1. Introduction












In the atmosphere, turbulence is much more complex than the homogeneous and isotropic one because of the influence of density stratification and rotation (Holton and Hakim 2012). Turbulence in such condition is known to develop a complex dynamics, with power-law energy spectra, as revealed by accurate numerical simulations and laboratory experiments (Levich and Tzvetkov 1985; Schertzer et al. 1997; Falkovich 1992; Pouquet and Marino 2013). Depending on the scale of the flow, energy transfers can be directed either toward smaller scales (direct cascade) or toward larger scales (inverse cascade; Bartello 1995). To date, there is in fact no general consensus about the direction of cascades in the atmosphere. Observed energy spectra in the troposphere and in the lower stratosphere (Nastrom and Gage 1985) exhibit a
A way to clarify the situation is to compute directly the energy fluxes. In the classical picture of turbulence, such energy transfers are related directly to the skewness of velocity increments







This is now possible through an important breakthrough made by Duchon and Robert (2000), who reformulated the energy budget of the Navier–Stokes equations into a form allowing for the definition of energy transfers local in space and time and valid for any geometry, including when strong inhomogeneity and anisotropy are present. Its ability to provide interesting information about energy transfers at a given scale
The paper is structured as follows. After presenting the indicator, we will study these transfers in the National Centers for Environmental Prediction–National Center for Atmospheric Research (NCEP–NCAR) reanalysis and ERA-Interim—to study the sensitivity of the results to the resolution—for the year 2005. This year is ideal, as it does not correspond to major ENSO events or volcanic eruptions. We investigate (i) the vertical and horizontal global averages and (ii) the distribution of energy transfers at different scales. Results are displayed in arbitrary units, but in the same scale, for the NCEP–NCAR reanalysis and ERA-Interim. Finally, we discuss the implications of our results on a theoretical and practical level.
2. Methods











where G is a smooth filtering function, nonnegative, spatially localized and such that

As noticed by Duchon and Robert, the average of
By construction, the intrinsic weak formulation of













The sign and geometry of the zones associated with high and low values of
3. Analysis
For this study, outputs of ERA-Interim and NCEP–NCAR Reanalysis-1 have been used. ERA-Interim is a modern-generation reanalysis with a much-higher-resolution model. NCEP–NCAR Reanalysis-1 was pioneering when it was developed but is run at a comparatively low resolution and does not take advantage of as many observations as ERA-Interim.
ERA-Interim is the currently operational reanalysis product at the European Centre for Medium-Range Weather Forecasts (ECMWF; Dee et al. 2011). Released in 2007, it provides reanalyzed data from 1979 to the present, stored at an original T255 spectral resolution (about 80-km horizontal resolution), with 60 vertical hybrid model levels. A 12-h four-dimensional variational (4D-Var) data assimilation is adopted. As a forecast model, the Integrated Forecast System (IFS), cy31r2 release, is used, fully coupling modules for the atmosphere, ocean waves, and land surface. Sea surface temperatures (SSTs) and sea ice concentration (SIC) are ingested as boundary conditions and interpolated on a reduced Gaussian grid as needed. In our case, zonal, meridional, and vertical wind components are considered at a 0.75° × 0.75° horizontal resolution over 12 pressure levels between 1000 and 100 hPa. A 12-h time step is considered. Known problems concerning these datasets are the lack of dry mass conservation (Berrisford et al. 2011) and the slight asymmetry between evaporation and precipitation (Dee et al. 2011). The turbulent fluxes are based on the tiled ECMWF scheme for surface exchanges over land (Viterbo and Beljaars 1995; Viterbo and Betts 1999). Each grid box is divided into up to six fractions (over land) depending on the type of surface, having different transfer coefficients based on a Monin–Obukhov formulation. Similarly, over oceans, two different coefficients are used for stable and unstable conditions (Beljaars 1995).
NCEP–NCAR Reanalysis-1 has been developed in a joint effort by NCEP and NCAR (Kalnay et al. 1996). The simulation has been operational since January 1995, covering a period from 1948 to the present. Data assimilation is performed via a 3D variational scheme (Parrish and Derber 1992). It features a T62 spectral resolution, corresponding to a 2.5° × 2.5° horizontal grid (about 200-km horizontal resolution), with 28 sigma levels. Most of the major physical processes involving the climate system are parameterized. SST, SIC, snow cover, albedo, soil wetness, and roughness length are ingested as boundary conditions. Data are archived at an original 6-h time step, and such a temporal resolution is retained for our analysis. The atmospheric model that provides the NCEP–NCAR reanalysis data uses bulk aerodynamic formulas to estimate the turbulent fluxes, with exchange coefficients depending on empirical profiles extending the Monin–Obukhov similarity relationship (Miyakoda and Sirutis 1986). For more details on the comparison between different subgrid parameterization of surface fluxes, refer to Brunke et al. (2011).
