1. Introduction
Atmospheric turbulence is a complex phenomenon, characterized by the presence of a plethora of scales (eddies) ranging from small viscosity-dominated to large, energy-containing structures. Turbulence may undergo space and time variations due to rapidly changing external conditions, it may be locally suppressed or enhanced. To describe characteristic features of turbulence, statistical theories are sought for. In this context, a number of recent research works (Vassilicos 2015; Mazellier and Vassilicos 2008; Bos and Rubinstein 2017; Bos et al. 2007; McComb et al. 2010) address the problem of the equilibrium Taylor law (Taylor 1935) and its failure in the presence of rapid changes of the system. A new, nonclassical, although universal scaling is introduced to describe the latter.
The general validity of the “equilibrium” equations, Eqs. (1) and (2), was coming into question, as laboratory studies by Antonia and Pearson (2000) and Burattini et al. (2005) revealed that Cϵ varies considerably. This behavior was first explained by the dependence on the inflow and boundary conditions. In another work by Bos et al. (2007) it was shown based on theoretical arguments that Cϵ takes different values in forced and decaying turbulence, even though the classical, Kolmogorov −5/3 form of the spectrum was assumed.
Rubinstein and Clark (2017) argued that turbulence characterized by the classical laws (1) and (2) should be called “equilibrium” turbulence. In such type of flow the energy spectrum takes a self-similar form and the classical −5/3 scaling holds (Kolmogorov 1941). Nonequilibrium states of a flow field appear after a sudden change of external conditions (e.g., forcing) when the system evolves toward another equilibrium (Mahrt and Bou-Zeid 2020; Wacławczyk 2021, 2022). In the nonequilibrium turbulence, the turbulence dissipation coefficient Cϵ and the integral-to-Taylor-scale ratios are described by relations (3) and (4) and a nonequilibrium correction should be added to the spectrum. Bos and Rubinstein (2017) derived the form of this correction, as well as the nonequilibrium scaling laws for Cϵ and
The nonequilibrium relations (3) and (4) are substantially different from their equilibrium counterparts in (1) and (2). Hence, classification of turbulence in a given area can be made on this basis. Previous research works which focus on such analysis concern laboratory or numerical experiments (e.g., decaying grid turbulence, turbulent boundary layer) under controlled conditions (Valente and Vassilicos 2015; Obligado et al. 2016; Cafiero and Vassilicos 2019; Obligado et al. 2022). The main idea of this work is to use this procedure to analyze data from in situ airborne measurements of atmospheric turbulence, where the flow conditions in the investigated region are unknown a priori and only limited information, that is 1D intersections of the flow field along the flight track are available. In principle, such analysis could be performed on any high-resolution observations of the atmospheric boundary layer. In this study we emphasize theoretical aspects of the method to be used; hence, we decided to analyze data already known from previous campaigns, but in a way that adds one more important aspect of turbulence. Our choice is the stratocumulus-topped boundary layer (STBL). Several research campaigns and subsequent data analyses were devoted to the study of the turbulence in STBL (see, e.g., Stevens et al. 2003; Malinowski et al. 2013; Siebert et al. 2010; Jen-La Plante et al. 2016). We focus on the data from the Azores Stratocumulus Measurements of Radiation, Turbulence and Aerosols (ACORES) campaign collected in the eastern North Atlantic (Siebert et al. 2021). Nowak et al. (2021) compared the properties of turbulence between the cases of coupled and decoupled STBLs [see Wood et al. (2015) for the definition of decoupling]. They suggested that the observed differences in inertial range scaling might be related to different stages of turbulence lifetime (e.g., development and decay) but did not study these mechanisms in detail. In this work, from the same dataset, we derive all turbulence statistics necessary to calculate Cϵ,
We perform a sensitivity study to show that all statistics, including ϵ, can be estimated from the wind velocity records, even in the case of nonequilibrium, provided that the resolution of the data is good enough and the flight segments are sufficiently long. These conditions are fulfilled by the ACORES data. Moreover, we extend analysis of Bos (2020) and estimate the upper bound of Cϵ for flows with nonzero buoyancy.
The present paper is structured as follows. In the following section 2 formal definitions of the “equilibrium” and “nonequilibrium” turbulence are presented and the theoretical derivations of Bos and Rubinstein (2017) are recalled. In section 3 the investigated data are briefly presented. This is followed by the detailed description of the methods applied in the data analysis. Results are presented in section 4, and finally, conclusions and perspectives are discussed in section 5.
2. Theory of nonequilibrium turbulence
a. Nonequilibrium scaling laws
This section is devoted to the theory presented previously in the works of Bos and Rubinstein (2017) and Rubinstein and Clark (2017). Their derivations are briefly recalled here for the sake of clarity.
