Finally, to illustrate the dependence of the amplitude and variability of the potential temperature perturbations within the ABL upon atmospheric stability, the vertical distributions of

Vertical distributions of the maximum amplitude of potential temperature perturbation
Citation: Monthly Weather Review 146, 6; 10.1175/MWR-D-18-0077.1
5. Conclusions
The generation of forcing-consistent turbulence in nested mesoscale-to-LES simulations of atmospheric boundary layers is one of the challenges hindering seamless multiscale modeling capabilities. Herein, we have extended our previous work on the cell perturbation method for nonneutral ABLs (stable and convective) and provided insight into the transition mechanisms. This is a relevant aspect, since in most cases the atmospheric boundary layer is influenced by stability effects, with neutral conditions mostly occurring on inland locations during the morning and evening transitions of the diurnal cycle. To that end, a set of nested LES within a mesoscale flow simulation has been performed, carefully designing the experiments to cover a broad range of the conditions that can be simulated using LES of ABLs with large yet affordable computing capabilities.
A modified Richardson number scaling is proposed for the stable ABL, which accounts for buoyancy suppression from the mesoscale inflow (i.e., positive vertical potential temperature gradient) and the instability effect of the thermal perturbations introduced under the CP method. Generally, it is found that
In the context of convective ABLs, perturbation amplitudes based on scaling of potential temperature variance in the mixed layer are proposed:
With the present work, we have provided understanding of the turbulence transition mechanisms in large-eddy simulations nested within nonneutral mesoscale-driven ABLs. Furthermore, we have presented stability-aware scalings of the optimum amplitude of the potential temperature perturbations with the CP method. While recent application of the CP method for a diurnal cycle by Muñoz-Esparza et al. (2017) has shown promise in enabling realistic multiscale modeling of ABL flows, we expect the current extensions of the method to further optimize the inflow turbulence generation aspect of coupled mesoscale–LES. We intend this contribution to enable more realistic multiscale simulations of flows in atmospheric models. We plan to provide the improvements developed herein to the community release version of WRF in the near term.
We would like to acknowledge high-performance computing support from Cheyenne (https://doi.org/10.5065/D6RX99HX) provided by NCAR’s Computational and Information Systems Laboratory. The National Center for Atmospheric Research is sponsored by the National Science Foundation. Computing time was provided by NCAR Strategic Capability Computing Grant NRAL0016. The authors are grateful to Jeremy Sauer and Robert Sharman for providing comments on an earlier version of the manuscript, as well as to two anonymous reviewers for their constructive comments.
REFERENCES
Allaerts, D., and J. Meyers, 2018: Gravity waves and wind-farm efficiency in neutral and stable conditions. Bound.-Layer Meteor., 166, 269–299, https://doi.org/10.1007/s10546-017-0307-5.
Basu, S., and F. Porté-Agel, 2006: Large-eddy simulation of stably stratified atmospheric boundary layer turbulence: A scale-dependent dynamic modeling approach. J. Atmos. Sci., 63, 2074–2091, https://doi.org/10.1175/JAS3734.1.
Basu, S., and A. Lacser, 2017: A cautionary note on the use of Monin–Obukhov similarity theory in very high-resolution large-eddy simulations. Bound.-Layer Meteor., 163, 351–355, https://doi.org/10.1007/s10546-016-0225-y.
Basu, S., A. A. Holtslag, B. J. Van De Wiel, A. F. Moene, and G.-J. Steeneveld, 2008: An inconvenient “truth” about using sensible heat flux as a surface boundary condition in models under stably stratified regimes. Acta Geophys., 56, 88–99, https://doi.org/10.2478/s11600-007-0038-y.
Beare, R. J., and Coauthors, 2006: An intercomparison of large-eddy simulations of the stable boundary layer. Bound.-Layer Meteor., 118, 247–272, https://doi.org/10.1007/s10546-004-2820-6.
Buckingham, S., L. Koloszar, Y. Bartosiewicz, and G. Winckelmans, 2017: Optimized temperature perturbation method to generate turbulent inflow conditions for LES/DNS simulations. Comput. Fluids, 154, 44–59, https://doi.org/10.1016/j.compfluid.2017.05.015.
Deardorff, J. W., 1980: Stratocumulus-capped mixed layers derived from a three-dimensional model. Bound.-Layer Meteor., 18, 495–527, https://doi.org/10.1007/BF00119502.
di Mare, L., M. Klein, W. Jones, and J. Janicka, 2006: Synthetic turbulence inflow conditions for large-eddy simulation. Phys. Fluids, 18, 025107, https://doi.org/10.1063/1.2130744.
Gibbs, J. A., E. Fedorovich, and A. Shapiro, 2015: Revisiting surface heat-flux and temperature boundary conditions in models of stably stratified boundary-layer flows. Bound.-Layer Meteor., 154, 171–187, https://doi.org/10.1007/s10546-014-9970-y.
Golaz, J.-C., J. D. Doyle, and S. Wang, 2009: One-way nested large-eddy simulation over the Askervein hill. J. Adv. Model. Earth Syst., 1, 6, https://doi.org/10.3894/JAMES.2009.1.6.
