1. Introduction
It is now generally accepted through observation (Lenschow et al. 1980) and through large eddy simulation (LES) studies (Moeng 1984) that turbulent diffusion in the convective atmospheric boundary layer (CABL) is associated with the nonlocal integral properties of the boundary layer. Therefore, the usual down-gradient parameterization of second and third moments for the CABL is not appropriate since such parameterizations do not adequately represent the complex mixing processes, including counter-gradient and entrainment effects, that affect the profiles of mean quantities.
The work of Deardorff (1972a) drew attention to the importance of accounting for counter-gradient effects in the formulation for the vertical heat flux in the atmospheric boundary layer (ABL). More recently, Holtslag and Moeng (1991) included the transport term in the heat flux equation, determined empirically from LES results, to obtain a different form for the counter-gradient term for the heat flux. They also proposed a nonlocal eddy diffusivity based on the earlier work of Troen and Mahrt (1986). Although these works, among others, have established the importance of accounting for such nonlocal and counter-gradient effects in specifying vertical fluxes, they are often not included in the ABL formulations used in atmospheric general circulation models (GCMs).
Part of the motivation for the present work is to develop a more physically realistic ABL parameterization for use in the atmospheric general circulation model of the Canadian Centre for Climate Modelling and Analysis. The scheme currently used is based on a local-K theory and does not include counter-gradient effects (McFarlane et al. 1992). It is qualitatively similar to other such schemes that, until relatively recently, have been widely employed in GCMs. The defects of such a scheme for GCM applications have been discussed by Holtslag and Boville (1993). These authors present evidence to demonstrate that use of a nonlocal ABL scheme, based on the work of Troen and Mahrt (1986), in the National Center for Atmospheric Research (NCAR) CCM2 leads to improvements in climate simulations, particularly over the tropical oceans.
Recently, Ayotte et al. (1996) compared one-dimensional cloud-free forms of a number of boundary layer parameterization schemes that are currently in use within GCMs, as well as fully nonlocal schemes such as the one proposed by Holtslag and Moeng (1991) and local turbulence schemes such as the level-2.5 (MY25) scheme of Mellor and Yamada (1974). They show that while most of the fully nonlocal schemes tend to substantially overpredict entrainment, most of the local schemes have a strong tendency to underpredict entrainment. In free-convective cases, none of the schemes were able to reasonably reproduce the LES results.
In this paper we present a new parameterization that includes a nonlocal eddy diffusivity and counter-gradient terms for the heat flux and the temperature variance. To derive this scheme we parameterize the third-order moments of scalar quantities (the flux of the heat flux and the flux of the temperature variance) by utilizing a general form suggested by a mass-flux concept (Randal et al. 1992). Rather than implementing the full mass-flux concept, we have used it as a guide to suggest the general form of the third moments. This is in part motivated by the desire to simplify the complex nonlinear system that results from the mass-flux approach. The simplification is obtained by utilizing results from LES (Moeng and Wyngaard 1989) and field experiments as documented in Lenschow et al. (1980). This leads to a nonlocal formulation for vertical fluxes in the atmospheric boundary layer in which a counter-gradient term appears in the expression for the vertical heat flux and the velocity variance. It turns out that for the heat flux the counter-gradient term contains both Deardorff’s (1972a) and Holtslag and Moeng’s (1991) counter-gradient terms as limiting forms, as well as an additional new term that is a result of the new formulation for the third moments. The nonlocal diffusivity is numerically close to the one suggested by Holtslag and Moeng (1991) but is quite different in form.
In principle, the full nonlinear system that is obtained by the mass-flux approach can be numerically solved. Although we have not done this, we have evaluated, a posteriori, using results from the simplified scheme, the various third-moment quantities as implied by the mass-flux approach and we find them to be in reasonable agreement with the simplified ones.
In the next section we present the turbulent scheme and discuss its derivation. In section 3 we present numerical results of comparisons with LES simulations. Concluding remarks are given in section 4.
