1. Introduction
Thermal stratification also influences the SGS energy spectrum of turbulence, which in turn violates Lilly’s assumption of a long inertial subrange in deriving cs. In particular, the coefficient has to be decreased in stably stratified conditions. This trend is reflected in Deardorff’s (1980) empirical model, as well as in the model of Brown et al. (1994), who derive a stability-dependent model from the SGS energy equation assuming a state of local equilibrium. Canuto and Cheng (1997) employ a two-point closure to construct the SGS energy spectrum under the influence of shear and buoyancy. From the SGS energy spectrum they derive an analytical expression for the reduction of cs under shear and buoyancy. Like stratification, the presence of mean shear also requires decreasing the Smagorinsky coefficient.
Another approach, quite different from using analytical expressions that model the postulated dependence of cs as a function of flow parameters, is the so-called dynamic model of Germano et al. (1991). In the dynamic model the resolved turbulence at scales between Δ and 2Δ is analyzed statistically during the simulation, and coefficients are deduced from appropriate averaging operations. There is considerable evidence that the dynamic model is able to provide realistic predictions of the coefficient, at least when the filter scale Δ and the test-filter scale 2Δ are smaller than the local integral scale of turbulence. Porté-Agel et al. (2000a) generalize the dynamic model to include scale dependence and show successful application to a neutral boundary layer.
In addition to the dependence of cs on these parameters, the great variability of turbulence dynamics in general, and of atmospheric dynamics in particular, raises the issue of how the averaging procedures needed in evaluating terms in Eq. (7) should be performed, and how meaningful the results are. Variability in cs is caused by the inherent intermittency of turbulence, and of ABL flow patterns in particular. It is well known that the SGS dissipation Πmeas in turbulence is highly intermittent. This was already shown for isotropic turbulence using DNS by Cerutti and Meneveau (1998) and for the ABL in the context of the SGS dissipation of scalar variance by previous experiments described in Porté-Agel et al. (2000b, 2001a,b). To examine the effects of intermittency upon eddy-viscosity coefficients, the averages in the nominator and denominator of Eq. (7) can be computed over different timescales Tc. Then cs is no longer a single value but fluctuates from one time period (of length Tc) to another. We wish to examine how this variability is affected by varying Tc under different flow conditions. Moreover, in LES using the Lagrangian dynamic model (Meneveau et al. 1996), one needs to prescribe a timescale. This timescale is used in that model to set the duration of averaging over the history of turbulence following fluid trajectories.
Before proceeding to the details of the present data and analysis, it is worthwhile to delineate two aspects of the deliberately focussed scope of the present study: First, we restrict attention to the basic structure of the eddy-viscosity Smagorinsky closure. This closure is based on the assumption that the SGS stresses and fluxes are aligned to the gradients of velocity and temperatures. The drawbacks of this assumption have already been documented extensively in the literature. As reviewed in Meneveau and Katz (2000) and Tao et al. (2002) in the context of experimental studies in laboratory turbulence, the alignment hypothesis is not accurate. In the context of ABL turbulence, Higgins et al. (2003) confirm this limitation and show that addition of a so-called tensor eddy-diffusion model improves the alignment trends. Moreover, near the ground, Tong et al. (1999) show that the streamwise accelerations inherent in the eddy-viscosity closures cause unphysical couplings with the resolved velocity field. Even with these limitations, the eddy-viscosity closure is still the most often used in practical applications, providing continued interest in the dependence of cs on physical flow parameters as studied here.
The present paper is organized as follows. In section 2, we describe the field experiment and the dataset used in the present study. Section 3 presents results on the magnitude of the measured cs as function of atmospheric stability and distance to the ground. An empirical fit to represent the dependence of cs on the flow parameters for easy use in the context of LES is presented. Section 4 studies the dependence of cs on the local strain-rate magnitude. Section 5 examines in more detail the issue of statistical variability of cs, as function of averaging timescale and atmospheric stability. Section 6 shows results from a similar analysis for the SGS heat flux. Conclusions are presented in section 7.