Grossman, R. L., 1982: An analysis of vertical velocity spectra obtained in the BOMEX fair-weather, trade-wind boundary layer. Bound.-Layer Meteor., 23, 323–357, https://doi.org/10.1007/BF00121120.
Huang, J., and E. Bou-Zeid, 2013: Turbulence and vertical fluxes in the stable atmospheric boundary layer. Part I: A large-eddy simulation study. J. Atmos. Sci., 70, 1513–1527, https://doi.org/10.1175/JAS-D-12-0167.1.
Jähn, M., D. Muñoz-Esparza, F. Chouza, O. Reitebuch, O. Knoth, M. Haarig, and A. Ansmann, 2016: Investigations of boundary layer structure, cloud characteristics and vertical mixing of aerosols at Barbados with large eddy simulations. Atmos. Chem. Phys., 16, 651–674, https://doi.org/10.5194/acp-16-651-2016.
Jiang, P., Z. Wen, W. Sha, and G. Chen, 2017: Interaction between turbulent flow and sea breeze front over urban-like coast in large-eddy simulation. J. Geophys. Res. Atmos., 122, 5298–5315, https://doi.org/10.1002/2016JD026247.
Jiménez, P. A., J. Dudhia, J. F. González-Rouco, J. Navarro, J. P. Montávez, and E. García-Bustamante, 2012: A revised scheme for the WRF surface layer formulation. Mon. Wea. Rev., 140, 898–918, https://doi.org/10.1175/MWR-D-11-00056.1.
Kosović, B., 1997: Subgrid-scale modelling for the large-eddy simulation of high-Reynolds-number boundary layers. J. Fluid Mech., 336, 151–182, https://doi.org/10.1017/S0022112096004697.
Kosović, B., and J. A. Curry, 2000: A large eddy simulation study of a quasi-steady, stably stratified atmospheric boundary layer. J. Atmos. Sci., 57, 1052–1068, https://doi.org/10.1175/1520-0469(2000)057<1052:ALESSO>2.0.CO;2.
Kosović, B., D. Muñoz-Esparza, and J. Sauer, 2016: Improving spectral resolution of finite difference schemes for multiscale modeling applications using numerical weather prediction model. 22nd Symp. on Boundary Layers and Turbulence, Salt Lake City, UT, Amer. Meteor. Soc., 3B.5, https://ams.confex.com/ams/32AgF22BLT3BG/webprogram/Paper295892.html.
LeMone, M. A., 1973: The structure and dynamics of horizontal roll vortices in the planetary boundary layer. J. Atmos. Sci., 30, 1077–1091, https://doi.org/10.1175/1520-0469(1973)030<1077:TSADOH>2.0.CO;2.
Lilly, D. K., 1966: On the application of the eddy viscosity concept in the inertial sub-range of turbulence. NCAR Manuscript 123, 19 pp.
Lilly, D. K., 1967: The representation of small-scale turbulence in numerical simulation experiments. NCAR Manuscript 281, 24 pp.
Lund, T. S., X. Wu, and K. D. Squires, 1998: Generation of turbulent inflow data for spatially-developing boundary layer simulations. J. Comput. Phys., 140, 233–258, https://doi.org/10.1006/jcph.1998.5882.
Lundquist, J., and L. Bariteau, 2015: Dissipation of turbulence in the wake of a wind turbine. Bound.-Layer Meteor., 154, 229–241, https://doi.org/10.1007/s10546-014-9978-3.
Mayor, S. D., P. R. Spalart, and G. J. Tripoli, 2002: Application of a perturbation recycling method in the large-eddy simulation of a mesoscale convective internal boundary layer. J. Atmos. Sci., 59, 2385–2395, https://doi.org/10.1175/1520-0469(2002)059<2385:AOAPRM>2.0.CO;2.
Mazzaro, L. J., D. Muñoz-Esparza, J. K. Lundquist, and R. R. Linn, 2017: Nested mesoscale-to-LES modeling of the atmospheric boundary layer in the presence of under-resolved convective structures. J. Adv. Model. Earth Syst., 9, 1795–1810, https://doi.org/10.1002/2017MS000912.
Mirocha, J., J. Lundquist, and B. Kosović, 2010: Implementation of a nonlinear subfilter turbulence stress model for large-eddy simulation in the advanced research WRF Model. Mon. Wea. Rev., 138, 4212–4228, https://doi.org/10.1175/2010MWR3286.1.
Mirocha, J., B. Kosović, and G. Kirkil, 2014: Resolved turbulence characteristics in large-eddy simulations nested within mesoscale simulations using the Weather Research and Forecasting Model. Mon. Wea. Rev., 142, 806–831, https://doi.org/10.1175/MWR-D-13-00064.1.
Moeng, C.-H., 1984: A large-eddy-simulation model for the study of planetary boundary-layer turbulence. J. Atmos. Sci., 41, 2052–2062, https://doi.org/10.1175/1520-0469(1984)041<2052:ALESMF>2.0.CO;2.