2. The turbulence scheme
In this section we present the prognostic equations of the model and discuss the parameterization of turbulent quantities. The model includes prognostic equations for the mean variables and the turbulent kinetic energy (TKE). Momentum fluxes are derived from their steady-state equations by making assumptions on the relative importance of the various budgets in the momentum flux equation. For the momentum fluxes this leads to the usual down-gradient formulation. Other second-order moments are determined diagnostically through a parameterization of the third-order moments that is based on a convective mass-flux argument. Including these third-order moments in this way leads to parameterization for the heat flux that includes a nonlocal diffusivity and a counter-gradient term.
In this paper we limit attention to cloud-free boundary layers and we ignore for simplicity the effects of water vapor. We also ignore radiative effects. In these circumstances, the main effect of turbulence on the evolution of the larger scale mean state of the boundary layer is through convergence of the vertical fluxes of heat and momentum.
a. The third-order moments
A traditional closure approximation for the third moments in the foregoing equations involves representing these quantities as being entirely due to down-gradient diffusion (e.g., Stull 1988, 204). However, as noted, for example, by Zeman and Lumley (1976) and later Moeng and Wyngaard (1989), such a down-gradient diffusion assumption for the third moments, although adequate for stable and neutral conditions (when they are often relatively small), is inadequate for the convective regime in the planetary boundary layer (PBL). In the convective PBL, the transport of the third-order moments contributes significantly to the rate equations of the second moments. Therefore, an incorrect representation of the third-order moments will significantly degrade the accuracy of the fluxes and consequently of the mean state throughout the ABL, including the entrainment zone at its the top.
As an alternative approach we appeal to a convective mass-flux concept as a guide to representing the vertical transports of fluxes and variances of scalar quantities. As is demonstrated below, this leads to a parameterization of second-order moments in terms of nonlocal diffusivities and includes a counter-gradient term for the heat flux.
1) Parameterizing W ′2θ′ and W ′θ′2
The convective mass-flux concept has been used for parameterization of fluxes and variances in convectively active boundary layers in a number of studies (e.g., Betts 1973,1983; Wang and Albrecht 1990; Randal et al. 1992). It has also been employed in observational studies (e.g., Lenschow and Stephens 1982; Khasla and Greenhut 1985; Young 1988) and for analyses of the results of large eddy simulations (e.g., Moeng and Schumann 1991). Here we use the mass-flux concept as a guide to representing the third-moment terms
In Eq. (21) we retain the (z/h)1/3 dependence of
It is interesting to note that, using (21) and (22),
2) Parameterizing W ′q2 and W ′3
The assumptions of Therry and Lacarrère include
stationary equilibrium;
neglecting U, S, and M;
- neglecting the horizontal velocity variance contributions to the buoyancy term B, which implies
- after rewriting E asneglecting the terms that involve the gradient of the horizontal variances; and
- assuming a “return-to-isotropy” for the pressure termwhere c3 = 8.
Here it must be noted that this approximation for
3) The momentum flux transports
This balance is discussed by Therry and Lacarrère (1983), who depict the momentum flux budgets for the VOVES experiment. They also make the same assumption for their model. However, it must be noted that, although their results support the existence of such a balance in the lower part of the PBL, this is not so in the middle and upper part of the PBL. In fact, the transport term and the buoyancy term have the same sign in the middle and the upper part of the PBL. Therefore, a more accurate parameterization of the third-order moments in the momentum flux equation is needed. This will be investigated in future work.
b. The second-order moments
In this section we consider the second-order statistical moments. These include the heat flux
1) The heat flux
2) The momentum flux
As we will see in the results discussed in section 3, this formulation is not fully adequate for the convective boundary layer. The LES results show that nonlocal processes are significant in the transport of momentum. We are presently investigating a way of including these processes in the momentum flux formulation.
3) The turbulent kinetic energy
4) The vertical velocity and the temperature variances
In order to use the heat flux equation formulated earlier, we need to compute the vertical variance
Note that both
c. The timescale τ
d. The determination of the boundary layer height h
e. Summary of the boundary layer equations
Hereafter we designate the new boundary layer scheme that has been formulated above as the TKE model. In its one-dimensional version, this scheme includes the following main features.