2. The Horizontal Array Turbulence Study (HATS) dataset
In the context of LES for the atmospheric boundary layer, a number of field studies have aimed at measuring qi and τij from field data and at analyzing the results to improve SGS modeling. A study using data from a single 3D sonic anemometer (Porté-Agel et al. 1998) restricted the analysis to one-dimensional filtering (time filtering and interpreting the results as filtering in the x1 = x direction using Taylor’s hypothesis), whereas the filtering assumed in Eq. (1) is in all three directions. Henceforth, x1 = x, x2 = y, and x3 = z will be used interchangeably. Tong et al. (1998) proposed deploying a horizontal array of sensors and examined filtering issues using LES-data. Their results showed that filtering in two horizontal directions was required for quantitatively more accurate results. Experimental results from one horizontal array of sensors using two-dimensional filtering were reported in Tong et al. (1999) and Porté-Agel et al. (2000b). The latter paper showed that, while filter dimensionality did not have a strong effect on the previously reported trends based on one-dimensional filtering, atmospheric stability had strong effects on the results. Limiting the setup of Porté-Agel et al. (2000b) was the inability to compute vertical derivatives. This issue was addressed by using two vertically displaced horizontal arrays as proposed in Tong et al. (1999), and also in the Davis 1999 experiment (Porté-Agel et al. 2001a). As described in the next paragraphs, a similar setup is used in the present study [Horizontal Array Turbulence Study (HATS)] now including two more anemometers, and including more data under stable stratification, due to prevailing wind conditions at night.
HATS was conducted in the San Joaquin Valley close to Kettleman City, California, from 31 August until 1 October 2000. The field site was selected because of its homogeneous surface conditions with predictable wind directions. It was located 5.6 km east-northeast of Kettleman City at the southeast corner of an area of unplanted farmland. Homogeneous surface conditions ranged at least 2 km in the upwind (northwest) direction. Vegetation consisted of crop stubble and weeds for which the displacement height d0 and roughness length z0 were calculated to be 32 and 2 cm, respectively. As outlined in the introduction the goal of the experiment was the examination of SGS quantities for a wide range of stabilities Δ/L and array geometries Δ/z. The requirement of computing derivatives in all directions necessitated a setup of 3D sonic anemometers in two parallel horizontal arrays, which are separated in the vertical direction and centered in the lateral direction (see Fig. 1). Variation in Δ/z was achieved by selecting four setups with different geometrical arrangements (see Table 1), each of which was in the field for 6–9 days with continuous sampling in order to record data for a wide range of stabilities Δ/L. A total of 14 Campbell Scientific three-component sonic anemometer–thermometers (CSAT3) were partitioned into one array with nine sonics and another array with five instruments. The former allows for computation of double-filtered quantities and is named the subscript “d” array, while the latter is referred to as subscript “s” array as in single filtered. An additional two sonics were mounted on a reference tower to examine flow obstruction. For additional information see Horst et al. (2003, unpublished manuscript).
All 16 sonics were calibrated before and after the experiment in the National Center for Atmospheric Research (NCAR) wind tunnel and differences in the slope of regressions for the 16 sonics were in a range on the order of 2%. The standard deviation of the slope of the regressions was less than 0.5%. All sonics met the specification of the manufacturer of an intercept of less than 4 cm s–1, only one had an offset of 6 cm s–1 after the experiment. Other errors stem from the alignment of the sonic anemometers. Errors in the alignment of the x–y plane of the sonic anemometers parallel to the surface can be corrected for in postprocessing assuming that the mean wind vector is parallel to the local surface. This tilt was found to be less than 2°. The x axes of all sonics should be parallel to each other and perpendicular to the x–z array plane. The error in this alignment was measured onsite with a theodolite. After correcting the data with the theodolite measurements intercomparisons of horizontal wind components of the instruments still showed offsets of up to 6 cm s–1 and residual wind direction biases of up to 2°. This paragraph summarizes the descriptions in Horst et al. (2003, unpublished manuscript), where a more detailed data quality analysis is presented.
The temperature measurements are uncalibrated. However, the present analysis does not involve any vertical gradients of mean temperature, but only gradients of temperature fluctuations. By subtracting the mean temperature from each instrument’s measurement, the remaining error is due to experimental uncertainty in the output. It is specified by the manufacturer as 0.026 K.
