1. Introduction
Large-eddy simulation, or LES, is now very widely used in small-scale meteorology. Applications range from severe storms (Klemp 1987) to small-mesoscale phenomena (Cotton et al. 1993) and boundary layer meteorology (Moeng 1984; Mason and Thomson 1987; Nieuwstadt et al. 1993; Schumann 1993).
The strength of LES lies in its explicit calculation of the energy-containing eddies of a turbulent flow. The unresolvable or subgrid-scale (SGS) eddies, which are formally removed by spatial filtering of the governing equations (Leonard 1974), manifest themselves through SGS terms in the filtered or resolvable-scale equations. If the scale of the filter is small compared to the scale of the energy-containing eddies, the resolvable-scale eddies do contain most of the turbulent kinetic energy and fluxes. Furthermore, they are fairly insensitive to the details of the SGS model used to close the filtered equations. Thus, well-resolved LES has an appealing robustness.
Because the horizontal length scale of the vertical velocity fluctuations scales with distance from the surface, LES inevitably has inadequate spatial resolution near the surface. As a result, the SGS model takes on much more importance there than in other regions of the flow. Furthermore, the specification of resolvable-scale surface fluxes comes into question since the widely used surface-exchange coefficients are not obviously valid locally. Thus, the fidelity of LES within the surface layer is problematic.
In this paper we study three closely related aspects of LES in the surface layer. First we examine the behavior of the surface-exchange coefficients when the grid aspect ratio (horizontal dimension/vertical dimension) is of order 1, as in LES. As an alternative to surface-exchange coefficients, we develop conservation equations for the resolvable-scale surface fluxes. Simplifying these surface-flux equations requires our second development, scaling relations for LES variables in the surface layer. These determine how turbulent variances are apportioned between the resolvable and SGS components. Finally, in order to interpret the impact of the resolvable-scale surface-flux equations on LES fields we evaluate some aspects of the performance of the widely used eddy-diffusivity SGS model in the surface layer.
2. Surface-exchange coefficients for resolvable-scale fluxes
a. Background
The LES codes used in boundary layer meteorology (e.g., Nieuwstadt et al. 1993) involve the surface fluxes of momentum, temperature, water vapor, and trace constituents. Since these codes use local spatial averaging of the governing equations rather than ensemble averaging, these are resolvable-scale surface fluxes that contain not only the ensemble mean but also the fluctuations about this mean on scales up to the filter cutoff. Most such codes take this resolvable-scale surface flux as proportional to the product of the horizontal wind speed and the difference in the transported quantity between the surface and the first grid point. The proportionality factor is the surface-exchange coefficient.
b. Observed fluctuations in local surface-exchange coefficients
We evaluated (5) using LES data for the convective boundary layer and observations from the STORMFEST experiment. The LES domain was 5 km × 5 km in the horizontal and 2 km deep. It used a nested mesh in the surface layer so the resolution there was equivalent to that of a 2563 simulation (Khanna and Brasseur 1996). The inversion height zi was about 1000 m. In one run the geostrophic wind was 5.0 m s−1 and the surface temperature flux was 0.14 m s−1 K, giving −zi/L = 63. The other run had a geostrophic wind of 1.0 m s−1 and a surface temperature flux of 0.24 m s−1 K, giving −zi/L = 730. We took zs = 3.9 m, the lowest grid point, and evaluated the local surface-exchange coefficients at z1 = 12 and 27 m, using temperature as the scalar.
The STORMFEST data, taken by the NCAR Atmospheric Surface Transfer and Exchange Research facility, included time series of fluctuating temperature and the horizontal components of velocity from in situ sensors at z = 1 m, 4 m, and 10 m. The mean wind speed at 10 m was 10.7 m s−1, the surface temperature flux was 0.23 m s−1 K, the Monin–Obukhov length L was −63 m, and −zi/L ≈ 10. We took zs = 1.0 m, the lowest measurement point, and evaluated the local surface-exchange coefficients for temperature at z1 = 4 m and10 m.
