Advanced Search
Article Contents

Parameterization of Sheared Entrainment in a Well-Developed CBL. Part I: Evaluation of the Scheme through Large-Eddy Simulations


doi: 10.1007/s00376-016-5208-x

  • The entrainment flux ratio A e and the inversion layer (IL) thickness are two key parameters in a mixed layer model. A e is defined as the ratio of the entrainment heat flux at the mixed layer top to the surface heat flux. The IL is the layer between the mixed layer and the free atmosphere. In this study, a parameterization of A e is derived from the TKE budget in the first-order model for a well-developed CBL under the condition of linearly sheared geostrophic velocity with a zero value at the surface. It is also appropriate for a CBL under the condition of geostrophic velocity remaining constant with height. LESs are conducted under the above two conditions to determine the coefficients in the parameterization scheme. Results suggest that about 43% of the shear-produced TKE in the IL is available for entrainment, while the shear-produced TKE in the mixed layer and surface layer have little effect on entrainment. Based on this scheme, a new scale of convective turbulence velocity is proposed and applied to parameterize the IL thickness. The LES outputs for the CBLs under the condition of linearly sheared geostrophic velocity with a non-zero surface value are used to verify the performance of the parameterization scheme. It is found that the parameterized A e and IL thickness agree well with the LES outputs.
  • 加载中
  • Arya S. P. S., J. C. Wyngaard, 1975: Effect of baroclinicity on wind profiles and the geostrophic drag law for the convective planetary boundary layer. J. Atmos. Sci., 32, 767- 778.10.1175/1520-0469(1975)032<0767:EOBOWP>2.0.CO;2cafee80efe2adbeb012c25cd57a6cc6ehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1975jats...32..767ahttp://adsabs.harvard.edu/abs/1975jats...32..767aBy using a simple physical model of the baroclinic convective planetary boundary layer, the similarity functions of the geostrophic drag law are expressed as sums of a barotropic part, dependent only on the stability and boundary layer height parameters, and a baroclinicity dependent part. The latter are predicted to he sinusoidal functions of the angle between surface wind and geostrophic shear, their amplitudes being proportional to the normalized magnitude of geostrophic shear. These drag laws are confirmed by the results of a more sophisticated higher-order closure model, which also predict the magnitude of actual wind shears in the bulk of the mixed layer remaining much smaller than the magnitude of imposed geostrophic shear. The results are shown to be supported by some observations from the recent Wangara and ATFX experiments. The surface cross-isobar angle is predicted to increase toward the equator, a trend well confirmed by observations, but in obvious conflict with the drag laws proposed by others who have ignored the height of the lowest inversion base from their similarity considerations.
    Batchvarova E., S.-E. Gryning, 1994: An applied model for the height of the daytime mixed layer and the entrainment zone. Bound.-Layer Meteor., 71, 311- 323.10.1007/BF007137448d58ec7b7321fe88a676d495107746cchttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1994bolme..71..311bhttp://adsabs.harvard.edu/abs/1994bolme..71..311bA model is presented for the height of the mixed layer and the depth of the entrainment zone under near-neutral and unstable atmospheric conditions. It is based on the zero-order mixed-layer height model of Batchvarova and Gryning (1991) and the parameterization of the entrainment zone depth proposed by Gryning and Batchvarova (1994). However, most zero-order slab type models of mixed-layer height may be applied. The use of the model requires only information on those meteorological parameters that are needed in operational applications of ordinary zero-order slab type models of mixed-layer height: friction velocity, kinematic heat flux near the ground and potential temperature gradient in the free atmosphere above the entrainment zone. When information is available on the horizontal divergence of the large-scale flow field, the model also takes into account the effect of subsidence, although this is usually neglected in operational models of mixed-layer height owing to lack of data. Model performance is tested using data from the CIRCE experiment.