The arrays were oriented in a way that southeastward winds (315°) were perpendicular to the arrays and caused the least interinstrumental flow obstruction. For our analysis, all time periods with an angle of the downstream pointing array normal and 6.8-min averaged wind vector of –30° < α < 30° are considered. Excluding all data violating this criterion leaves us with the amount of data specified in the second column of Table 1. During data processing, the array is rotated to a position perpendicular to the prevailing wind using Taylor’s hypothesis. The center of rotation for both arrays is the center sonic (same y coordinate). The new (rotated) velocity for a sonic with distance δy from the center sonic for given mean horizontal velocity vector 〈u〉 and angle of average wind vector with the array normal α is
Filtered quantities that were defined as a continuum in Eq. (1) have to be computed using discrete filters as specified in Table 1. Many LES codes use a 2D spectral cutoff filter in horizontal planes. However, this filter is not suited for our analysis, because its slow x–1 decay in physical space aggravates its approximation with O(5) sensors. Moreover the spatial cutoff filter produces a spatially nonlocal impact when filtering spatially localized phenomena (“ringing”). Thus, we choose to use spatially localized filters, which can be well represented by the experimental arrangement. In the lateral (y) direction trapezoidal filter functions are used with the exception of array 4, for which a top-hat filter is used for the d array in order to match the filter sizes of s and d arrays. For increased smoothness Gaussian filter functions are applied in the streamwise (x) direction where a higher resolution is available due to the 20-Hz sampling that corresponds to a sampling distance of about 0.12 m, using Taylor’s hypothesis. Filtering is done in wave space using the Fourier transform of the Gaussian filter function ĜΔ = exp[–(
3. Dependence of cs on stability and height
In order to study the effect of stability and height, the data are divided into segments of length TL (we mostly use TL = 6.8-min-long segments containing 213 points), that are classified in terms of Obukhov length L [defined according to Eq. (8)], and height Δ/z. To illustrate the total amount of data, the cumulative duration of all segments in each Δ/L bin and Δ/z bin is shown in Fig. 2. As can be seen, more data are available in the near-neutral bins while less data are available in the more stable bins. There are ∼40 h of useful data for each array, which implies that there is more data available for the Δ/z < 0.7 case, because data from arrays 3 and 4 are combined in this bin. As outlined in the introduction, in the paper various averaging timescales Tc will be used to compute cs from Eq. (7).
Figure 3a shows that the most likely value of
The mean and the variability of
Figure 4 shows results for cs from averaging over segments of length Tc = TL = 13.7 min ∼ 283Δ/〈u〉 for the four different arrays. The data for arrays 3 and 4 are combined since they correspond to similar values of Δ/z. As is visible, even after averaging over times corresponding to 283 filter length scales, there is significant variability. Nevertheless, it is seen that for all stabilities, the cs values for large Δ/z tend to fall below those for low Δ/z, a trend that is consistent with previous results (Mason 1994; Porté-Agel et al. 2000a, 2001b). In order to identify more clearly the trends with Δ/L and Δ/z, averages are performed over the entire data available.
Figure 5 shows results for cs from averaging SGS energy dissipations over all segments within each Δ/L bin of Fig. 2. Thus, these results correspond to using Tc equal to the times indicated in Fig. 2 in each case. A very clear dependence of the coefficient on Δ/L and Δ/z can be identified. Considering the heterogeneity of the data within one bin with respect to wind angle, turbulence intensity, mean velocity, etc., it is reassuring that such clear trends emerge from the data. From its neutral value, cs decreases strongly under stable atmospheric conditions. Moreover, a larger Δ/z leads to a decrease in the model coefficient, consistent with the use of damping functions for cs close to the wall, where z becomes equal to or smaller than Δ.
To further examine the validity of the proposed expression we consider the simultaneous limit of large Δ/L and large Δ/z. For this limit (and n ≥ 1, say), Eq. (17) reduces to cs ∼ (Δ/L)–1 (Δ/z)–1. To test this asymptotic trend, in Fig. 6b, cs is plotted versus Δ/z × Δ/L for all arrays. Indeed, for large Δ/L and large Δ/z cs follows closely the line cs ∼ (Δ2/(Lz))–1, justifying the proposed fit in Eq. (17). This suggests that for Δ ≫ L and Δ ≫ z the value of cs is determined by the product of the two length scales L and z rather than by the smaller of the two.