The randomness factor in (5) involves wave-cutoff filtering at wavenumber magnitude κc in the horizontal wavenumber plane. We did this straightforwardly with the LES data, using several values of κc = π/Δ. This gave zs/Δ ⩽ 0.13. For the STORMFEST data we approximated the averaging by wave-cutoff filtering of the time series at frequency ωc, using Taylor’s hypothesis in the form κc = ωc/U. Here zs/Δ ⩽ 0.10.
Figure 2 shows the fluctuations in the local surface-exchange coefficient C̃H and its counterpart for momentum, C̃D, evaluated from measured time series and LES data, using (5) and its counterpart for C̃D. The fluctuation level increases as grid aspect ratio decreases, as we anticipated through the coupling-eddies notion depicted in Fig. 1, and it increases as −zi/L increases. It can be quite large.
This suggests that the surface-exchange coefficient fluctuations are due to the decreasing correlation between resolvable-scale fields very near the surface and at z1 as κcz1 → 1. Further analysis of these LES and STORMFEST data by Wyngaard and Peltier (1996) supports this notion.
An alternative to treating the resolvable-scale surface fluxes through surface-exchange coefficients is predicting them directly through their conservation equations. We discuss this next, focusing on the scalar flux. We treat the momentum flux in appendix C.
3. The subgrid-scale flux near the surface
The terms in (14), being resolvable scale, are bounded as z → 0 (we will use x3 and z interchangeably to denote height above the surface). In particular,
The spectrum (19) has a peak at κ ∼ 1/l and a κ−5/3 inertial range asymptote at κ ≫ l−1. Peltier et al. chose the constants c1 and c2 by requiring the variance and the inertial-range spectral level to agree with observations. They chose the length and intensity scales l and s as z and u∗ under neutral conditions. In free convection they chose the scales as the PBL depth zi and the convective velocity scale w∗ = (gQ0zi/T0)1/3, rather than the local-free-convection scales z and uf = (gQ0z/T0)1/3 (Wyngaard et al. 1971); here Q0 is the ensemble-mean surface temperature flux. As justification they cited the evidence that horizontal velocity fluctuations in the unstable surface layer are not Monin–Obukhov (M–O) similar, but instead scale with the mixed-layer scales zi and w∗ (Kaimal 1978; Wyngaard 1988).
Peltier et al. then combined the neutral and free convection forms to give an interpolation formula for the entire stability range between them. The spectrum E(κ) cannot be directly measured with conventional instruments, but Peltier et al. showed that the corresponding one-dimensional spectra agree well with existing measurements.
The modeled two-dimensional spectra are shown in Fig. 3 in area-preserving coordinates. For all but near-neutral conditions the horizontal spectrum Eh, which integrates to (
4. Scaling LES fields in the surface layer
We will use (29) to determine how the variances of uh, c, and u3 for z/Δ ∼ κcz ≪ 1 are partitioned into their resolvable and SGS parts. We will use (30) to determine how the variances of
a. Horizontal velocity and scalars
b. Vertical velocity
c. Vertical gradients
From (33), (35), and (51) we see that
5. Simplifying the resolvable-scale surface-flux budget
a. The “constant-SGS-flux” layer
We conclude that at small z/Δ, the rms difference in SGS flux between the surface and height z, Eq. (55), is small compared to the rms fluctuation level in SGS flux at height z, Eq. (56). If so, when z/Δ ≪ 1 the SGS flux at z is a reliable surrogate for the resolvable-scale surface flux.