A nonlocal eddy diffusivity (K) for the buoyancy flux as given by Eq. (39);
a counter-gradient term for the buoyancy flux as computed from Eq. (41);
a prognostic equation for the TKE with a corrective buoyancy term in the TKE flux as in Eq. (28) and with Therry and Lacarrère’s dissipation length scale as given by Eqs. (46) and (47); and
a down-gradient formulation for the momentum fluxes as given by Eqs. (43) and (44).
A complete list of the model equations is presented in appendix C.
3. Comparison with LES results
In this section the results from the TKE model simulation are compared with the LES output of Ayotte et al. (1996), who carried out an intercomparison of PBL parameterizations relative to the LES output. The LES code they used to generate the LES database was developed by Moeng (1984) and has 963 grid points. Particulars of the LES database and details on the constructions of the simulations are given in Ayotte et al. (1996).
Large-eddy simulation models explicitly simulate the larger-scale eddies, while they parameterize the net effects of the subgrid-scale eddies using inertial subrange theory. Since most of the turbulent energy in the PBL is due to the large-scale eddies, LES models have been very successful in capturing the observed features of convective PBL (e.g., Deardorff 1972b; Moeng and Wyngaard 1989; Mason 1989), shear-driven PBL (e.g., McWilliams et al. 1993; and Andrén and Moeng 1993), and stratus-topped PBL (e.g., Moeng et al. 1992).
The TKE simulation is initiated with the horizontal averages of the LES model fields that are obtained after running the LES model for approximately five large-eddy turnover times. The simulation is then carried out for 10 large-eddy turnover times and compared with the LES run, which is also integrated for 10 more large-eddy turnover times. The boundary conditions used in the TKE model are the same as those used, in the LES simulations including a specified surface heat flux. Although the results we present are run with a 40-m uniform vertical resolutions for the TKE model, as discussed below, the model behaves well and produces reasonable results at low resolutions as well. In comparing the results, we consider three different cases with distinct surface conditions, geostrophic wind, and initial mean profiles. The first case involves a strongly buoyant flow with a small shear, the second involves a shear driven flow with no surface heat flux, and the third case involves a free convection flow.
All our comparisons are carried out with the resolved-scale components of the LES output. The contribution of the subgrid-scale component is negligible except in the surface layer where our model utilizes similarity theory. At the top of the surface layer, we match the similarity results with our parameterization. Therefore the discrepancy we see in the surface layer is a result of not including the subgrid-scale contributions.
a. Case I: Buoyancy-driven PBL
Let us first consider the case where the PBL is buoyancy dominated with relatively small shear. The depth of the PBL in this case is on the order of 1000 m with a convective velocity scale (w∗) of approximately 2 m s−1, a friction velocity (u∗) of about 0.7 m s−1, and where the ratio h/L is about −15. Figure 2 shows the initial mean variable profiles for this case. The potential temperature profile shows a strong capping inversion on the order of 6 K and the ū and the
As shown in Fig. 3, where the final potential temperature and the heat flux profiles obtained from the TKE and the LES models are depicted, the TKE model is able to accurately reproduce the structure of the PBL, including the entrainment zone. The mean temperature profile is well mixed, which is typical of buoyancy-driven PBLs where the updrafts that are on the order of the PBL height cause nonlocal mixing. The heat flux is a linear function of height with negative values at the inversion layer that is associated with the entrainment of the overlying air. Consistent with the entrainment calculation, the downward heat flux at the inversion is about 20% of the upward surface heat flux.
The cross points in Figs. 2 and 3 are computed using a low vertical resolution similar to the third-generation Canadian Centre for Climate Modelling and Analysis GCM for which the new scheme is designed. The lowest level is at 60 m and the other levels have a uniform resolution of 200 m such that there are only 5 grid points in the convective PBL. Even at such a low vertical resolution, the model behaves well, and it is in good agreement with the higher resolution results and the LES output. Similar results were obtained for the other cases.