To fit the parameters of Eq. (17) to the data in Fig. 5 we set n = 3 and fit c0 and α using multidimensional unconstrained nonlinear optimization from MATLAB. Mason and Brown (1999) suggest n = 2, but the small differences between the cs of different arrays in neutral and unstable conditions are indication of a slower decrease of cs with Δ/z, which requires a larger n. From the optimization with n = 3, we obtain c0 = 0.1347, and α = 0.1289. Since the difference between c0 and α is within the range of experimental uncertainty, we assume α = c0 = 0.135. The resulting equation is used for the fits in Fig. 5, as well as in the preceding Figs. 3 and 4.
The proposed fit is tested by comparison with a different set of data, namely from array 1 in which a box filter is applied on four adjacent sonics in the s array and the corresponding sonics in the d array. This results in a filter scale of Δ = 26.8 m and a value of Δ/z = 8.6. Using a one-sided derivative in the y direction and a centered derivative in the x direction, the quantities needed to compute
As a further test of the proposed fit, Fig. 7 compares the measured cs for an averaging time Tc = TL = 13.7 min with the value obtained from Eq. (17). It can be concluded that the empirical fit represents the mean trends in the data also for the shorter (compared to Fig. 5) averaging time. However, for unstable conditions (large cs), deviations between the modeled and the measured cs occur due to the large variability of the measured cs, whereas the model fit yields a constant value of cs for any given value of Δ/z. Also, for arrays 3 and 4 (Δ/z < 0.7) the scatter in the data is larger than for arrays 1 and 2. This might be caused by the difference in setup geometry of array 3. There, the single-filtered array is below the double-filtered array (see Table 1), which influences and possibly overestimates vertical derivatives compared to the other setups. For array 4, different filter types in the lateral direction are used for the single- and double-filtered arrays, as indicated in Table 1.
Analyses by other investigators have revealed similar results. Deardorff (1971) and Piomelli et al. (1988) both found cs ≈ 0.1 for small Δ/z. Porté-Agel et al. (2001a) found cs ≈ 0.08 which is about 35% smaller than ours, but the tendency of an increase of the coefficient with Δ/z is the same.
The proposed expression in Eq. (17) can be easily used in LES, since Δ/L and Δ/z are known parameters that are imposed in the simulations a priori by the choice of mesh spacing, wall shear stress, and heat flux at the boundary. If the dependence on stratification is to be expressed as function of Richardson number, relationships between Ri and L/z can be used such as those appearing in Businger et al. (1971). However, most of the recent work dealing with stability of the lower atmosphere has tended to be in terms of L (Brutsaert 1982).
Finally, we report the coefficient values that are obtained from matching momentum flux instead of dissipation, according to Eq. (13). Figure 8 shows the coefficients so determined for various Δ/z and Δ/L. Comparing with Fig. 5, we see that the coefficients are much larger. LES with such values are known to be overly damped and thus we conclude that the condition of correct energy dissipation is more appropriate for the data analysis. The impossibility to choose a cs that satisfies both the requirements of producing the correct rate of kinetic energy transfer from the resolved to the subgrid-scales Π and the correct subgrid-scale stress τij is a basic flaw of the eddy-viscosity model. For further information consult Meneveau (1994), Pope (2000, p. 603), and Juneja and Brasseur (1999).
4. Dependence of cs on local strain-rate magnitude
The basic scaling inherent in the Smagorinsky model, predicated upon inertial-range dimensional arguments, assumes that the eddy viscosity is linearly proportional to the local strain-rate magnitude | S̃ | [see Eq. (3)]. Whether this concept is justified can be examined by evaluating cs from subsets of the data in which | S̃ | has certain values. If the Smagorinsky scaling is correct, the measured value of cs should be independent of strain-rate magnitude. Thus, in this section we further classify the available data according to the local strain-rate magnitudes for conditional sampling. Since the data must also be classified into different ranges of stabilities, the limited amounts of data under each condition become an issue. In order to assure sufficient amounts of data in each condition, data segments of TL = 6.8 min are classified into six ranges of stability—unstable to neutral (Δ/L ≤ 0)—and several ranges of increasing stability—(0 < Δ/L < 0.1, 0.1 < Δ/L < 0.5, 0.5 < Δ/L < 1.5, 1.5 < Δ/L < 3, and Δ/L > 3).