Model for spectral energy density in the equilibrium (black dashed line) and in nonequilibrium (solid red line) considered in Bos and Rubinstein (2017).
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
All considerations of Bos and Rubinstein (2017) were performed under the assumptions of homogeneity and local isotropy. These assumptions are usually employed to calculate turbulence dissipation rate from in situ measurements performed by aircrafts, where it is also assumed that the energy spectrum follows the −5/3 law. We also use the assumption of local isotropy in our analyses; however, following Bos and Rubinstein (2017) we account for possible deviations from the Kolmogorov’s scaling due to nonstationarity. The next step, left for further work, would be to account for inhomogeneity and anisotropy. This could be done along the line of the studies of Chen and Vassilicos (2022), where transport equations for the second-order structure functions are considered. We note that the nonequilibrium scaling relations (11) and (12) are also present in the nonhomogeneous flows, like, e.g., turbulent boundary layers or wakes, cf. Obligado et al. (2016, 2022), Nedić et al. (2017); hence, we could assume that the dependence of Cϵ and
b. Estimation of Cϵ from turbulence models
Values of Cϵ/Cϵ0 estimated from Eq. (16).
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
Generally when turbulence becomes locally stronger, Cϵ is smaller than its equilibrium counterpart. The opposite is true when turbulence becomes locally weaker. It is to note that the value Cϵ = 1.8Cϵ0 remains an upper bound also in the presence of negative buoyancy flux (positive Rif), which follows from the fact that (A1 − A3Rif) in Eq. (16) remains positive for the maximal value of Rif = 0.5. It could, however, be possible that some limited regions of zero turbulence production P and negative buoyancy G < 0 would be present. We can expect turbulence would be in a strong nonequilibrium, decaying fast in such zones.
Larger values of Cϵ were obtained, e.g., in the study of Nedić et al. (2017) from direct numerical simulations data of boundary layer flows at relatively low Reynolds number. However, the derivations of Bos and Rubinstein (2017) were performed for simplified high-Re form of the spectrum, cf. Fig. 1. Hence, the results may not be valid at low Re, where the dissipative part of the spectrum is nonnegligible. Low values of Reλ are unlikely to be found in the atmospheric boundary layer turbulence; hence, we would rather expect Cϵ to remain bounded from above by 1.8Cϵ0.
3. Data and methods
To verify, whether the presented above analysis can be applied to atmospheric turbulence, we used high-resolution data collected in July 2017 during the ACORES campaign in the eastern North Atlantic around the island of Graciosa (Siebert et al. 2021). Measurements were performed with the Airborne Cloud Turbulence Observation System (ACTOS) (Siebert et al. 2006) mounted 170 m below the BO-105 helicopter. We analyze the vertical wind velocity w corrected for platform motion and attitude (cf. Nowak et al. 2021) provided by the ultrasonic anemometer–thermometer (Siebert and Muschinski 2001). The sensor was mounted on ACTOS and has the sampling frequency of 100 Hz. The standard deviations due to uncorrelated noise for velocity measurements are 0.002 m s−1. Relatively low velocity of the helicopter (∼20 m s−1) and high frequency of the sensor allows us to reliably resolve small scales of turbulence down to ∼0.5 m. The further analysis is based on the vertical velocity component only, as the horizontal components still showed some influence of platform attitude and motion which could not be completely removed with standard procedures. These issues caused problems with Reynolds decomposition and subsequent analysis. We note that Obligado et al. (2022) compared estimates of Cϵ with
Helicopter flights during ACORES were performed over the ocean inside the 10 km × 10 km square adjacent to the Graciosa island. The typical flight duration was 2 h. The trajectory included vertical profiles up to 2000 m and horizontal legs of the length of several kilometers. In our analysis we use data from the horizontal segments for the same flights as selected in Nowak et al. (2021): flight 5 on 8 July 2017 and flight 14 on 18 July 2017, distinctive by the stratocumulus presence and STBL stratification (considerably well mixed in flight 5, considerably decoupled in flight 14). Detailed analysis of turbulence statistics based on these data was performed in Nowak et al. (2021). The study of nondimensional dissipation coefficient and
a. Modified spectra
Due to finite frequency of ACTOS turbulence sensor, the wind velocity is measured with a spectral cutoff. As mentioned, ACORES data have the spatial resolution of ∼0.5 m, which is still larger than the typical size of the smallest Kolmogorov’s eddies (∼0.001 m to ∼0.01 m). With this, part of the energy spectrum remains unresolved and to recover value of ϵ indirect methods, which rely on the Kolmogorov’s hypotheses, should be used.