For the purposes of comparison, we also present in Fig. 5 the nondimensional counter-gradient term of our formulation and that of Holtslag and Moeng, which is given by Eq. (42). In the central part of the PBL they both give a value of 5, which corresponds to the dimensional value of 0.6 × 10−3 K m−1. This value is close to the original value proposed by Deardorff (1972a), which was 0.7 × 10−3 K m−1.
The vertical velocity variance normalized by
The temperature variance is shown in Fig. 7a. It decreases rapidly with height in the surface layer and becomes nearly constant in most of the mixed layer. There is a local maxima near the inversion level due to the large vertical inhomogeneity in the temperature that is caused by the large temperature jump in this part of the PBL.
The magnitude of the mean horizontal velocities ū and
The dimensionless profile of the sum of the three velocity variances,
b. Case II: Shear-driven PBL
We now consider the case with zero surface-flux and a strong shear. The PBL height h, in this case, is on the order of 500 m and the flow is strongly capped as in the previous case. Since there is no surface heat flux, both w∗ and θ∗ are zero, which, according to our parameterization of third-order terms, implies neglecting the turbulent transport terms in the heat flux and temperature variance equations and assuming a down-gradient transport for the turbulent kinetic energy. Therefore, the nonlocal characteristics of our parameterization are lost and all the turbulent fluxes and variances become down-gradient with a local eddy diffusivity Kloc that, in effect, reduces our parameterization to an equivalent of the level 2.5 Mellor–Yamada scheme (Mellor and Yamada 1974). The level 2.5 scheme is among a hierarchy of turbulent schemes developed by Mellor and Yamada, who used a systematic expansion procedure to analyze the departure of turbulent flows from the state of local isotropy, and it has been widely used in geophysical applications (Yamada 1977; Miyakoda et al. 1983; Gaspar et al. 1988; Crawford 1993). Although the level 2.5 model has a major deficiency in that it underestimates mixing and entrainment in the convective boundary layers, it has successfully been used to simulate the neutral and the stable boundary layers. Therefore we expect our model to perform well in these conditions as well.
As shown in Fig. 12, where the final profiles of the heat flux and the vertical velocity variance are depicted, good agreement is obtained between the LES and the TKE model. All the other turbulent quantities and the mean variables computed by the model are also in close agreement with the corresponding LES results.
c. Case III: Free-convective PBL
We now consider a free-convective PBL with the surface heating equal to Case I but with no geostrophic wind and u∗ = 0. As pointed out in Ayotte et al. (1996), most PBL models, even those with nonlocal and counter-gradient features substantially underpredict entrainment for this case. This may be due to the fact that fully nonlocal parameterizations ignore the partial local mixing that exists in the entrainment zone. On the other hand, single point models, such as the level 2.5 Mellor–Yamada, where the heat flux parameterization is based mainly on the local Richardson number, essentially ignore the nonlocal aspect of the mixed layer structure, resulting in weak entrainment.
The entrainment ratio R computed using Eq. (59) gives a value of 0.19 for our scheme, which is close to the LES computed value of 0.2. The heat flux profile and the profile of the vertical velocity variance for this case are also shown in Fig. 13. Although the vertical variance is slightly underpredicted in the central part of the PBL by the TKE model, the heat fluxes computed by the TKE model and the LES model are in agreement throughout the PBL. Similarly, Fig. 14 shows the normalized total variance
d. Skewness and third moments
As we have discussed in section 2, the use of the mass-flux concept leads to the parameterizations of the third-order moments
Figure 15 shows the skewness computed using the final profiles of
Figure 16 depicts the third moments computed using Eqs. (19) and (20) with the skewness and the variances calculated a posteriori from the model and the third moments computed with the approximated functions given by Eqs. (20) and (21), as well as the third moments computed from the LES model. Both
Although this comparison indicates that the simplified assumptions that reduced the complex skewness expression to the more manageable form are reasonable, it does not reveal the full impact of using the complete nonlinear set of differential equations for the third-order moments. This is because the full nonlinear feedback is ignored in this a posteriori computation.