In order to isolate the effect of strain-rate magnitude, the conditional
The implications for the Smagorinsky model are as follows. We conclude that the deeper Δ is in the inertial range [Δ ≪ min(z, L)] the more
5. Variability of cs
In this section we address the question “how variable is cs?” Results shown in section 3, specifically Figs. 3a and 3b, suggest that while the most likely value of
Figure 12a shows that the spread in the pdf of cs increases for decreasing Tc for unstable atmospheric stability. Reassuringly, however, the most likely value of cs and the median (as shown in Fig. 13a) do not depend on Tc. For stable conditions (Fig. 12b), the most likely value and the median (Fig. 13a) of cs are constant with Tc and smaller than for unstable conditions, in agreement with the findings in section 3. The fact that the medians of cs are independent of Tc for stable and unstable conditions is encouraging for LES with dynamic SGS models which, as discussed in the introduction, often use some kind of averaging procedures, either in space (e.g., horizontal planes) or time [e.g., the Lagrangian dynamic model (Meneveau et al. 1996)] to compute the coefficient. Our results suggest that correct median coefficients can be obtained even for fairly short averaging timescales. Rather surprisingly, however, in the case of stable conditions it appears that the spread in the pdf does not decrease for increasing Tc.
Figure 13b presents a quantification of the width of the pdfs as function of Tc. Instead of computing the rms value (which tends to be biased due to some outliers in the distribution), we quantify the spread of the pdfs with quartiles. The figure shows the difference between the third and first quartile of the distribution, normalized by the second quartile (thus giving a dimensionless measure of the variability that is not strongly affected by atypical outliers). The relative width of the pdf for the stable bin does not decrease as Tc is increased. This result shows strong variability of the real and/or modeled SGS dissipation under stable atmospheric conditions indicating that fluctuations occur over very long timescales. This may be related to the strong intermittency in stable atmospheric conditions.
The fraction of segments of length Tc that display average backscatter (with negative
6. Results for coefficients in scalar models
As introduced in Eq. (12), the coefficient for the Smagorinsky model for the SGS heat flux
By dividing
Repeating the analysis of section 3,
In order to quantify the variability of PrT, the analysis of section 5 is repeated. All data segments with –2.0 < Δ/L < 0.0 (unstable bin) and 1.5 < Δ/L > 5.5 (stable bin) are selected and PrT(Δ/L) is computed with varying averaging times Tc. Then the quartiles of the resulting probability distribution of PrT are obtained and the median q2 is plotted in Fig. 17a. In contrast to our findings concerning cs, the median of the Prandtl number is not constant, but increases with Tc. This explains the difference between Figs. 16b and 15b, in which PrT computed from averages over several hours in Fig. 16b was significantly larger than PrT computed from 102.4-s averages in Fig. 15b. The increase with averaging time appears to level off for Tc > 102 s. For all Tc, the median for very stable conditions is larger than the median for unstable conditions, but they seem to converge for large Tc. A similar behavior (but with different magnitudes of Prandtl numbers) is observed for the other arrays. The dependence of the median of PrT on the averaging time and the large scatter in Fig. 15b complicate the development of empirical expressions for PrT and
In comparing with prior results, we can remark that for small Δ/z, Mason and Derbyshire (1990), Moin et al. (1991), and Porté-Agel et al. (2001a) found PrT ∼ 0.4, which is within the range of uncertainity around our value of PrT(Δ/z < 0.7) = 0.49. For large Δ/z, Porté-Agel et al. (2001a) examined two 30-min segments whose Δ/z roughly correspond to the values for our arrays 1 and 2. For the setup similar to our array 2 they obtain PrT ∼ 0.5 for Δ/L = –0.26, their analysis of the setup similar to our array 1 results in PrT ∼ 0.6 for Δ/L = –1.18. Our results from Table 2 suggest PrT = 0.60 and PrT = 0.67, which is qualitatively consistent and within the range of experimental uncertainty.