It remains to evaluate the importance of the cross contributions to the SGS scalar flux, Eq. (17). In appendix B we show their rms values are of the order of F0 (z/Δ)3. Thus, at height z ≪ Δ they are not larger than the rms difference in SGS flux between the surface and height z. Since we judged the latter to be negligible, we will continue to neglect the cross contributions to SGS flux as well.
b. Scaling the flux budget
We can now scale some of the terms in the resolvable-scale surface flux budget (24). We choose the x1 direction to be that of the mean wind so that Ui = (U1, 0, 0). The scaling results are summarized in Table 1. We cannot directly scale pressure destruction, but from experience with the Reynolds flux budget close to the surface (Wyngaard et al. 1971) we expect that it is a leading term.
c. A hierarchy of resolvable-scale surface-flux budgets
According to the scaling results of Table 1 the budget of surface scalar flux takes different forms depending on the value of z1/Δ.
1) z1/Δ ≪ 1
The turbulence active in the processes represented in (58) is subgrid scale; it has a dominant length scale z and timescales Ts ∼ z/u∗ and z/uf in the neutral and free-convection limits, respectively. The final filtering of each term can be thought of roughly as averaging over horizontal spatial scales Δ ≫ z. The time-change term is negligible because its timescale is much greater than Ts. Similarly, the horizontal inhomogeneity of filtered variables is weak because it occurs on a spatial scale much larger than z. Thus, we argue that for z/Δ ≪ 1 the resolvable-scale surface flux budget (58) represents a quasi-steady, locally homogeneous state of“local grid-volume equilibrium.”
2) z1/Δ small but not negligible
3) z1/Δ ⩽ 1
According to our scaling analysis in appendix B, the fractional contribution of the cross terms to the SGS flux is one to two orders smaller in z/Δ than the SGS flux fluctuations produced by the terms retained in (74). Thus, there seems no need to include the cross contributions to SGS flux.
d. Resolution in z
Our scaling analysis of section 4 treats the effects of filtering the fields in the horizontal plane. Such filtering is done explicitly in Moeng’s (1984) LES code, for example. Our analysis does not treat the effects of filtering or finite-difference approximations in z. Thus, our surface-layer scaling results are directly applicable only to LES with high vertical resolution.
The vertical resolution of LES also figures in the choice of a resolvable-scale surface flux model. The tradeoff is clear: if z1/Δ ≪ 1, which requires the first grid point to be very close to the surface, a standard surface-exchange model for flux, Eq. (60), can be used. If z1/Δ ≈ 1, the first grid point can be much higher but then more physics enters the resolvable-scale surface flux conservation equation, as indicated in Eq. (75). Thus, when z1/Δ ≪ 1, the complicated connection between the resolvable-scale surface flux and the resolvable-scale flow at z ∼ Δ must be made entirely through the SGS model; when z1/Δ ≈ 1, some elements of that connection can be made through the resolvable-scale surface flux model.
6. Use of resolvable-scale surface-flux equations in LES
a. Implementation and results
Since z1/Δ is a critical parameter in our analysis, we carried out three sets of simulations: 643, 1282 × 64, and 1922 × 64, with z1/Δ of 0.13, 0.27, and 0.4, respectively, for a moderately convective boundary layer (−zi/L ≈ 10). Figure 2 shows the fluctuation levels in the surface-exchange coefficients obtained by using conservation equations for resolvable-scale surface fluxes in the simulations. These levels agree well with the findings of Wyngaard and Peltier (1996). The fluctuation levels in
Figures 4–6 show the nondimensional mean shear ϕm, mean potential temperature gradient ϕh, and vertical velocity variance, respectively, from 643 and 1282 × 64 simulations using the conservation equations (curves A and B) and the standard surface-exchange coefficients (curves 1 and 2). The profiles observed in past field experiments are also plotted in these figures.
Consistent with the experience of Mason and Thomson (1992) and Sullivan et al. (1994) ϕm is overpredicted, with respect to the observations, by the Smagorinsky-based SGS model (used in Moeng’s code) with surface-exchange coefficients. Moreover, the profile kink moves closer to the surface as the horizontal resolution is increased.