In this subsection we show that in spite of the presence of the nonequilibrium correction, the turbulence kinetic energy dissipation rate can still be calculated from the one-dimensional spectra by fitting to the −5/3 slope, provided that the fitting range is moved toward larger wavenumbers. This is possible, because the nonequilibrium correction affects the one-dimensional spectra only slightly and its influence is negligible for large wavenumbers. Hence, we can still estimate ϵ from the measured velocity signals even in the presence of the spectral cutoff at the Nyquist frequency fs/2 = 50 Hz.
One-dimensional spectra calculated for the case P/ϵ = 2, G/ϵ = 0 (developing turbulence with the production twice larger than the dissipation): solid red line; P/ϵ = 0, G/ϵ = 0 (decaying turbulence, no turbulence production): dot–dashed blue line; compared with K41 spectrum: dashed black line.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
One-dimensional transverse structure functions estimated from airborne measurements, from the vertical wind velocity component recorded at height 280 m (region of large turbulence production): solid red line; at height 990 m (region inside the cloud with locally weak turbulence production): dot–dashed blue line; compared with K41 spectrum: dashed black line. Vertical lines mark the fitting range.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
Parameters estimate for different values of P/ϵ and G/ϵ.
b. Characteristic time scales
c. Effects of Reynolds decomposition
To calculate turbulence statistics the instantaneous velocity should first be decomposed into the mean and fluctuating parts. Here, the ensemble average is approximated by the space average along the flight track. The size of averaging window used for detrending, AWD, should be much larger than the characteristic turbulence length scale but much smaller than the length associated with the mean changes. In practice, in the atmospheric turbulence, the presence of large-scale convective motions, internal waves, and changes of atmospheric conditions along the flight track makes the choice of the averaging window difficult. In this work we perform the analysis for vertical velocity components, for which the changes of the mean component were of lesser magnitude. In Nowak et al. (2021), AWD = 50 s, which correspond to, approximately 1 km length, were used for the horizontal segments. Such window was approximately 10 times larger than the estimated integral turbulence length scales which were of order 100 m.
The analysis is performed for the subcloud segment F05LEG307 from flight 5, characterized by low values of Cϵ, which suggests nonequilibrium, developing turbulence, a part of subcloud segment F14LEG287 from flight 14, where turbulence is close to equilibrium and cloud segment F14LEG992 from flight 14 with Cϵ > Cϵ0. For the first two segments the conditions remain approximately constant along the flight track. In the third segment, larger variations of turbulence properties along the flight track are observed; however, the mean value of Cϵ still exceeds the equilibrium value, see section 6. After detrending another averaging window AWS should be chosen for the calculation of statistics. This window should be sufficiently long to reduce the bias and the random errors to acceptable levels (Lenschow et al. 1994). In the following subsection, we estimate the size of AWS sufficient to estimate Cϵ with a good accuracy. Here, we perform averaging over the whole considered parts of the segments, that is, AWS = 270 s for flight 5 and AWS = 300 s for both segments of flight 14. These lengths correspond to at least
Results are presented in Table 2. As can be seen, the larger AWD is chosen, the larger resulting
Estimates of turbulence statistics for different AWD.
d. Effect of averaging windows
Keeping AWD = 50s constant we next investigated influence of the results to the choice of AWS, that is, the window used to calculate statistics from the previously detrended signal. The analysis is performed for the first half of F14LEG287 from flight 14 in the decoupled STBL, where Cϵ remains approximately constant. The largest AWS = 200 s corresponds to, approximately
Cϵ/Cϵ0 calculated for different AWS from part of signal from flight 14, F14LEG287. Horizontal coordinate x denotes the position of the center of the averaging window.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
4. Results: Analysis of turbulence states in STBL
We first address and compare two horizontal flight segments of comparable, low altitudes of 287 m (decoupled STBL, F14LEG287) and 307 m (coupled STBL, F05LEG307). In the former case the averaging window AWS = 150 s was used, in the latter, the window was increased to AWS = 170 s due to larger length scales. The window was moved every 5 s (≈100 m) along the flight track and each time turbulence statistics were calculated. Figure 6a presents values of Cϵ as a function of the position of the center of the averaging window. In the figure two parallel lines of a constant Cϵ0 = 0.45 and Cϵ = 0.8 are plotted. The first one would correspond to the equilibrium, stationary case. We note that this equilibrium value may in fact differ slightly for different flows, as it is weakly influenced by the large turbulence structures and their scaling (Valente and Vassilicos 2012; Bos et al. 2007). This influence was not taken into account in the theoretical considerations in section 2, as it was assumed that the energy density has sharp spectral cutoffs at κL and κη. Moreover, possible flow anisotropy may also play a role and modify the proportionality constant. We choose Cϵ0 = 0.45 as it is the value obtained for F14LEG287 at small AWD = 15 s, see Table 2, for which nonequilibrium effects were mostly filtered out, as it was discussed in section 5d. The second value, Cϵ = 0.8 corresponds to a freely decaying turbulence with P = 0, where the equilibrium, self-similar state was reached (see Table 1).