Analysis of local energy transfers
1) One year’s and seasons’ averages of local energy transfers
We begin the analysis by studying the latitudinal averages and the spatial features of the DR indicator for both ERA-Interim and the coarser NCEP–NCAR reanalysis. To enable comparison between the two datasets, one has to choose the analysis length larger than the resolution scale of NCEP–NCAR (200 km) since going below the resolution size introduces spurious effects dependent on the filter design. On the other hand, since we want to obtain as much detail as possible, we have to choose the smallest scale consistent with those requirements. Here, we thus adopt a scale of
Results obtained for both reanalyses are consistent with each other, as can be checked from Fig. 1 (ERA-Interim) and Fig. 2 (NCEP–NCAR). The gross features do not depend on whether one undertakes an annual average (Figs. 1a,d and 2a,d) or seasonal average (Figs. 1b,c,e,f and 2b,c,e,f): in Figs. 1a–c and 2a–c, which show height dependence of the longitudinally averaged
Distribution of
Citation: Journal of the Atmospheric Sciences 75, 7; 10.1175/JAS-D-17-0114.1
As in Fig. 1, but for NCEP–NCAR reanalysis.
Citation: Journal of the Atmospheric Sciences 75, 7; 10.1175/JAS-D-17-0114.1
In the middle troposphere (P = 500 hPa), the behavior of
2) Correlation with energy spectrum
The above result shows that the kinetic energy flux is globally positive in the troposphere, indicating a direct kinetic energy cascade, while they are negative in the lower stratosphere, indicating an inverse kinetic energy cascade. Our results are therefore consistent with those found by Peng et al. (2015), who also found upscale transfer in the lower stratosphere at outer mesoscale length scales and downscale transfers at scales smaller than 360 (KE) or 200 km (APE).
To get some insight on these cascades, we have further computed the kinetic horizontal energy spectra where k is the inverse of the wavelength from the horizontal velocity fields at different pressure levels in the two reanalyses. They are reported in Fig. 3. One sees that for P ≤ 500 hPa (corresponding to the stratosphere), the energy spectrum is mostly scaling like
(a) NCEP–NCAR reanalysis and (b) ERA-Interim. Spectra
Citation: Journal of the Atmospheric Sciences 75, 7; 10.1175/JAS-D-17-0114.1
3) Probability distribution functions of instantaneous local energy transfers
In addition to time average, it is also interesting to study the probability distribution function of instantaneous local energy transfers,
Empirical
Citation: Journal of the Atmospheric Sciences 75, 7; 10.1175/JAS-D-17-0114.1
As in Fig. 4, but for NCEP–NCAR reanalysis.
Citation: Journal of the Atmospheric Sciences 75, 7; 10.1175/JAS-D-17-0114.1
Overall, all distributions are skewed and exhibit fat tails. The sign of the skewness depends on the height: for both the total and kinetic component, it is positive in the lower troposphere and negative for P < 500 hPa, in agreement with the time averages. For the thermodynamic part, the behavior is opposite, with a negative skewness at low altitude (P > 700 hPa) and positive skewness at large altitude. In such case, the distribution is totally asymmetric and includes only positive transfer, indicating that in the higher part of the atmosphere, the density fluctuations only contribute to a downscale energy transfer. Although there is agreement between the ERA-Interim and NCEP–NCAR data, the latter show fatter tails. This might be due to either the different resolution of the datasets and/or the different physical parameterizations.
Looking now at the dependence with scale at a fixed height, we see that both the kinetic and total local energy transfer display similar behavior, with a tendency to have fatter tails with decreasing scales. This means that the energy imbalance of the reanalysis is reduced when we look at motions whose characteristic scales are larger. This type of behavior, also observed in local energy transfers measured in a laboratory turbulent von Kármán flow (Saw et al. 2016), might be because, at larger scales, the atmosphere becomes more wavelike (and so less turbulent; Rhines 1979). Regarding the thermodynamical part of the transfer, the scale dependence is much milder on the positive side of the distribution and even absent in the negative part of the distribution.