In the 643 simulation with surface-exchange coefficients, ϕh is somewhat lower than the measured value at the first grid point; above that it is slightly overpredicted. Increasing the horizontal resolution to 1282, while keeping the vertical resolution the same, limits the overprediction of ϕh to the second grid point. The values at other grid points agree well with the field observations.
The nondimensional vertical-velocity variance in both the 643 and the 1282 × 64 simulations with surface-exchange coefficients are slightly lower than the field measurements reported by Panofsky et al. (1977). The change in resolution does not alter the profile significantly.
The use of flux-conservation equations instead of the surface-exchange coefficients yields slight improvements in all three profiles, most noticeably in that of vertical-velocity variance. However, the improvement is only marginal and not wholly satisfactory. We believe that the mild influence of improved lower boundary conditions on the mean surface-layer structure is due to deficiencies in the SGS model. The objective of surface-flux budgets is to capture more reliably the local structure of the resolvable-scale surface fluxes, but, as we show next, the Smagorinsky-based SGS model represents the surface-layer physics poorly and therefore presumably responds to enhanced structure in surface fluxes improperly as well.
b. Analysis of the SGS model from highly resolved surface-layer fields
The Smagorinsky-type SGS model has been standard in LES since the early work of Lilly (1967) and Deardorff (1970). Its popularity stems from its simplicity and, most importantly, the insensitivity of resolved-scale fields to the SGS model in well-resolved turbulent flows. Near the surface, however, the vertical motions are always inadequately resolved and as a result the SGS closure becomes crucial there.
We analyzed the accuracy of Smagorinsky closure in the surface layer using highly resolved LES with nested meshes. Such tests have been done previously using direct numerical simulation (DNS) of isotropic turbulence (Clark et al. 1979; Bardina et al. 1983) and the neutral turbulent boundary layer (Piomelli et al. 1991). The unstable atmospheric surface layer, however, has certain distinct features not present in homogeneous turbulence and neutral boundary layers. Specifically, the horizontal motions in the surface layer of an unstable atmospheric boundary layer (ABL) are strongly influenced by the zi-scaled mixed-layer eddies, causing a strong anisotropy of length and intensity scales between the horizontal and vertical motions. There is also a substantial horizontal mean temperature flux in the surface layer that is absent in the neutral boundary layer. Thus, past studies using DNS fields are not directly relevant for our analysis.
We generated highly resolved surface-layer fields using a three-level nested-mesh LES discussed in Khanna and Brasseur (1996). The full boundary layer was simulated at −zi/L ≈ 10 using a 1283 mesh covering a domain of 5zi × 5zi × 2zi. The next level of refinement in the surface layer was attained by an effective 2563 mesh covering a domain of 5zi × 5zi × 0.125zi, and the final level was attained by an effective 5123 mesh covering 5zi × 5zi × 0.06125zi of the boundary layer. The upper boundary conditions for the embedded meshes were obtained from the next-level coarser domains with a one-way communication between the domains. At a height where the effective 5123 simulation resolved approximately 90% of the vertical fluxes and variances (the sixth grid level, at z/zi ≈ 0.02) the resolved variables were treated as fully resolved fields. They were decomposed into a resolvable part and a subgrid part by two-dimensional horizontal wave-cutoff filtering with a cutoff wavenumber corresponding to a 962 horizontal mesh.
Table 2 compares the calculated SGS stresses and fluxes and their divergences with the values predicted by the SGS model used by Moeng (1984). The magnitudes of the cross-correlation coefficients of the modeled and actual SGS flux components are very low (less than 0.20). Those for SGS flux divergences, which appear in the dynamical equations, are particularly low (less than 0.05 for the horizontal components of the divergences). The SGS model also fails in capturing the mean diagonal SGS stress components, presumably due to its inability to reproduce the anisotropic distribution of SGS energy, and in capturing the mean horizontal SGS temperature flux, due to the failure of the eddy-viscosity model of that flux (Wyngaard et al. 1971). These deficiencies in the SGS model, we hypothesize, can mask the effects of improved resolution of the structure of resolvable-scale surface fluxes.
c. Results with improved subgrid-scale model
Mason and Thomson (1992) found that adding stochastic fluctuations to the SGS stress divergence improved the nondimensional mean shear in a neutral boundary layer calculated through LES. These fluctuations simulate the local transfer of energy from unresolved to resolved scale, or “backscatter,” in the Smagorinsky SGS closure. We implemented this modification to the SGS model used in Moeng’s code along with the surface-flux conservation equations.