Cϵ as a function of the position of the center of the averaging window. Red rectangles mark nonequilibrium region. Data for decoupled STBL: (a) F14LEG287, (b) F14LEG448, (c) F14LEG992.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
As it is seen in Fig. 6a, initially, Cϵ ≈ Cϵ0. As the averaging window is moved further along the flight track, a moderate increase of Cϵ is first observed, followed by a sharp decrease toward smaller values. It would suggest that a weak turbulence decay, followed by a region of strong turbulence production was detected. In the analogous figure for the coupled STBL, Fig. 7a, only Cϵ < Cϵ0 are detected, which can be again interpreted as a region of strong turbulence production.
As in Fig. 6, but for coupled STBL: (a) F05LEG307, (b) F05LEG553, (c) F05LEG819.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
We next compared two subcloud flight segments: F14LEG448 in the decoupled STBL of height 448 m and F05LEG553 in the coupled STBL of height 553 m. In Nowak et al. (2021) nearly zero buoyancy production at those levels was reported. The weak turbulence production is manifested through larger average values of Cϵ, both in decoupled and coupled case, cf. Figs. 6b and 7b, respectively. Insufficient turbulence production at this altitude may lead to decoupling, when the eddies fail to mix air over the entire depth of the STBL, and consequently STBL separates into two parts: cloud driven at the top and surface driven at the bottom, cf. Nowak et al. (2021).
Finally, we consider in-cloud segments F14LEG992 in the decoupled STBL and F05LEG819 in the coupled STBL. The signals were recorded at heights 992 and 819 m. In both cases values of Cϵ vary considerably, cf. Figs. 6c and 7c; however, they are somewhat larger in the decoupled case which indicates weaker turbulence production and is again in line with the observations of Nowak et al. (2021). The results of Cϵ can be compared with experimental estimations of Zheng et al. (2021) in high-Re turbulence behind a grid, where Cϵ first decreased in the production region and next increased in the nonequilibrium decay region toward value Cϵ ≈ 0.8. We also mention here the direct numerical simulation study of Gallana et al. (2022), who calculated Cϵ in stably and unstably stratified flows. In the former case Cϵ increased toward Cϵ ≈ 0.8, indicating turbulence suppression and in the unstable conditions decreased toward Cϵ ≈ 0.4, below the equilibrium value. In STBL turbulence is manly driven by radiative and evaporative cooling at the cloud top, leading to instability. However, boundary layer is also capped from above by a stably stratified layer. This may explain large variations of Cϵ along the in-cloud flight tracks.
To classify whether turbulence is in its equilibrium or nonequilibrium state, we calculated the length scales and the local Reynolds numbers and plotted
Cϵ as a function of Reλ. Solid black lines: equilibrium scalings Cϵ0 = 0.45 and Cϵ1 = 1.8Cϵ0; dashed red line: nonequilibrium scaling, Eq. (11); calculated statistics: symbols. Decoupled STBL: (a) F14LEG287, (b) F14LEG448, (c) F14LEG992.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
As in Fig. 8, but for coupled STBL: (a) F05LEG307, (b) F05LEG553, (c) F05LEG819. In (c) color coded are different local equilibrium states.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
Figures 8b and 9b refer to the flight segments F14LEG448 and F05LEG553 below the cloud. Figure 8b suggests that part of the signal in the decoupled STBL is a record of equilibrium turbulence, while statistics of another part fit the nonequilibrium formulas. For the coupled STBL the statistics follow rather the equilibrium predictions, cf. Fig. 9b. A large scatter of results is visible in Fig. 9b such that it is not easy to draw conclusions. Note that the flight segment F05LEG553 is relatively short and so is the range of the detected Reλ values. Equilibrium scaling may suggest that in spite of the weak turbulence production, the changes of atmospheric conditions are relatively slow, such that the turbulence spectra already relaxed to the self-similar form (5). Similar, nearly constant levels of Cϵ were reported, e.g., by Neunaber et al. (2022) in the study of turbulence in wakes behind turbines.