4) Possible interpretation
A possible way to explain the sign of the DR indicators is to invoke the relation between baroclinic and barotropic flows and direct and inverse cascades. In Tung and Orlando (2003), it is argued that the baroclinic motions responsible for the genesis and decay of extratropical cyclones are mostly associated with direct cascades [corresponding to positive
4. Discussion
Weather and climate models do not resolve the viscous scales, which for the atmospheric motions are of order 0.1 mm (Priestley 1959). To date, their resolution ranges from ≃2 km for the regional weather models to ≃100 km for the global climate models. To correctly represent dissipation effects at a scale
The quantity
It is evident that the resolution plays an important role in determining spurious energy fluxes by looking at the difference in the
Several wave phenomena in the atmosphere—gravity and Rossby waves—are tied to these horizontal density variations and are associated with energy conversion between available potential and kinetic energy. One key question is whether this diagnostic may incorrectly assess such energy conversion as an energy transfer across scales. For future research directions, it might be worth applying the diagnostic to a simple gravity or Rossby wave model.
Acknowledgments
D. Faranda was supported by ERC Grant 338965. V. Lembo was funded by the Collaborative Research Centre TRR181 “Energy Transfers in Atmosphere and Ocean” funded by the German Research Foundation (DFG). D. F. thanks G. Messori and N. Vercauteren for useful discussions and comments on the paper.
APPENDIX
Derivation of the Local Duchon–Robert Equation for Boussinesq Equations

















REFERENCES
Augier, P., and E. Lindborg, 2013: A new formulation of the spectral energy budget of the atmosphere, with application to two high-resolution general circulation models. J. Atmos. Sci., 70, 2293–2308, https://doi.org/10.1175/JAS-D-12-0281.1.
Augier, P., S. Galtier, and P. Billant, 2012: Kolmogorov laws for stratified turbulence. J. Fluid Mech., 709, 659–670, https://doi.org/10.1017/jfm.2012.379.
Bartello, P., 1995: Geostrophic adjustment and inverse cascades in rotating stratified turbulence. J. Atmos. Sci., 52, 4410–4428, https://doi.org/10.1175/1520-0469(1995)052<4410:GAAICI>2.0.CO;2.
Beljaars, A., 1995: The parametrization of surface fluxes in large-scale models under free convection. Quart. J. Roy. Meteor. Soc., 121, 255–270, https://doi.org/10.1002/qj.49712152203.
Berrisford, P., P. Kållberg, S. Kobayashi, D. Dee, S. Uppala, J. Simmons, P. Poli, and H. Sato, 2011: Atmospheric conservation properties in ERA-Interim. Quart. J. Roy. Meteor. Soc., 137, 1381–1399, https://doi.org/10.1002/qj.864.
Bou-Zeid, E., C. Meneveau, and M. B. Parlange, 2004: Large-eddy simulation of neutral atmospheric boundary layer flow over heterogeneous surfaces: Blending height and effective surface roughness. Water Resour. Res., 40, W02505, https://doi.org/10.1029/2003WR002475.
Brunke, M. A., Z. Wang, X. Zeng, M. Bosilovich, and C.-L. Shie, 2011: An assessment of the uncertainties in ocean surface turbulent fluxes in 11 reanalysis, satellite-derived, and combined global datasets. J. Climate, 24, 5469–5493, https://doi.org/10.1175/2011JCLI4223.1.
Burgess, B. H., A. R. Erler, and T. G. Shepherd, 2013: The troposphere-to-stratosphere transition in kinetic energy spectra and nonlinear spectral fluxes as seen in ECMWF analyses. J. Atmos. Sci., 70, 669–687, https://doi.org/10.1175/JAS-D-12-0129.1.
Campagne, A., 2015: Cascades dénergie et turbulence dondes dans une expérience de turbulence en rotation. Ph.D. thesis, Université Paris Sud, 129 pp.
Campagne, A., B. Gallet, F. Moisy, and P.-P. Cortet, 2014: Direct and inverse energy cascades in a forced rotating turbulence experiment. Phys. Fluids, 26, 125112, https://doi.org/10.1063/1.4904957.
Charney, J. G., 1971: Geostrophic turbulence. J. Atmos. Sci., 28, 1087–1095, https://doi.org/10.1175/1520-0469(1971)028<1087:GT>2.0.CO;2.
Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553–597, https://doi.org/10.1002/qj.828.
Duchon, J., and R. Robert, 2000: Inertial energy dissipation for weak solutions of incompressible Euler and Navier-Stokes equations. Nonlinearity, 13, 249–255, https://doi.org/10.1088/0951-7715/13/1/312.