Figure 7 shows the effect of the SGS model and lower boundary conditions on the ϕm profile in the 1282 × 64 simulation of the moderately convective boundary layer. The original SGS model and drag coefficients overestimate ϕm at the first grid point (z/zi ≈ 0.03) by 50%. Use of surface-flux conservation equations with the original SGS model makes a marginal improvement. The modified SGS model (with stochastic backscatter) with the drag coefficients gives a smooth profile for ϕm, although it is not entirely in agreement with the observed profile. The simulated value increases with z/L before falling off, while the observed profile decreases monotonically. Our results are not in complete agreement with those of Mason and Thomson, presumably due to three factors: first, we are simulating a moderately convective boundary layer (with both buoyancy and shear effects) while their work concerned neutral boundary layers; second, we use a 1282 × 64 mesh, whereas Mason and Thomson used a much finer vertical mesh compared to the horizontal mesh; finally, some of the parameters used in our implementation of the stochastic-backscatter model are different from those of Mason and Thomson. Nevertheless, we infer from these results that the subgrid-scale model does have a significant influence on the mean structure of the surface layer. Use of surface-flux conservation equations along with the stochastic backscatter model makes a marginal improvement. We conclude that the improved lower boundary conditions need a compatible subgrid-scale model to make better predictions of atmospheric surface layers.
7. Summary and conclusions
We argued physically and showed through analysis of observations and LES data that the local surface-exchange coefficient relating the resolvable-scale surface flux to resolvable-scale properties of the overlying flow is a random variable. Its fluctuation level increases with z/Δ, where z is height above the surface and Δ is the spatial scale of the filter that separates fields into resolvable-scale and subgrid-scale parts. As z/Δ → 0, the surface-exchange coefficient approaches its traditional definition.
An alternative to using a local surface-exchange coefficient to diagnose the resolvable-scale surface flux is predicting that flux through its conservation equation. We showed that near the surface this flux is dominated by the s–s component and we derived the equation for that component. We used the surface-layer spectral model of Peltier et al. (1996) to develop scaling expressions for the partitioning of the variances of surface-layer fields between the resolvable and subgrid-scale components. We used these scaling expressions to simplify the flux conservation equations in three limits.
For high aspect ratio grids, z1/Δ ≪ 1, the surface-flux conservation equation is in a state of local grid volume equilibrium and the mean surface-exchange coefficient can be used locally, as is generally done in LES. For smaller aspect ratio grids it has horizontal advection and time-change terms, consistent with the appearance of fluctuations in the local surface-exchange coefficient in this regime. As the grid aspect ratio approaches unity, the flux conservation equation gains a production term proportional to the convergence of horizontal velocity near the surface. The observed fluctuations in the local surface-exchange coefficient in this regime are quite large.
Based on Bradshaw's (1969) suggestion that horizontal inhomogeneity of the surface layer causes the local mean gradient of a transported quantity to deviate from its equilibrium value, we proposed a simple closure for the flux conservation equations and implemented them in the Moeng (1984) LES code.
Use of the surface flux conservation equations yields slight improvements in the nondimensional mean shear, mean potential temperature gradient, and vertical-velocity variance profiles. The commonly observed kink in the mean shear and temperature gradient profiles predicted by Smagorinsky-based SGS models at z ≈ Δ is reduced. This improvement, however, is not substantial.