    Betts A. K., 1974: Reply to comment on the paper "Non-precipitating cumulus convection and its parameterization". Quart. J. Roy. Meteor. Soc., 100, 469- 471.10.1002/qj.4971004251714c5d6194e87e61204c8e10412533060http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49710042517%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1002/qj.49710042517/citedbyNot Available
    Boers R., E. W. Eloranta, and R. L. Coulter, 1984: Lidar observations of mixed layer dynamics: Tests of parameterized entrainment models of mixed layer growth rate. J. Climate Appl. Meteor., 23, 247- 266.10.1175/1520-0450(1984)0232.0.CO;21d4cb8c1ab531612a8bf9f0cb22fb47dhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1984JApMe..23..247Bhttp://adsabs.harvard.edu/abs/1984JApMe..23..247BGround based lidar measurements of the atmospheric mixed layer depth, the entrainment zone depth and the wind speed and wind direction were used to test various parameterized entrainment models of mixed layer growth rate. Six case studies under clear air convective conditions over flat terrain in central Illinois are presented. It is shown that surface heating alone accounts for a major portion of the rise of the mixed layer on all days. A new set of entrainment model constants was determined which optimized height predictions for the dataset. Under convective conditions, the shape of the mixed layer height prediction curves closely resembled the observed shapes. Under conditions when significant wind shear was present, the shape of the height prediction curve departed from the data suggesting deficiencies in the parameterization of shear production. Development of small cumulus clouds on top of the layer is shown to affect mixed layer depths in the afternoon growth phase.
    Conzemius R. J., E. Fedorovich, 2006a: Dynamics of sheared convective boundary layer entrainment. Part I: Methodological background and large-eddy simulations. J. Atmos. Sci., 63, 1151- 1178.10.1175/JAS3691.1dc567ba9229c2a4a6242a7158880b645http%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D2216927http://www.ams.org/mathscinet-getitem?mr=2216927Abstract The reported study examines the dynamics of entrainment and its effects on the evolution of the dry atmospheric convective boundary layer (CBL) when wind shear is present. The sheared CBL can be studied by means of direct measurements in the atmosphere, laboratory studies, and numerical techniques. The advantages and disadvantages of each technique are discussed in the present paper, which also describes the methodological background for studying the dynamics of entrainment in sheared CBLs. For the reported study, large-eddy simulation (LES) was chosen as the primary method of convective entrainment investigation. Twenty-four LES runs were conducted for CBLs growing under varying conditions of surface buoyancy flux, free-atmospheric stratification, and wind shear. The simulations were divided into three categories: CBL with no mean wind (NS), CBL with a height-constant geostrophic wind of 20 m s鈭1 (GC), and CBL with geostrophic wind shear (GS). In the simulated cases, the sheared CBLs grew fastes...
    Conzemius R. J., E. Fedorovich, 2006b: Dynamics of sheared convective boundary layer entrainment. Part II: Evaluation of bulk model predictions of entrainment flux. J. Atmos. Sci., 63, 1179- 1199.10.1175/JAS3696.1b40c3287-a005-4131-98f4-e4e1940fd58ca0ed3d2267609603cc6daf82ae61a4d1http%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D2216928refpaperuri:(39f545aff06888e3d4a6d9c0cb9971b8)http://www.ams.org/mathscinet-getitem?mr=2216928Abstract Several bulk model–based entrainment parameterizations for the atmospheric convective boundary layer (CBL) with wind shear are reviewed and tested against large-eddy simulation (LES) data to evaluate their ability to model one of the basic integral parameters of convective entrainment—the entrainment flux ratio. Test results indicate that many of these parameterizations fail to correctly reproduce entrainment flux in the presence of strong shear because they underestimate the dissipation of turbulence kinetic energy (TKE) produced by shear in the entrainment zone. It is also found that surface shear generation of TKE may be neglected in the entrainment parameterization because it is largely balanced by dissipation. However, the surface friction has an indirect effect on the entrainment through the modification of momentum distribution in the mixed layer and regulation of shear across the entrainment zone. Because of this effect, parameterizations that take into account the surface friction veloci...