Falkovich, G., 1992: Inverse cascade and wave condensate in mesoscale atmospheric turbulence. Phys. Rev. Lett., 69, 3173–3176, https://doi.org/10.1103/PhysRevLett.69.3173.
Frisch, U., 1995: Turbulence: The Legacy of A. N. Kolmogorov. Cambridge University Press, 296 pp.
Gage, K., 1979: Evidence for a k−5/3 law inertial range in mesoscale two-dimensional turbulence. J. Atmos. Sci., 36, 1950–1954, https://doi.org/10.1175/1520-0469(1979)036<1950:EFALIR>2.0.CO;2.
Higgins, C. W., 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., 109, 59–83, https://doi.org/10.1023/A:1025484500899.
Holton, J. R., and G. J. Hakim, 2012: An Introduction to Dynamic Meteorology. 5th ed. International Geophysics Series, Vol. 88, Academic Press, 552 pp.
Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77, 437–471, https://doi.org/10.1175/1520-0477(1996)077<0437:TNYRP>2.0.CO;2.
Kitamura, Y., and Y. Matsuda, 2006: The kH −3 and kH −5/3 energy spectra in stratified turbulence. Geophys. Res. Lett., 33, L05809, https://doi.org/10.1029/2005GL024996.
Klein, R., N. Botta, T. Schneider, C.-D. Munz, S. Roller, A. Meister, L. Hoffmann, and T. Sonar, 2001: Asymptotic adaptive methods for multi-scale problems in fluid mechanics. Practical Asymptotics, Springer, 261–343.
Kolmogorov, A. N., 1941: Dissipation of energy in locally isotropic turbulence. Dokl. Akad. Nauk SSSR, 32, 16–18.
Koshyk, J. N., B. A. Boville, K. Hamilton, E. Manzini, and K. Shibata, 1999: Kinetic energy spectrum of horizontal motions in middle-atmosphere models. J. Geophys. Res., 104, 27 177–27 190, https://doi.org/10.1029/1999JD900814.
Kuzzay, D., D. Faranda, and B. Dubrulle, 2015: Global vs local energy dissipation: The energy cycle of the turbulent von Kármán flow. Phys. Fluids, 27, 075105, https://doi.org/10.1063/1.4923750.
Levich, E., and E. Tzvetkov, 1985: Helical inverse cascade in three-dimensional turbulence as a fundamental dominant mechanism in mesoscale atmospheric phenomena. Phys. Rep., 128, 1–37, https://doi.org/10.1016/0370-1573(85)90036-5.
Lilly, D. K., 1983: Stratified turbulence and the mesoscale variability of the atmosphere. J. Atmos. Sci., 40, 749–761, https://doi.org/10.1175/1520-0469(1983)040<0749:STATMV>2.0.CO;2.
Lindborg, E., 2005: The effect of rotation on the mesoscale energy cascade in the free atmosphere. Geophys. Res. Lett., 32, L01809, https://doi.org/10.1029/2004GL021319.
Lindborg, E., 2006: The energy cascade in a strongly stratified fluid. J. Fluid Mech., 550, 207–242, https://doi.org/10.1017/S0022112005008128.
Lindborg, E., and J. Y. N. Cho, 2001: Horizontal velocity structure functions in the upper troposphere and lower stratosphere: 2. Theoretical considerations. J. Geophys. Res., 106, 10 233–10 241, https://doi.org/10.1029/2000JD900815.
Lorenz, E. N., 1955: Available potential energy and the maintenance of the general circulation. Tellus, 7, 157–167, https://doi.org/10.3402/tellusa.v7i2.8796.
Lovejoy, S., and D. Schertzer, 2010: Towards a new synthesis for atmospheric dynamics: Space–time cascades. Atmos. Res., 96, 1–52, https://doi.org/10.1016/j.atmosres.2010.01.004.
Lübken, F.-J., 1997: Seasonal variation of turbulent energy dissipation rates at high latitudes as determined by in situ measurements of neutral density fluctuations. J. Geophys. Res., 102, 13 441–13 456, https://doi.org/10.1029/97JD00853.
Lucarini, V., and F. Ragone, 2011: Energetics of climate models: Net energy balance and meridional enthalpy transport. Rev. Geophys., 49, RG1001, https://doi.org/10.1029/2009RG000323.
Miller, L. M., F. Gans, and A. Kleidon, 2011: Estimating maximum global land surface wind power extractability and associated climatic consequences. Earth Syst. Dyn., 2, 1–12, https://doi.org/10.5194/esd-2-1-2011.