4. Concluding remarks
We have developed a new second-order closure turbulence scheme for cloud-free boundary layers. While this scheme is similar in complexity to the Mellor–Yamada level 2.5 scheme, it differs significantly in the treatment of the convectively active boundary layer. In particular, the traditional down-gradient diffusion representation of second moments is replaced by a nonlocal representation that is derived from a new approach in the representation of third moments. This representation, while inspired by mass-flux concepts, also utilizes results from LES and field experiments.
The nonlocal character of the scheme is seen most clearly in the expression for the vertical heat flux. This quantity contains a counter-gradient term that combines previously derived expressions with an additional new term that results from accounting for the role of third moments in the heat flux equation.
Although the relevant third-moment terms have been parameterized in a simple way, this had led to a boundary layer scheme that accurately reproduces the results of LES for the vertical structure of both mean and turbulent quantities in cloud free boundary layers.
While the mass-flux approach was not used in detail in the present work, a posteriori calculations are in agreement with the basic concepts of the mass-flux approach, namely, that turbulence in convectively active boundary layers is dominated by deep plumelike eddies.
While the present work has been concerned mainly with cloud-free boundary layers where the turbulent quantities of most interest are the vertical fluxes of heat and momentum, in future work we will extend the scheme developed here to include effects of moisture and clouds. The approach used will be similar to that pioneered by Sommeria (1976) and Sommeria and Deardorff (1977), and used extensively for simulating the trade-wind cumulus boundary layer (Bougealt 1981) and for mesoscale applications (Bechtold 1992a).
Acknowledgments
The authors would like to thank Keith W. Ayotte and Peter P. Sullivan for providing the LES output used in the comparison. Special thanks to Peter Sullivan for redoing the LES runs in order to provide us with the LES third-order moments. We also like to thank the three anonymous referees for their suggestions for the improvement of the manuscript.
REFERENCES
André, J. C., G. De Moor, P. Lacarrère, G. Therry, and R. du Vachat, 1978: Modeling the 24-hour evolution of the mean and turbulent structures of the planetary boundary layer. J. Atmos. Sci.,35, 1861–1883.
Andrén, A., and C.-H. Moeng, 1993: Single-point closures in a neutrally stratified boundary layer. J. Atmos. Sci.,50, 3366–3379.
Arakawa, A., 1969: Parameterization of cumulus convection. Proc. WMO/IUGG Symp. Numerical Weather Prediction, Tokyo, Japan, Japan Meteorological Agency, 1–6.
Ayotte, K. W., and Coauthors, 1996: An evaluation of neutral and convective planetary boundary layer parameterizations relative to large eddy simulations. Bound.-Layer Meteor.,79, 131–175.
Bechtold, P., C. Fravalo, and J.-P. Pinty, 1992: A model of marine boundary-layer cloudiness for mesoscale applications. J. Atmos. Sci.,49, 1723–1744.
Betts, A. K., 1973: Nonprecipitating cumulus convection and its parameterization. Quart. J. Roy. Meteor. Soc.,99, 178–196.
——, 1983: Thermodynamics of mixed stratocumulus layers: Saturation point budgets. J. Atmos. Sci.,40, 2655–2670.
Bougeault, P., 1981a: Modeling the trade-wind cumulus boundary layer. Part I: Testing the ensemble cloud relations against numerical data. J. Atmos. Sci.,38, 2414–2428.
——, 1981b: Modeling the trade-wind cumulus boundary layer. Part II: A higher order one-dimensional model. J. Atmos. Sci.,38, 2429–2439.
Canuto, V. M., F. Minotti, C. Ronchi, R. M. Ypma, and O. Zeman, 1994: Second-order closure PBL with new third-order moments: Comparison with LES data. J. Atmos. Sci.,51, 1605–1618.
Crawford, G. C., 1993: Upper ocean response to storms—A resonant system. Ph.D. dissertation, University of British Columbia, 277 pp.