    Conzemius R. J., E. Fedorovich, 2007: Bulk models of the sheared convective boundary layer: Evaluation through large eddy simulations. J. Atmos. Sci., 64, 786- 807.10.1175/JAS3870.15f2cb93756df7f36d1efd427378c30c7http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2007JAtS...64..786Chttp://adsabs.harvard.edu/abs/2007JAtS...64..786CAbstract A set of first-order model (FOM) equations, describing the sheared convective boundary layer (CBL) evolution, is derived. The model output is compared with predictions of the zero-order bulk model (ZOM) for the same CBL type. Large eddy simulation (LES) data are employed to test both models. The results show an advantage of the FOM over the ZOM in the prediction of entrainment, but in many CBL cases, the predictions by the two models are fairly close. Despite its relative simplicity, the ZOM is able to quantify the effects of shear production and dissipation in an integral sense鈥攁s long as the constants describing the integral dissipation of shear- and buoyancy-produced turbulence kinetic energy (TKE) are prescribed appropriately and the shear is weak enough that the denominator of the ZOM entrainment equation does not approach zero, causing a numerical instability in the solutions. Overall, the FOM better predicts the entrainment rate due to its ability to avoid this instability. Also, the FOM in a more physically consistent manner reproduces the sheared CBL entrainment zone, whose depth is controlled by a balance among shear generation, buoyancy consumption, and dissipation of TKE. Such balance is manifested by nearly constant values of Richardson numbers observed in the entrainment zone of simulated sheared CBLs. Conducted model tests support the conclusion that the surface shear generation of TKE and its corresponding dissipation, as well as the nonstationary terms, can be omitted from the integral TKE balance equation.
    Deardorff J. W., 1979: Prediction of convective mixed-layer entrainment for realistic capping inversion structure. J. Atmos. Sci., 36, 424- 436.10.1175/1520-0469(1979)0362.0.CO;2ea378b025e44c61379785777e54138b5http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FADS%3Fid%3D1979JAtS...36..424Dhttp://onlinelibrary.wiley.com/resolve/reference/ADS?id=1979JAtS...36..424DThe first-order jump model for the potential temperature or buoyancy variable at the capping inversion atop a convectively mixed layer is reexamined and found to imply existence of an entrainment rate equation which is unreliable. The model is therefore extended here to allow all the negative buoyancy flux of entrainment to occur within the interfacial layer of thickness Δand to allow realistic thermal structure within the layer. The new model yields a well behaved entrainment rate equation requiring scarcely any closure assumption in the cases of steady-state entrainment with large-scale subsidence, and pseudo-encroachment. For nonsteady entrainment the closure assumption need only be made on (Δ)/in order to obtain the entrainment rates at both the outer and inner edges of the interfacial layer. A particular closure assumption for (Δ)/is tested against five laboratory experiments and found to yield favorable results for both Δand the mixed-layer thickness if the initial value of Δis known. It is also compared against predictions from two zero-order jump models which do not attempt prediction of Δand one first-order jump model.