The spread of the pdf of PrT is shown in Fig. 17b as a function of Tc. For unstable atmospheric stability conditions, (q3 – q1)/q2 decreases from a value of 1.7 to 0.3 for Tc ranging from Tc = 0.05 s to 6.8 min. For very stable conditions, the variability is constant between 0.3 and 0.6 for the entire range of Tc. This is in agreement with findings for the variability of cs in section 5. Possibly due to the intermittency in stable conditions the variability does not decrease for larger averaging times, while in unstable conditions the variability decreases significantly. The results for arrays 1, 3, and 4 are very similar.
7. Conclusions
Parameters of the Smagorinsky model for the SGS shear stress and the SGS heat flux have been studied based on a statistical analysis of a large dataset (157 h) of ABL turbulence. Model coefficients have been measured based on the condition of equivalence between real and modeled SGS dissipation of kinetic energy and scalar variance. Several trends have been identified. Consistent with prior results in the literature, near the ground it is found that cs depends on the ratio of filter length and height above the ground Δ/z and decreases as Δ/z is increased. Moreover, cs depends strongly on atmospheric stability as parameterized by the length-scale ratio Δ/L. The previously postulated decrease of cs in stable stratification and shear (Deardorff 1980; Canuto and Cheng 1997) is quantified from the data, and an empirical formula [Eq. (17)] for cs is proposed. By varying the time Tc over which the SGS energy dissipations are averaged, we find that the variability in cs decreases with increasing Tc for unstable to neutral conditions, whereas, in very stable conditions, the variability in cs is independent of averaging time. The fact that in either case the median of cs is independent of averaging time confirms the robustness of the results. It also supports the assumption inherent in Lagrangian dynamic SGS models that coefficients can be obtained from data by averaging over timescales that are not overly long.
The dependence of cs on local strain-rate magnitude has also been studied here. Since the Smagorinsky model already assumes proportionality of the eddy viscosity νT to strain-rate magnitude | S̃ | , cs should be independent of strain-rate magnitude. The data suggest that this is correct for unstable to neutral conditions or for small strain-rate magnitudes. However, in stable conditions and for large strain-rate magnitudes, cs decreases with strain-rate magnitude. In very stable conditions the data are consistent with a
A similar analysis is carried out for the coefficient of the SGS heat flux
Finally, the basic flaws of the eddy-viscosity models need to be pointed out. Even perfect knowledge of the coefficient does not result in correct prediction of both energy transfer from the resolved scales to the subgrid scales and the momentum fluxes associated with the SGS stress. Moreover, the basic proportionality assumption of the Smagorinsky model τij ∝ Δ2 | S̃ | S̃ij is contradicted by tensorial misalignment between SGS stress and strain rate (Tao et al. 2002), independent of the value of cs.
Acknowledgments
The authors wish to thank Tom Horst, Donald Lenschow, Chin-Hoh Moeng, Peter Sullivan, and Jeffrey Weil from the NCAR ATD and MMM divisions for this fruitful collaboration during the field experiment. Thanks also to Profs. W. Eichinger, F. Porté-Agel, S. Richardson, and J. Wyngaard for the loan of sonic anemometers. The authors gratefully acknowledge funding from the National Science Foundation Grant NSF-ATM 01300766.
REFERENCES
Brown, A. R., P. J. Mason, and S. H. Derbyshire, 1994: Large-eddy simulation of stable atmospheric boundary layers with a revised stochastic subgrid model. Quart. J. Roy. Meteor. Soc., 120:1485–1512.
Brutsaert, W., 1982: Evaporation into the Atmosphere: Theory, History and Applications. D. Reidel, 299 pp.
Businger, J. A., J. C. Wyngaard, Y. Izumi, and E. F. Bradley, 1971: Flux–profile relationships in the atmospheric boundary layer. J. Atmos. Sci., 28:181–189.
Canuto, V. M. and Y. Cheng, 1997: Determination of the Smagorinsky–Lilly constant cs. Phys. Fluids, 7:1368–1378.
Carati, D., G. Winckelmans, and H. Jeanmart, 2001: On the modelling of the subgrid-scale and filtered-scale stress tensors in large-eddy simulation. J. Fluid Mech., 441:119–138.
Cerutti, S. and C. Meneveau, 1998: Intermittency and relative scaling of the subgrid dissipation rate in turbulence. Phys. Fluids, 10:928–937.
Cerutti, S. and C. Meneveau, 2000: Statistics of filtered velocity in grid and wake turbulence. Phys. Fluids, 12:1143–1165.