Deardorff, J. W., 1972a: Theoretical expression for the counter-gradient vertical heat flux. J. Geophys. Res.,77, 5900–5904.
——, 1972b: Numerical investigation of neutral and unstable planetary boundary layers. J. Atmos. Sci.,29, 91–115.
Gasper, P. Y., 1988: Modeling the seasonal cycle of the upper ocean. J. Phys. Oceanogr.,18, 161–180.
Holtslag, A. A. M., and C.-H. Moeng, 1991: Eddy diffusivity and countergradient transport in the convective atmospheric boundary layer. J. Atmos. Sci.,48, 1690–1698.
——, and B. A. Boville, 1993: Local versus nonlocal boundary layer diffusion in a global climate model. J. Climate,6, 1825–1842.
Kaimal, J. C., J. C. Wyngaard, D. A. Haugen, O. R. Coté, Y. Izumi, S. J. Caughey, and C. J. Readings, 1976: Turbulence structure in the convective boundary layer. J. Atmos. Sci.,33, 2152–2169.
Khasla, S. J. S., and G. K. Greenhut, 1985: Conditional sampling of updrafts and downdrafts in the marine atmospheric boundary layer. J. Atmos. Sci.,42, 2550–2562.
Lenschow, D. H., and P. L. Stephens, 1982: Mean vertical velocity and turbulence intensity inside and outside thermals. Atmos. Environ.,16, 761–774.
——, J. C. Wyngaard, and W. T. Pennel, 1980: Mean-field and second-moment budgets in baroclinic, convective boundary layer. J. Atmos. Sci.,37, 1313–1326.
Lumley, J. L., 1978: Computational modelling of turbulent flows. Adv. Appl. Mech.,18, 123–176.
Mason, P. J., 1989: Large eddy simulation of the convective atmospheric boundary layer. J. Atmos. Sci.,46, 1492–1516.
McFarlane, N. A., G. J. Boer, J. P. Blanchet, and M. Lazar, 1992: The Canadian Climate Centre second-generation general circulation model and its equilibrium climate. J. Climate,5, 1013–1043.
McWilliams, J. C., P. C. Gallacher, C.-H. Moeng, and J. C. Wyngaard, 1993: Modeling the oceanic planetary boundary layer. Large Eddy Simulation of Complex Engineering and Geophysical Flows, B. Galpperin and S. A. Orszag, Eds., Cambridge University Press, 600 pp.
Mellor, G. L., and T. Yamada, 1974: A hierarchy of turbulence closure models for planetary boundary layers. J. Atmos. Sci.,31, 1791–1806.
Miyakoda, K., T. Gordon, R. Caverly, W. Stern, J. Sirutis, and W. Bourke, 1983: Simulation of a blocking event in January 1977. Mon. Wea. Rev.,111, 846–869.
Moeng, C.-H., 1984: A large-eddy-simulation model for the study of planetary boundary layer turbulence. J. Atmos. Sci.,41, 2052–2062.
——, and J. C. Wyngaard, 1986: An analysis of closures for pressure-scalar covariances in the convective boundary layer. J. Atmos. Sci.,43, 2499–2513.
——, and ——, 1989: Evaluation of turbulent transport and dissipation closures in second-order modeling. J. Atmos. Sci.,46, 2311–2330.
——, and U. Schumann, 1991: Composite structure of plumes in stratus-topped boundary layers. J. Atmos. Sci.,48, 2280–2291.
——, S. Shen, and D. A. Randal, 1992: Physical processes within the nocturnal stratus-topped boundary layer. J. Atmos. Sci.,49, 2384–2401.
Randal, D. A., Q. Shao, and C.-H. Moeng, 1992: A second-order bulk boundary-layer model. J. Atmos. Sci.,49, 1903–1923.
Rotta, J. C., 1951: Statistische theorie nichthomogener turbulenz. Arch. Phys.,129, 547–572.