    Deardorff J. W., 1980: Stratocumulus-capped mixed layers derived from a three-dimensional model. Bound.-Layer Meteor., 18, 495- 527.6618e7dc0e5515dbf5ac81d7f5e9eb65http%3A%2F%2Fwww.nrcresearchpress.com%2Fservlet%2Flinkout%3Fsuffix%3Drefg4%2Fref4%26dbid%3D16%26doi%3D10.1139%252Fcjfr-2014-0184%26key%3D10.1007%252FBF00119502http://xueshu.baidu.com/s?wd=paperuri%3A%282cb7d26b0f739a344dfcfa43abf9f3ee%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.nrcresearchpress.com%2Fservlet%2Flinkout%3Fsuffix%3Drefg4%2Fref4%26dbid%3D16%26doi%3D10.1139%252Fcjfr-2014-0184%26key%3D10.1007%252FBF00119502&ie=utf-8&sc_us=9920505749051549919
    Deardorff J. W., G. E. Willis, and B. H. Stockton, 1980: Laboratory studies of the entrainment zone of a convectively mixed layer. J. Fluid Mech., 100, 41- 64.10.1017/S0022112080001000f1c617fa1d6f83097048125d24ab6792http%3A%2F%2Fjournals.cambridge.org%2Farticle_S0022112080001000http://journals.cambridge.org/article_S0022112080001000Not Available
    Driedonks A. G. M., 1982: Models and observations of the growth of the atmospheric boundary layer. Bound.-Layer Meteor., 23, 283- 306.10.1007/BF001211179b1feedd63d1e4c5ad37dca0b80d9b13http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1982BoLMe..23..283Dhttp://adsabs.harvard.edu/abs/1982BoLMe..23..283DThe evolution of the mixed layer during a clear day can be described with a slab model. The model equations have to be closed by a parameterization of the turbulent kinetic energy budget. Several possibilities for this parameterization have been proposed. In order to assess the practical applicability of these models for the atmosphere, field experiments were carried out on ten clear days in 1977 and 1978. Within the accuracy of the measurements the mixed-layer height in fully convective conditions (at noon on clear days) is well predicted taking a constant heat flux ratio [ - overline {θ w_h } = 0.2overline {θ w_s } ]. In the early morning hours mechanical entrainment is also important. Good overall results are obtained with the entrainment formulation [ - overline {θ w_h } = 0.2overline {θ w_s } + 5u_ * ^3 T/gh]. Only large differences in the entrainment coefficients lead to significantly different results. Making the entrainment model more complex does not lead to substantial improvement. The mean potential temperature in the mixed layer is reproduced within 0.5 °C. This temperature is insensitive to the choice of a particular entrainment formulation and depends more on the surface heat input and the temperature gradient in the stable air aloft.
    Fedorovich E., 1995: Modeling the atmospheric convective boundary layer within a zero-order jump approach: An extended theoretical framework. J. Appl. Meteor., 34, 1916- 1928.10.1175/1520-0450(1995)0342.0.CO;2960bb803447a2f112425d5b5e9e2fdfbhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1995JApMe..34.1916Fhttp://adsabs.harvard.edu/abs/1995JApMe..34.1916FThe paper presents an extended theoretical background for applied modeling of the atmospheric convective boundary layer within the so-called zero-order jump approach, which implies vertical homogeneity of meteorological fields in the bulk of convective boundary layer (CBL) and zero-order discontinuities of variables at the interface of the layer.The zero-order jump model equations for the most typical cases of CBL are derived. The models of nonsteady, horizontally homogeneous CBL with and without shear, extensively studied in the past with the aid of zero-order jump models, are shown to be particular cases of the general zero-order jump theoretical framework. The integral budgets of momentum and heat are considered for different types of dry CBL. The profiles of vertical turbulent fluxes are presented and analyzed. The general version of the equation of CBL depth growth rate (entrainment rate equation) is obtained by the integration of the turbulence kinetic energy balance equation, invoking basic assumptions of the zero-order parameterizations of the CBL vertical structure. The problems of parameterizing the turbulence vertical structure and closure of the entrainment rate equation for specific cases of CBL are discussed. A parameterization scheme for the horizontal turbulent exchange in zero-order jump models of CBL is proposed. The developed theory is generalized for the case of CBL over irregular terrain.