Cerutti, S., C. Meneveau, and O. M. Knio, 2000: Spectral and hyper eddy viscosity in high-Reynolds-number turbulence. J. Fluid Mech., 421:307–338.
Clark, R. A., J. H. Ferziger, and W. C. Reynolds, 1979: Evaluation of subgrid models using an accurately simulated turbulent flow. J. Fluid Mech., 91:1–16.
Deardorff, J. W., 1970: A numerical study of three-dimensional turbulent channel flow at large Reynolds numbers. J. Fluid Mech., 41:453–480.
Deardorff, J. W., 1971: On the magnitude of the subgrid-scale eddy coefficient. J. Comput. Phys., 7:120–133.
Deardorff, J. W., 1980: Stratocumulus-capped mixed layers derived from a three dimensional model. Bound.-Layer Meteor., 18:495–527.
Germano, M., U. Piomelli, P. Moin, and W. H. Cabot, 1991: A dynamic subgrid-scale eddy viscosity model. Phys. Fluids, 3A:1760–1765.
Higgins, C., M. B. Parlange, and C. Meneveau, 2003: Alignment trends of velocity gradients and subgrid scale fluxes in the turbulent atmospheric boundary layer. Bound.-Layer Meteor., in press.
Hunt, J. C. R., D. D. Stretch, and R. E. Britter, 1988: Length scales in stably stratified turbulent flows and their use in turbulence models. Stably Stratified Flow and Dense Gas Dispersion, J.S. Puttock, Ed., Clarendon Press, Oxford, 285–322.
Juneja, A. and J. Brasseur, 1999: Characteristics of subgrid-resolved-scale dynamics in anisotropic turbulence with application to rough-wall boundary layers. Phys. Fluids, 11:3054–3068.
Kang, H. S. and C. Meneveau, 2002: Universality of large eddy simulation model parameters across a turbulent wake behind a heated cylinder. J. Turbul., 3.paper 032.
Lesieur, M. and O. Metais, 1996: New trends in large-eddy simulations of turbulence. Annu. Rev. Fluid Mech., 28:45–82.
Lilly, D. K., 1967: The representation of small-scale turbulence in numerical simulation experiments. Proc. IBM Scientific Computing Symp. on Environmental Sciences, Yorktown Heights, NY, Thomas J. Watson Research Center, 195–209.
Mason, P. J., 1994: Large-eddy simulation: A critical review of the technique. Quart. J. Roy. Meteor. Soc., 120:1–26.
Mason, P. J. and S. H. Derbyshire, 1990: Large eddy simulation of the stably-stratified atmospheric boundary layer. Bound.-Layer Meteor., 53:117–162.
Mason, P. J. and D. J. Thomson, 1992: Stochastic backscatter in large-eddy simulations of boundary layers. J. Fluid Mech., 242:51–78.
Mason, P. J. and A. R. Brown, 1999: On subgrid models and filter operation in large eddy simulation. J. Atmos. Sci., 56:2101–2114.
Meneveau, C., 1994: Statistics of turbulence subgrid-scale stresses: Necessary conditions and experimental tests. Phys. Fluids, 6A:815–833.
Meneveau, C. and J. Katz, 2000: Scale-invariance and turbulence models for large-eddy-simulation. Annu. Rev. Fluid Mech., 32:1–32.
Meneveau, C., T. Lund, and W. Cabot, 1996: A Lagrangian dynamic subgrid-scale model of turbulence. J. Fluid Mech., 319:353–385.
Moeng, C-H., 1984: A large-eddy-simulation model for the study of planetary boundary-layer turbulence. J. Atmos. Sci., 41:2052–2062.
Moin, P. and J. Kim, 1982: Numerical investigation of channel flow. J. Fluid Mech., 118:341–377.
Moin, P., K. Squires, W. Cabot, and S. Lee, 1991: A dynamic subgrid-scale model for compressible turbulence and scalar transport. Phys. Fluids, 3A:2746–2757.
Nieuwstadt, F. T. M., 1984: Some aspects of the turbulent stable boundary-layer. Bound.-Layer Meteor., 30:31–55.