Sommeria, G., 1976: Three-dimensional simulation of turbulent processes in an undisturbed trade wind boundary layer. J. Atmos. Sci.,33, 216–241.
——, and J. W. Deardorff, 1977: Subgrid-scale condensation in models of nonprecipitating clouds. J. Atmos. Sci.,34, 344–355.
Stull, R. B., 1984: Transilient turbulence theory. Part I: The concept of eddy-mixing across finite distances. J. Atmos. Sci.,41, 3351–3367.
——, 1988: An Introduction to Boundary Layer Meteorology. Kluwer Academic, 666 pp.
Therry, G., and P. Lacarrère, 1983: Improving the eddy kinetic energy model for planetary boundary layer description. Bound.-Layer Meteor.,25, 63–88.
Troen, I., and L. Mahrt, 1986: A simple model of the atmospheric boundary layer: Sensitivity to surface evaporation. Bound.-Layer Meteor.,37, 129–148.
Wang, S., and B. A. Albrecht, 1990: A mean-gradient model of the dry convective boundary layer. J. Atmos. Sci.,47, 126–138.
Wyngaard, J. C., and O. R. Cote, 1975: The evolution of the convective planetary boundary layer—A higher-order closure model study. Bound-Layer Meteor.,7, 289–308.
Yamada, T., 1977: A numerical experiment on pollutant dispersion in a horizontally-homogeneous atmospheric boundary layer. Atmos. Environ.,11, 1015–1024.
Young, G. S., 1988: Turbulence structure of the convective boundary layer. Part II: Phoenix 78 aircraft observations of thermals and their environment. J. Atmos. Sci.,45, 727–735.
Zeman, O., and J. L. Lumley, 1976: Modeling buoyancy driven mixed layers. J. Atmos. Sci.,33, 1974–1988.
——, and ——, 1979: Turbulent Shear Flows. Vol. 1, Springer, 295 pp.
APPENDIX A
Solving the Heat Flux Equation
APPENDIX B
Surface Layer Matching
In order to utilize the closures presented in section 2 that apply above the surface layer, we match the PBL parameterizations with surface layer parameterizations from which we obtain the appropriate expressions for the Prandtl number Pr and the parameter α.
Unstable conditions
Neutral and stable conditions
The Prandtl number in this case is assumed to be 1.
APPENDIX C
Model Equations
The function S(η) defined by Eq. (34) and computed using the LES data for a typical convective PBL.
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
Mean initial profile for the buoyancy-driven PBL (Case I). (a) The mean potential temperature
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
Final profiles of (a) potential temperature, (b) nondimensionalized heat flux for the buoyancy-driven PBL (Case I) as calculated by the TKE model (dashed line), as calculated by the LES (solid line), and as calculated by the TKE model with very low resolution.
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The nondimensional eddy diffusivity K for heat as calculated by the TKE model (dashed line) and as calculated using the formulation of Holtslag et al. (1991) according to Eq. (62) (solid line).
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The nondimensional counter-gradient term for heat as calculated by the TKE model(dashed line) and as calculated using the formulation of Holtslag et al. (1991) according to Eq. (42) (solid line).
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profile of the nondimensional vertical velocity variance for the buoyancy-driven PBL (Case I) computed by the TKE model (dashed line) and by the LES (solid line).
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profile of the potential temperature variance for the buoyancy-driven PBL (Case I) computed by the TKE model (dashed line) and by the LES (solid line). (a) Without merging with the top-down variance, (b) with the merging of the top-down variance.
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profiles of the mean winds (a) ū and (b)
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profiles of the momentum fluxes (a)
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profile of the normalized
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profile of the nondimensional vertical flux of turbulent kinetic energy as computed by the TKE model.
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profiles of (a) the heat flux
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profiles of the nondimensional (a) heat flux
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The final profile of the normalized
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The vertical distribution of the skewness Sw as computed by the model.
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2
The third moments (a)
Citation: Journal of the Atmospheric Sciences 54, 14; 10.1175/1520-0469(1997)054<1850:ANSOTC>2.0.CO;2