We showed through highly resolved LES that the Smagorinsky-based SGS models perform poorly in the atmospheric surface layer; a better SGS model is needed. We are currently extending the present analysis to derive conservation equations for SGS stresses and fluxes in the atmospheric surface layer, which, combined with dynamic lower boundary conditions, will hopefully make significant improvements in the LES predictions of atmospheric surface-layer structure.
Acknowledgments
We are grateful to T. Horst and G. Maclean of NCAR SSSF for kindly providing ASTER data from the STORMFEST experiment; to Andrew R. Brown of U.K. Meteorological Office for helpful discussions on stochastic backscatter; and to J. Brasseur, P. Mourad, and P. Sullivan for making helpful suggestions on the manuscript. This work was supported by Army Research Office Grant DAAL03-92-G-0117 and by Office of Naval Research Grant N00014-92-J-1688.
REFERENCES
Bardina, J., J. H. Ferziger, and W. C. Reynolds, 1983: Improved turbulence models based on large-eddy simulation of homogeneous, incompressible, turbulent flows. Tech. Rep. TF-19, Stanford University, Stanford, CA, 174 pp.
Bradshaw, P., 1969: Comments on “On the relation between the shear stress and the velocity profile after change in surface roughness.” J. Atmos. Sci.,26, 1353–1354.
Businger, J. A., J. C. Wyngaard, Y. Izumi, and E. F. Bradley, 1971: Flux-profile relationships in the atmospheric surface layer. J. Atmos. Sci.,28, 181–189.
Clark, R. A., J. H. Ferziger, and W. C. Reynolds, 1979: Evaluation of subgrid-scale models using an accurately simulated turbulent flow. J. Fluid Mech.,91, 1–16.
Cotton, W. R., R. L. Walko, K. R. Costigan, P. J. Flatau, and R. A. Pielke, 1993: Using the Regional Atmospheric Modeling System in the large-eddy-simulation mode: From inhomogeneous surfaces to cirrus clouds. Large-Eddy Simulation of Complex Engineering and Geophysical Flows, B. Galperin and S. Orszag, Eds., Cambridge University Press, 369–398.
Deardorff, J. W., 1970: A numerical study of three-dimensional turbulent channel flow at large Reynolds numbers. J. Fluid Mech.,41, 453–480.
——, 1980: Stratocumulus-capped mixed layers derived from a three-dimensional model. Bound.-Layer Meteor.,18, 495–527.
Kaimal, J. C., 1978: Horizontal velocity spectra in an unstable surface layer. J. Atmos. Sci.,35, 18–24.
Khanna, S., and J. G. Brasseur, 1996: Analysis of Monin–Obukhov similarity from large-eddy simulation. J. Fluid Mech.,345, 251–286.
Klemp, J. B., 1987: Dynamics of tornadic thunderstorms. Annu. Rev. Fluid Mech.,19, 369–402.
Leonard, A., 1974: Energy cascade in large eddy simulation of turbulent flows. Advances in Geophysics, Vol. 18A, Academic Press, 237–248.
Lilly, D. K., 1967: The representation of small-scale turbulence in numerical simulation experiments. Proc. Tenth IBM Scientific Computing Symp. on Environmental Sciences, IBM Thomas J. Watson Research Center, Yorktown Heights, NY, 195–210.
Lumley, J. L., and H. A. Panofsky, 1964: The Structure of Atmospheric Turbulence. Interscience, 239 pp.
Mason, P. J., and N. S. Callen, 1986: On the magnitude of the SGS eddy coefficient in LES of turbulent channel flows. J. Fluid Mech.,162, 439–462.
——, and D. J. Thomson, 1987: Large-eddy simulation of the neutral-static-stability planetary boundary layer. Quart. J. Roy. Meteor. Soc.,113, 413–443.
——, and ——, 1992: Stochastic backscatter in large-eddy simulations of boundary layers. J. Fluid Mech.,242, 51–78.