Miller, L. M., N. A. Brunsell, D. B. Mechem, F. Gans, A. J. Monaghan, R. Vautard, D. W. Keith, and A. Kleidon, 2015: Two methods for estimating limits to large-scale wind power generation. Proc. Natl. Acad. Sci. USA, 112, 11 169–11 174, https://doi.org/10.1073/pnas.1408251112.
Miyakoda, K., and J. Sirutis, 1986: Manual of the e-physics. Geophysical Fluid Dynamics Laboratory Rep., 122 pp.
Nastrom, G., and K. S. Gage, 1985: A climatology of atmospheric wavenumber spectra of wind and temperature observed by commercial aircraft. J. Atmos. Sci., 42, 950–960, https://doi.org/10.1175/1520-0469(1985)042<0950:ACOAWS>2.0.CO;2.
Parrish, D. F., and J. C. Derber, 1992: The National Meteorological Center’s spectral statistical-interpolation analysis system. Mon. Wea. Rev., 120, 1747–1763, https://doi.org/10.1175/1520-0493(1992)120<1747:TNMCSS>2.0.CO;2.
Peng, J., L. Zhang, and J. Guan, 2015: Applications of a moist nonhydrostatic formulation of the spectral energy budget to baroclinic waves. Part I: The lower-stratospheric energy spectra. J. Atmos. Sci., 72, 2090–2108, https://doi.org/10.1175/JAS-D-14-0306.1.
Pouquet, A., and R. Marino, 2013: Geophysical turbulence and the duality of the energy flow across scales. Phys. Rev. Lett., 111, 234501, https://doi.org/10.1103/PhysRevLett.111.234501.
Pouquet, A., R. Marino, P. D. Mininni, and D. Rosenberg, 2017: Dual constant-flux energy cascades to both large scales and small scales. Phys. Fluids, 29, 111108, https://doi.org/10.1063/1.5000730.
Priestley, C. H. B., 1959: Turbulent Transfer in the Lower Atmosphere. University of Chicago Press, 130 pp.
Rhines, P. B., 1979: Geostrophic turbulence. Annu. Rev. Fluid Mech., 11, 401–441, https://doi.org/10.1146/annurev.fl.11.010179.002153.
Salby, M. L., 1996: Fundamentals of Atmospheric Physics. International Geophysics Series, Vol. 61, Academic Press, 627 pp.
Saw, E.-W., D. Kuzzay, D. Faranda, A. Guittonneau, F. Daviaud, C. Wiertel-Gasquet, V. Padilla, and B. Dubrulle, 2016: Experimental characterization of extreme events of inertial dissipation in a turbulent swirling flow. Nat. Commun., 7, 12466, https://doi.org/10.1038/ncomms12466.
Schertzer, D., S. Lovejoy, F. Schmitt, Y. Chigirinskaya, and D. Marsan, 1997: Multifractal cascade dynamics and turbulent intermittency. Fractals, 5, 427–471, https://doi.org/10.1142/S0218348X97000371.
Seinfeld, J. H., and S. N. Pandis, 2016: Atmospheric Chemistry and Physics: From Air Pollution to Climate Change. John Wiley and Sons, 1152 pp.
Sellers, W. D., 1969: A global climatic model based on the energy balance of the earth-atmosphere system. J. Appl. Meteor., 8, 392–400, https://doi.org/10.1175/1520-0450(1969)008<0392:AGCMBO>2.0.CO;2.
Tung, K. K., and W. W. Orlando, 2003: The k−3 and k−5/3 energy spectrum of atmospheric turbulence: Quasigeostrophic two-level model simulation. J. Atmos. Sci., 60, 824–835, https://doi.org/10.1175/1520-0469(2003)060<0824:TKAKES>2.0.CO;2.
Viterbo, P., and A. C. Beljaars, 1995: An improved land surface parameterization scheme in the ECMWF model and its validation. J. Climate, 8, 2716–2748, https://doi.org/10.1175/1520-0442(1995)008<2716:AILSPS>2.0.CO;2.
Viterbo, P., and A. K. Betts, 1999: Impact on ECMWF forecasts of changes to the albedo of the boreal forests in the presence of snow. J. Geophys. Res., 104, 27 803–27 810, https://doi.org/10.1029/1998JD200076.
Žagar, N., D. Jelić, M. Blaauw, and P. Bechtold, 2017: Energy spectra and inertia–gravity waves in global analyses. J. Atmos. Sci., 74, 2447–2466, https://doi.org/10.1175/JAS-D-16-0341.1.