Particularly interesting is the study of the in-cloud segments F14LEG992 and F05LEG819. In the decoupled STBL most of the points follow the nonequilibrium scaling relations, cf. Fig. 8c. Exceptions are several points corresponding to large Cϵ ≈ 1.8Cϵ0, which would indicate regions of no turbulence production and/or stable stratification. In the coupled STBL on the other hand, in spite of large variations of Cϵ, the points seem to follow three different lines of constant Cϵ, namely, Cϵ ≈ 0.8, Cϵ ≈ 0.6, and Cϵ ≈ 0.4, cf. Fig. 9c. We may call such state of the system as a quasi-equilibrium state, which means that the changes of external conditions are slow enough, such that the system undergoes through several local equilibria.
We note here that for certain points the assignment to equilibrium or nonequilibrium region may be somewhat arbitrary. This concerns especially those points which represent statistics calculated at the boundary between the equilibrium and nonequilibrium region. Sometimes this assignment becomes more definite when the dependence of the length scale ratio
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
As in Fig. 10, but for coupled STBL: (a) F05LEG307, (b) F05LEG553, (c) F05LEG819. In (c) color codes are different local equilibrium states.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
Nonequilibrium scaling relations were found in many laboratory studies, as examples we mention here the works of Zheng et al. (2021), which concerns high-Re flows behind a grid, and of Nedić et al. (2017), which concerns boundary layer flows.
As in Fig. 6, but for the decoupled STBL, F14LEG143.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
(a) Cϵ as a function of Reλ, (b)
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
As in Fig. 13, but for rescaled Reλ.
Citation: Journal of the Atmospheric Sciences 79, 10; 10.1175/JAS-D-22-0028.1
5. Conclusions
The main purpose of this work was to show that certain information about the temporal variations of turbulence and the budget of the turbulence kinetic energy in the STBL can be acquired by analyzing statistics of wind velocity based on in situ measurement data. The key indicators are the nondimensional dissipation coefficient Cϵ and the integral-to-Taylor-scale ratio
Moreover, as it was discussed in the previous works (Bos and Rubinstein 2017) two different states of turbulence system can be identified: equilibrium, where the turbulence kinetic energy spectrum has a self-similar shape, (5), and a nonequilibrium where a correction to the spectrum should be considered; see Eq. (6). The equilibrium state can possibly be nonstationary; however, the changes of external conditions (forcing) are slow enough such that the system has enough time to relax and the energy spectrum takes the self-similar form (5). To detect the nonequilibrium states we studied dependence of Cϵ and
In this work we focus on the analysis of airborne measurement data, which often suffer from various deficiencies, related to finite frequencies of the sensors or short available averaging windows (Wacławczyk et al. 2017; Akinlabi et al. 2019). We showed that despite these deficiencies both Cϵ and
The analysis performed here allows for additional comparison of turbulence in the coupled and decoupled STBL’s, previously studied by Nowak et al. (2021). In both cases, Cϵ is smaller at lower altitudes due to strong shear and/or buoyancy production. On the other hand, in the subcloud parts Cϵ increases above Cϵ0 which suggests that in these regions turbulence decays. Large variations of Cϵ were detected inside the cloud in both cases. Apparently, along the flight track, regions where turbulence production surpass or is in balance with the dissipation alternate with areas of turbulence decay. Here, the main difference between the coupled and decoupled STBL was observed. Namely, in the former, turbulence seems to undergo a series of equilibrium states, where Cϵ ≈ const and
In this work only limited amount of data from two ACTOS flights were investigated. In future, more thorough analysis to better characterize turbulence in the coupled and decoupled STBL is foreseen. The same analysis can be repeated for various measurement data in the atmospheric boundary layers where the impact of transient states is nonnegligible (cf. Mahrt and Bou-Zeid 2020), e.g., to study the collapse of turbulent convective daytime boundary layer (El Guernaoui et al. 2019; Wang et al. 2020; Lothon et al. 2014), in and between various types of the clouds in order to better understand their dynamics and even in clear air turbulence to characterize its dynamic properties.
Another interesting aspect is the relation of the present study to the K62 theory (Kolmogorov 1962), which takes into account temporal and spatial variations of the instantaneous dissipation rate. This phenomenon is known as the “internal intermittency” and affects the scaling of higher-order structure functions. A careful analysis of the fine-scale structure of turbulence performed by Siebert et al. (2010) based on ACTOS data revealed the intermittency coefficient in cloud turbulence is consistent with values obtained in laboratory experiments. A question to be asked is how the nonequilibrium structure functions are affected by the small-scale intermittency and whether temporal variations of statistics considered here can be linked to the premises of the K62 theory.