Piomelli, U., P. Moin, and J. H. Ferziger, 1988: Model consistency in large eddy simulation of turbulent channel flows. Phys. Fluids, 31:1884–1891.
Pope, S. B., 2000: Turbulent Flows. Cambridge University Press, 971 pp.
Porté-Agel, F., C. Meneveau, and M. B. Parlange, 1998: Some basic properties of the surrogate subgrid-scale heat flux in the atmospheric boundary layer. Bound.-Layer Meteor., 88:425–444.
Porté-Agel, F., C. Meneveau, and M. B. Parlange, 2000a: A scale-dependent dynamic model for large-eddy simulation: Application to a neutral atmospheric boundary layer. J. Fluid Mech., 415:261–284.
Porté-Agel, F., M. B. Parlange, C. Meneveau, W. Eichinger, and M. Pahlow, 2000b: Subgrid-scale dissipation in the atmospheric surface layer: Effects of stability and filter dimension. J. Hydrometeor., 1:75–87.
Porté-Agel, F., M. Pahlow, C. Meneveau, and M. Parlange, 2001a: Atmospheric stability effect on subgrid scale physics of large-eddy simulation. Adv. Water Res., 24:1085–1102.
Porté-Agel, F., M. B. Parlange, C. Meneveau, and W. E. Eichinger, 2001b: A priori field study of the subgrid-scale heat fluxes and dissipation in the atmospheric surface layer. J. Atmos. Sci., 58:2673–2698.
Smagorinsky, J., 1963: General circulation experiments with the primitive equations. I. The basic experiment. Mon. Wea. Rev., 91:99–164.
Sullivan, P., J. 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.
Sullivan, P., T. W. Horst, D. H. Lenschow, C-H. Moeng, and J. C. Weil, 2003: Structure of subfilter-scale fluxes in the atmospheric surface layer with application to large eddy simulation modeling. J. Fluid Mech., 482:101–139.
Tao, B., J. Katz, and C. Meneveau, 2002: Statistical geometry of subgrid-scale stresses determined from holographic particle image velocimetry measurements. J. Fluid Mech., 457:35–78.
Tong, C., J. C. Wyngaard, S. Khanna, and J. G. Brasseur, 1998: Resolvable- and subgrid-scale measurement in the atmospheric surface layer: Technique and issues. J. Atmos. Sci., 55:3114–3126.
Tong, C., J. C. Wyngaard, and J. G. Brasseur, 1999: Experimental study of the subgrid-scale stresses in the atmospheric surface layer. J. Atmos. Sci., 56:2277–2292.
Zhang, J., B. Tao, and J. Katz, 1997: Turbulent flow measurement in a square duct with hybrid holographic PIV. Exp. Fluids, 23:373–381.
APPENDIX
Test of Filtered Velocity Gradient Accuracy
This divergence parameter η vanishes if the divergence-free condition is obeyed exactly. Moreover, for random data where the individual gradient terms are uncorrelated, η = 1. η is bound by 0 < η < 3. For our data, η varies from one data sample to another and so no unique value of η exists. Instead, as in Zhang et al. (1997) we measure the probability density function (pdf) of η and thus document the frequency of occurrence of different values of η. Parameter η is computed for the four different arrays over the entire dataset and pdfs are plotted in Fig. A1. Clearly η = 0 (satisfaction of continuity) is the most likely value. Between 50% (for array 1) and 65% (for array 4) of the data are between 0 < η < 0.5. Comparing the pdfs with each other one can state that accuracy of gradients decreases with increasing Δ/z. No conclusions can be made about relative errors of x, y, or z gradients, but we expect the largest contribution to the error to be from the first-order one-sided derivatives in the z direction. The level of error in evaluating derivatives apparent from this test can be considered reasonable (although it is not small).
Array properties for the HATS experiment: “d” is double-filtered array; “s” is single-filtered array; d 0 is displacement height; Δ is filter size. The last three columns specify the type of filter used in the x and y directions. The number following the filter type specifies the number of instruments over which the spatial average is computed. Note that, for the remainder of the paper, the data for arrays 3 and 4 are merged, because their z/Δ values are similar
Prandtl number Pr T conditioned on Δ/z computed from Eqs. (7) and (12), assuming that Pr T is not a function of stability. The averaging time is the total time available for each array (Tc > 35 h)