Moeng, C.-H., 1984: A large-eddy-simulation model for the study of planetary boundary layer turbulence. J. Atmos. Sci.,41, 2052–2062.
Nieuwstadt, F. T. M., P. J. Mason, C.-H. Moeng, and U. Schumann, 1993: Large-eddy simulation of the convective boundary layer:A comparison of four computer codes. Turbulent Shear Flows, F. Durst, R. Friedrich, B. E. Launder, F. W. Schmidt, U. Schumann, and J. H. Whitelaw, Eds., Vol. 8, Springer-Verlag, 343–368.
Panofsky, H. A., and J. A. Dutton, 1984: Atmospheric Turbulence. Wiley, 397 pp.
——, H. Tennekes, D. H. Lenschow, and J. C. Wyngaard, 1977: The characteristics of turbulent velocity components in the surface layer under convective conditions. Bound.-Layer Meteor.,11, 355–361.
Peltier, L. J., J. C. Wyngaard, S. Khanna, and J. G. Brasseur, 1996: Spectra in the unstable surface layer. J. Atmos. Sci.,53, 49–61.
Piomelli, U., W. H. Cabot, P. Moin, and S. Lee, 1991: Subgrid-scale backscatter in turbulent and transitional flows. Phys. Fluids A,3, 1766–1771.
Schumann, U., 1993: Large-eddy simulation of turbulent convection over flat and wavy surfaces. Large-Eddy Simulation of Complex Engineering and Geophysical Flows, B. Galperin and S. Orszag, Eds., Cambridge University Press, 399–421.
Stull, R., 1988: An Introduction to Boundary Layer Meteorology. Kluwer, 402 pp.
Sullivan, P., J. C. McWilliams, and C.-H. Moeng, 1994: A subgrid-scale model for large-eddy simulation of planetary boundary-layer flows. Bound.-Layer Meteor.,71, 247–276.
Tennekes, H., and J. L. Lumley, 1972: A First Course in Turbulence. The MIT Press, 300 pp.
Wyngaard, J. C., 1982: Planetary boundary layer modeling. Atmospheric Turbulence and Air Pollution Modelling, F. T. M. Nieuwstadt and H. van Dop, Eds., Reidel, 69–106.
——, 1988: Structure of the PBL. Lectures on Air-Pollution Modeling, A. Venkatram and J. C. Wyngaard, Eds., Amer. Meteor. Soc., 9–61.
——, and L. J. Peltier, 1996: Experimental micrometeorology in an era of turbulence simulation. Bound.-Layer Meteor.,78, 71–86.
——, O. R. Coté, and Y. Izumi, 1971: Local free convection, similarity, and the budgets of shear stress and heat flux. J. Atmos. Sci.,28, 1171–1182.
APPENDIX A
Simplifying the Resolvable-Scale Surface Scalar Flux Budget
Deriving the budget of resolvable-scale surface scalar flux as indicated in (23) yields
APPENDIX B
The Cross Components of SGS Flux
APPENDIX C
Extension to Resolvable-Scale Surface Stress
The resolvable-scale surface stress budget
The scaling procedure discussed in the text yields the estimates, shown in Table C1, of the magnitudes of the terms in (C5).
The constant-SGS-stress layer
In a similar analysis, Mason and Thomson (1992) assumed that as the surface is approached the vertical stress gradient is balanced by the horizontal pressure gradient, which they estimated to be roughly independent of height. Their estimate of vertical stress gradient, therefore, differs from ours. The conclusion regarding the constant-SGS-stress layer, however, still holds albeit for a smaller z/Δ. The discrepancy can only be resolved through direct numerical simulations.
The cross stresses
The surface stress budget for z1/Δ ≪ 1
The surface stress budget for z1/Δ < 1
Table C1 shows that in free convection there is not the clear separation in the order of the terms that existed for scalar flux. The next-order terms in (C5) are of order (z/Δ)2/3, but after that come a number of zero-mean,“noise terms” (e.g., RSCFI, RGCSI) of order z/Δ.