The presence of nonequilibrium in the stratocumulus clouds indicate that the common turbulence closures like the Smagorinsky LES model may fail to predict the dynamics of such clouds correctly. In particular, this may influence the prediction of boundary layer decoupling or cloud dissipation. A nonequilibrium turbulence model may largely improve these predictions. This is another direction of future studies.
Acknowledgments.
This research was supported by the National Science Centre, Poland, Project 2020/37/B/ST10/03695 (theory and methodology), and by the European Union’s Horizon 2020 research and innovation programme under Grant Agreement 101003470, project Next Generation Earth Modelling Systems (NextGEMS) (analyses of measurement data of turbulence in the stratocumulus-topped boundary layer). The field campaign was supported by the Deutsche Forschungsgesellschaft (DFG; Grant SI 1543/4-1).
Data availability statement.
All data from the ACORES study are available from the authors on request.
APPENDIX A
Derivation of Nonequilibrium Scaling Laws
APPENDIX B
Transport Equations for the Turbulence Kinetic Energy and Dissipation Rate
APPENDIX C
Estimation of the Integral Length Scale from Measurement Data
REFERENCES
Akinlabi, E. O., M. Wacławczyk, J. P. Mellado, and S. P. Malinowski, 2019: Estimating turbulence kinetic energy dissipation rates in the numerically simulated stratocumulus cloud-top mixing layer: Evaluation of different methods. J. Atmos. Sci., 76, 1471–1488, https://doi.org/10.1175/JAS-D-18-0146.1.
Antonia, R. A., and B. R. Pearson, 2000: Effect of initial conditions on the mean energy dissipation rate and the scaling exponent. Phys. Rev. E, 62, 8086, https://doi.org/10.1103/PhysRevE.62.8086.
Baum, E., and E. A. Caponi, 1992: Modeling the effects of buoyancy on the evolution of geophysical boundary layers. J. Geophys. Res., 97, 15 513–15 527, https://doi.org/10.1029/92JC01715.
Bos, W. J. T., 2020: Production and dissipation of kinetic energy in grid turbulence. Phys. Rev. Fluids, 5, 104607, https://doi.org/10.1103/PhysRevFluids.5.104607.
Bos, W. J. T., and R. Rubinstein, 2017: Dissipation in unsteady turbulence. Phys. Rev. Fluids, 2, 022601, https://doi.org/10.1103/PhysRevFluids.2.022601.
Bos, W. J. T., L. Shao, and J.-P. Bertoglio, 2007: Spectral imbalance and the normalized dissipation rate of turbulence. Phys. Fluids, 19, 045101, https://doi.org/10.1063/1.2714079.
Burattini, P., P. Lavoie, and R. A. Antonia, 2005: On the normalized turbulent energy dissipation rate. Phys. Fluids, 17, 098103, https://doi.org/10.1063/1.2055529.
Cafiero, G., and J. C. Vassilicos, 2019: Non-equilibrium turbulence scalings and self-similarity in turbulent planar jet. Proc. Roy. Soc. London, A475, 20190038, https://doi.org/10.1098/rspa.2019.0038.
Chen, J. G., and J. C. Vassilicos, 2022: Scalings of scale-by-scale turbulence energy in non-homogeneous turbulence. J. Fluid Mech., 938, A7, https://doi.org/10.1017/jfm.2022.153.
El Guernaoui, O., J. Reuder, I. Esau, T. Wolf, and B. Maronga, 2019: Scaling the decay of turbulence kinetic energy in the free-convective boundary layer. Bound.-Layer Meteor., 173, 79–97, https://doi.org/10.1007/s10546-019-00458-z.
Gallana, L., S. Abdunabiev, M. Golshan, and D. Tordella, 2022: Diffusion of turbulence following both stable and unstable step stratification perturbations. Phys. Fluids, 34, 065122, https://doi.org/10.1063/5.0090042.
Galperin, B., S. Sukoriansky, and P. S. Anderson, 2007: On the critical Richardson number in stably stratified turbulence. Atmos. Sci. Lett., 8, 65–69, https://doi.org/10.1002/asl.153.
Grachev, A. A., E. L. Andreas, C. W. Fairall, P. S. Guest, and P. O. G. Persson, 2013: The critical Richardson number and limits of applicability of local similarity theory in the stable boundary layer. Bound.-Layer Meteor., 147, 51–82, https://doi.org/10.1007/s10546-012-9771-0.
Jen-La Plante, I., and Coauthors, 2016: Physics of Stratocumulus Top (POST): Turbulence characteristics. Atmos. Chem. Phys., 16, 9711–9725, https://doi.org/10.5194/acp-16-9711-2016.