Muñoz-Esparza, D., B. Kosović, J. Mirocha, and J. van Beeck, 2014: Bridging the transition from mesoscale to microscale turbulence in numerical weather prediction models. Bound.-Layer Meteor., 153, 409–440, https://doi.org/10.1007/s10546-014-9956-9.
Muñoz-Esparza, D., B. Kosović, J. van Beeck, and J. Mirocha, 2015: A stochastic perturbation method to generate inflow turbulence in large-eddy simulation models: Application to neutrally stratified atmospheric boundary layers. Phys. Fluids, 27, 035102, https://doi.org/10.1063/1.4913572.
Muñoz-Esparza, D., J. K. Lundquist, J. A. Sauer, B. Kosović, and R. R. Linn, 2017: Coupled mesoscale-LES modeling of a diurnal cycle during the CWEX-13 field campaign: From weather to boundary-layer eddies. J. Adv. Model. Earth Syst., 9, 1572–1594, https://doi.org/10.1002/2017MS000960.
Muñoz-Esparza, D., R. D. Sharman, and J. K. Lundquist, 2018: Turbulence dissipation rate in the atmospheric boundary layer: Observations and WRF mesoscale modeling during the XPIA field campaign. Mon. Wea. Rev., 146, 351–371, https://doi.org/10.1175/MWR-D-17-0186.1.
Munters, W., C. Meneveau, and J. Meyers, 2016: Turbulent inflow precursor method with time-varying direction for large-eddy simulations and applications to wind farms. Bound.-Layer Meteor., 159, 305–328, https://doi.org/10.1007/s10546-016-0127-z.
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.
Nakayama, H., T. Takemi, and H. Nagai, 2012: Large-eddy simulation of urban boundary-layer flows by generating turbulent inflows from mesoscale meteorological simulations. Atmos. Sci. Lett., 13, 180–186, https://doi.org/10.1002/asl.377.
Salesky, S. T., M. Chamecki, and E. Bou-Zeid, 2017: On the nature of the transition between roll and cellular organization in the convective boundary layer. Bound.-Layer Meteor., 163, 41–68, https://doi.org/10.1007/s10546-016-0220-3.
Sauer, J. A., D. Muñoz-Esparza, J. M. Canfield, K. R. Costigan, R. R. Linn, and Y.-J. Kim, 2016: A large-eddy simulation study of atmospheric boundary layer influence on stratified flows over terrain. J. Atmos. Sci., 73, 2615–2632, https://doi.org/10.1175/JAS-D-15-0282.1.
Skamarock, W. C., 2004: Evaluating mesoscale NWP models using kinetic energy spectra. Mon. Wea. Rev., 132, 3019–3032, https://doi.org/10.1175/MWR2830.1.
Skamarock, W. C., and J. B. Klemp, 2008: A time-split nonhydrostatic atmospheric model for weather research and forecasting applications. J. Comput. Phys., 227, 3465–3485, https://doi.org/10.1016/j.jcp.2007.01.037.
Skamarock, W. C., and Coauthors, 2008: A description of the Advanced Research WRF version 3. NCAR Tech. Note TN-475+STR, 113 pp., https://doi.org/10.5065/D68S4MVH.
Sorbjan, Z., 1989: Structure of the Atmospheric Boundary Layer. Prentice Hall, 317 pp.
Sullivan, P. P., J. C. Weil, E. G. Patton, H. J. Jonker, and D. V. Mironov, 2016: Turbulent winds and temperature fronts in large-eddy simulations of the stable atmospheric boundary layer. J. Atmos. Sci., 73, 1815–1840, https://doi.org/10.1175/JAS-D-15-0339.1.
Vanderwende, B. J., J. K. Lundquist, M. E. Rhodes, E. S. Takle, and S. L. Irvin, 2015: Observing and simulating the summertime low-level jet in central Iowa. Mon. Wea. Rev., 143, 2319–2336, https://doi.org/10.1175/MWR-D-14-00325.1.
Weckwerth, T. M., T. W. Horst, and J. W. Wilson, 1999: An observational study of the evolution of horizontal convective rolls. Mon. Wea. Rev., 127, 2160–2179, https://doi.org/10.1175/1520-0493(1999)127<2160:AOSOTE>2.0.CO;2.
Xie, Z.-T., and I. P. Castro, 2008: Efficient generation of inflow conditions for large eddy simulation of street-scale flows. Flow Turbul. Combus., 81, 449–470, https://doi.org/10.1007/s10494-008-9151-5.
Zhou, B., and F. K. Chow, 2011: Large-eddy simulation of the stable boundary layer with explicit filtering and reconstruction turbulence modeling. J. Atmos. Sci., 68, 2142–2155, https://doi.org/10.1175/2011JAS3693.1.
Zhou, B., and F. K. Chow, 2014: Nested large-eddy simulations of the intermittently turbulent stable atmospheric boundary layer over real terrain. J. Atmos. Sci., 71, 1021–1039, https://doi.org/10.1175/JAS-D-13-0168.1.
The National Center for Atmospheric Research is sponsored by the National Science Foundation.