Despite this lack of separation, the two budgets are similar in that the next-order terms are large-scale advection, resolvable-scale gradient production, and the vertical part of resolvable-scale shear production. We argued in the paper that the time-change term is of the order of advection, which we will take to include that by the mean velocity.
We showed in section 2 of this appendix that the cross contribution is relatively more important for stress than it is for scalar flux. Thus, when z1/Δ is not very small so that a surface stress conservation equation rather than the usual surface-exchange expression is appropriate, it appears that the cross stress could also be significant. We showed that the dominant cross stress is (
A schematic of the “coupling eddies” in the grid volume adjacent to the surface. They couple the resolvable-scale flow at height z1 to the SGS surface flux. For a unity aspect ratio grid (left) there is only one such coupling eddy; for a large aspect ratio (right) there are many.
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
Behavior of the local surface-exchange coefficients as calculated from high-resolution LES and STORMFEST data taken by the NCAR ASTER facility. Circles: STORMFEST at 4 m (open) and 10 m (solid); −zi/L ≈ 10. Squares: LES at 12 m (open) and 27 m (solid); −zi/L ≈ 63. Triangles: LES at 12 m (open) and 27 m (solid); −zi/L ≈ 800. Plus symbols: LES at 16 m using surface-flux equations.
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
The two-dimensional, surface-layer spectra of a scalar (dashed line), vertical velocity (bold line), and horizontal velocity (fine line), as modeled by Peltier et al. (1996). The vertical lines indicate the κz value below which lies 50% of the variance. Top: neutral; center: sightly unstable; bottom: free convection.
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
Nondimensional mean shear in a moderately convective boundary layer (−zi/L ≈ 10). Curves 1 and 2 are from 643 and 1282 × 64 simulations, respectively, using surface-exchange coefficients;curves A and B are from 643 and 1282 × 64 simulations, respectively, using surface-flux equations; solid line is the empirical function of Businger et al. (1971)
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
Nondimensional mean temperature gradient in a moderately convective boundary layer (−zi/L ≈ 10). Curves 1 and 2 are from 643 and 1282 × 64 simulations, respectively, using surface-exchange coefficients; curves A and B are from 643 and 1282 × 64 simulations, respectively, using surface-flux equations; solid line is the empirical function of Businger et al. (1971).
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
Nondimensional vertical-velocity variance in a moderately convective boundary layer (−zi/L ≈ 10). Curves 1 and 2 are from 643 and 1282 × 64 simulations, respectively, using surface-exchange coefficients; curves A and B are from 643 and 1282 × 64 simulations, respectively, using surface-flux equations; solid line is the empirical function of Businger et al. (1971).
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
Effect of SGS model and lower boundary conditions on the nondimensional mean shear. Curves 1 and 2 are from the original SGS model (without backscatter) and curves A and B from the modified SGS model (with backscatter). In both sets, the first curve is based on surface-exchange coefficients and the second on the surface flux equations.
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
Fig. A1. The spectra of vertical temperature flux (long dashes), horizontal temperature flux (short dashes), and vertical velocity (solid) from high-resolution LES at z = 35 m (fine) and 66 m (bold) for −zi/L ≃ 65.
Citation: Journal of the Atmospheric Sciences 55, 10; 10.1175/1520-0469(1998)055<1733:LITSLS>2.0.CO;2
Scaling of terms in the budget of surface scalar flux.
A comparison of predicted and actual subgrid-scale temperature flux and stress components from an effective 5123 simulation, using nested meshes, of the atmospheric boundary layer (−zi/L ≈ 8). The comparisons are made at z/zi ≈ 0.04 and z/Δ ≈ 0.75; (q̃ = F̃i,i; ãi = τ̃ij,j).
Table C1. Scaling of terms in the budget of surface stress.