Jones, W. P., and B. E. Launder, 1972: The prediction of laminarization with a two-equation model of turbulence. Int. J. Heat Mass Transfer, 15, 301–314, https://doi.org/10.1016/0017-9310(72)90076-2.
Kolmogorov, A. N., 1941: Dissipation of energy in the locally isotropic turbulence. Proc. Roy. Soc. London, A434, 15–17, https://doi.org/10.1098/rspa.1991.0076.
Kolmogorov, A. N., 1962: A refinement of previous hypotheses concerning the local structure of turbulence in a viscous incompressible fluid at high Reynolds number. J. Fluid Mech., 13, 82–85, https://doi.org/10.1017/S0022112062000518.
Lenschow, D. H., J. Mann, and L. Kristensen, 1994: How long is long enough when measuring fluxes and other turbulence statistics? J. Atmos. Oceanic Technol., 11, 661–673, https://doi.org/10.1175/1520-0426(1994)011<0661:HLILEW>2.0.CO;2.
Lothon, M., and Coauthors, 2014: The BLLAST field experiment: Boundary-layer late afternoon and sunset turbulence. Atmos. Chem. Phys., 14, 10 931–10 960, https://doi.org/10.5194/acp-14-10931-2014.
Mahrt, L., and E. Bou-Zeid, 2020: Non-stationary boundary layers. Bound.-Layer Meteor., 177, 189–204, https://doi.org/10.1007/s10546-020-00533-w.
Malinowski, S. P., and Coauthors, 2013: Physics of Stratocumulus Top (POST): Turbulent mixing across capping inversion. Atmos. Chem. Phys., 13, 12 171–12 186, https://doi.org/10.5194/acp-13-12171-2013.
Mazellier, N., and J. C. Vassilicos, 2008: The turbulence dissipation constant is not universal because of its universal dependence on large-scale flow topology. Phys. Fluids, 20, 015101, https://doi.org/10.1063/1.2832778.
McComb, W. D., A. Berera, M. Salewski, and S. Yoffe, 2010: Taylor’s (1935) dissipation surrogate reinterpreted. Phys. Fluids, 22, 061704, https://doi.org/10.1063/1.3450299.
Nedić, J., S. Tavoularis, and I. Marusic, 2017: Dissipation scaling in constant-pressure turbulent boundary layers. Phys. Rev. Fluids, 2, 032601, https://doi.org/10.1103/PhysRevFluids.2.032601.
Neunaber, I., J. Peinke, and M. Obligado, 2022: Application of the Townsend–George theory for free shear flows to single and double wind turbine wakes—A wind tunnel study. Wind Energy Sci., 7, 201–219, https://doi.org/10.5194/wes-7-201-2022.
Nowak, J. L., H. Siebert, K.-E. Szodry, and S. P. Malinowski, 2021: Coupled and decoupled stratocumulus-topped boundary layers: Turbulence properties. Atmos. Chem. Phys., 21, 10 965–10 991, https://doi.org/10.5194/acp-21-10965-2021.
Obligado, M., T. Dairay, and J. C. Vassilicos, 2016: Nonequilibrium scalings of turbulent wakes. Phys. Rev. Fluids, 1, 044409, https://doi.org/10.1103/PhysRevFluids.1.044409.
Obligado, M., C. Brun, J. H. Silvestrini, and E. B. C. Schettini, 2022: Dissipation scalings in the turbulent boundary layer at moderate Reθ. Flow Turbul. Combust., 108, 105–122, https://doi.org/10.1007/s10494-021-00270-1.
Pope, S. B., 2000: Turbulent Flows. 1st ed. Cambridge University Press, 802 pp.
Rodi, W., 1987: Turbulence Models and Their Application in Hydraulics: A State of the Art Review. IAHR, 104 pp.
Rubinstein, R., and T. T. Clark, 2017: “Equilibrium” and “non-equilibrium” turbulence. Theor. Appl. Mech. Lett., 7, 301–305, https://doi.org/10.1016/j.taml.2017.09.010.
Seoud, R. E., and J. C. Vassilicos, 2007: Dissipation and decay of fractal-generated turbulence. Phys. Fluids, 19, 105108, https://doi.org/10.1063/1.2795211.
Siebert, H., and A. Muschinski, 2001: Relevance of a tuning-fork effect for temperature measurements with the Gill Solent HS ultrasonic anemometer–thermometer. J. Atmos. Oceanic Technol., 18, 1367–1376, https://doi.org/10.1175/1520-0426(2001)018<1367:ROATFE>2.0.CO;2.
Siebert, H., K. Lehmann, and M. Wendisch, 2006: Observations of small-scale turbulence and energy dissipation rates in the cloudy boundary layer. J. Atmos. Sci., 63, 1451–1466, https://doi.org/10.1175/JAS3687.1.
Siebert, H., R. A. Shaw, and Z. Warhaft, 2010: Statistics of small-scale velocity fluctuations and internal intermittency in marine stratocumulus clouds. J. Atmos. Sci., 67, 262–273, https://doi.org/10.1175/2009JAS3200.1.
Siebert, H., and Coauthors, 2021: Observations of aerosol, cloud, turbulence, and radiation properties at the top of the marine boundary layer over the eastern North Atlantic Ocean: The ACORES campaign. Bull. Amer. Meteor. Soc., 102, E123–E147, https://doi.org/10.1175/BAMS-D-19-0191.1.
Stevens, B., and Coauthors, 2003: Dynamics and Chemistry of Marine Stratocumulus—DYCOMS-II. Bull. Amer. Meteor. Soc., 84, 579–594, https://doi.org/10.1175/BAMS-84-5-579.
Taylor, G. I., 1935: Statistical theory of turbulence. Proc. Roy. Soc. London, A151, 421–444, https://doi.org/10.1098/rspa.1935.0158.
Valente, P. C., and J. C. Vassilicos, 2011: The decay of turbulence generated by a class of multiscale grids. J. Fluid Mech., 687, 300–340, https://doi.org/10.1017/jfm.2011.353.
Valente, P. C., and J. C. Vassilicos, 2012: Universal dissipation scaling for nonequilibrium turbulence. Phys. Rev. Lett., 108, 214503, https://doi.org/10.1103/PhysRevLett.108.214503.
Valente, P. C., and J. C. Vassilicos, 2015: The energy cascade in grid-generated non-equilibrium decaying turbulence. Phys. Fluids, 27, 045103, https://doi.org/10.1063/1.4916628.
Vassilicos, J. C., 2015: Dissipation in turbulent flows. Annu. Rev. Fluid Mech., 47, 95–114, https://doi.org/10.1146/annurev-fluid-010814-014637.
Wacławczyk, M., Y.-F. Ma, J. M. Kopeć, and S. P. Malinowski, 2017: Novel approaches to estimating the turbulent kinetic energy dissipation rate from low- and moderate-resolution velocity fluctuation time series. Atmos. Meas. Tech., 10, 4573–4585, https://doi.org/10.5194/amt-10-4573-2017.
Wacławczyk, M., A. S. Gozingan, J. Nzotungishaka, M. Mohammadi, and S. P. Malinowski, 2020: Comparison of different techniques to calculate properties of atmospheric turbulence from low-resolution data. Atmosphere, 11, 199, https://doi.org/10.3390/atmos11020199.
Wacławczyk, T., 2021: Modeling of non-equilibrium effects in intermittency region between two phases. Int. J. Multiphase Flow, 134, 103459, https://doi.org/10.1016/j.ijmultiphaseflow.2020.103459.
Wacławczyk, T., 2022: On differences between deterministic and statistical models of the interphase region. Acta Mech. Sin., 38, 722045, https://doi.org/10.1007/s10409-022-22045-w.
Wang, D., I. S. Stachlewska, X. Song, B. Heese, and A. Nemuc, 2020: Variability of the boundary layer over an urban continental site based on 10 years of active remote sensing observations in Warsaw. Remote Sens., 12, 340, https://doi.org/10.3390/rs12020340.
Wood, R., and Coauthors, 2015: Clouds, aerosols, and precipitation in the marine boundary layer: An ARM Mobile Facility deployment. Bull. Amer. Meteor. Soc., 96, 419–440, https://doi.org/10.1175/BAMS-D-13-00180.1.
Yang, P.-F., A. Pumir, and H. Xu, 2018: Generalized self-similar spectrum and the effect of large-scale in decaying homogeneous isotropic turbulence. New J. Phys., 20, 103035, https://doi.org/10.1088/1367-2630/aae72d.
Zheng, S., P. J. K. Bruce, C. Cuvier, J.-M. Foucaut, J. M. R. Graham, and J. C. Vassilicos, 2021: Nonequilibrium dissipation scaling in large Reynolds number turbulence generated by rectangular fractal grids. Phys. Rev. Fluids, 6, 054613, https://doi.org/10.1103/PhysRevFluids.6.054613.
Zilitinkevich, S., and A. Baklanov, 2002: Calculation of the height of the stable boundary layer in practical applications. Bound.-Layer Meteor., 105, 389–409, https://doi.org/10.1023/A:1020376832738.