    Fedorovich E. E., D. V. Mironov, 1995: A model for a shear-free convective boundary layer with parameterized capping inversion structure. J. Atmos. Sci., 52, 83- 95.10.1175/1520-0469(1995)0522.0.CO;28486a60388d6003b2f04f303266a9cc2http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1995JAtS...52...83Fhttp://adsabs.harvard.edu/abs/1995JAtS...52...83FThe paper extends Deardorff's general structure parameterization for a shear-free convective boundary layer. The model suggested employs the mixed layer hypothesis that the buoyancy (which is defined as b = g(rho(sub 0) - rho/rho(sub 0) where rho is the density, rho(sub 0) is the reference density, and g is the acceleration due to gravity) is constant with height within the mixed layer. The buoyancy flux zero-crossing height is taken as the mixed layer. The buoyancy flux zero-crossing height is taken as the mixed layer depth. The vertical buoyancy profile within the capping inversion, where the buoyancy flux is negative due to entrainment, is made dimensionless, using the buoyancy difference across the inversion and its thickness as appropiate scales. The auhtors examine the idea against the data from atmospheric measurements, laboratory experiments with buoyancy-agitated turbulence, and large-eddy simulations. The rate equations for the mixed layer and inversion layer depths are derived using the turbulent kinetic energy equation and Deardorff's scaling hypothesis refined to account for the inversion layer structure. The constants of the model are evaluated from the data of atmospheric, oceanic, and laboratory measurements, and large-eddy simulations. The causes of divergence of the estimates based on data of different origin are discussed. The model is applied to simulate convective entrainment in laboratory experiments. A reasonable explanation for ambiguous behavior of the entrainment zone in the experiments with a two-layer fluid is suggested. The model simulates transition regimes of convective entrainment in multilayer fluid strongly affected by the nonstationary of the entrainment zone.
    Fedorovich E., R. Conzemius, and D. Mironov, 2004a: Convective entrainment into a shear-free, linearly stratified atmosphere: bulk models reevaluated through large eddy simulations. J. Atmos. Sci., 61, 281- 295.10.1175/1520-0469(2004)0612.0.CO;262a794565bb20787780f79ab6f52e61bhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004JAtS...61..281Fhttp://adsabs.harvard.edu/abs/2004JAtS...61..281FRelationships between parameters of convective entrainment into a shear-free, linearly stratified atmosphere predicted by the zero-order jump and general-structure bulk models of entrainment are reexamined using data from large eddy simulations (LESs). Relevant data from other numerical simulations, water tank experiments, and atmospheric measurements are also incorporated in the analysis. Simulations have been performed for 10 values of the buoyancy gradient in the free atmosphere covering a typical atmospheric stability range. The entrainment parameters derived from LES and relationships between them are found to be sensitive to the model framework employed for their interpretation. Methods of determining bulk model entrainment parameters from the LES output are proposed and discussed. Within the range of investigated free-atmosphere stratifications, the LES predictions of the inversion height and buoyancy increment across the inversion are found to be close to the analytical solutions for the equilibrium entrainment regime, which is realized when the rate of time change of the CBL-mean turbulence kinetic energy and the energy drain from the CBL top are both negligibly small. The zero-order model entrainment ratio of about 0.2 for this regime is generally supported by the LES data. However, the zero-order parameterization of the entrainment layer thickness is found insufficient. A set of relationships between the general-structure entrainment parameters for typical atmospheric stability conditions is retrieved from the LES. Dimensionless constants in these relationships are estimated from the LES and laboratory data. Power-law approximations for relationships between the entrainment parameters in the zero-order jump and general-structure bulk models are evaluated based on the conducted LES. In the regime of equilibrium entrainment, the stratification parameter of the entrainment layer, which is the ratio of the buoyancy gradient in the free atmosphere to the overall buoyancy gradient across the entrainment layer, appears to be a constant of about 1.2.
    Fedorovich E., Coauthors, 2004b: Entrainment into sheared convective boundary layers as predicted by different large eddy simulation codes. 16th Symposium on Boundary Layers and Turbulence, Portland, ME, Amer. Meteor. Soc.6b1aea7f319dfadcff87c4cd5d77141fhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F40124905_Entrainment_into_sheared_convective_boundary_layers_as_predicted_by_different_large_eddy_simulation_codeshttp://www.researchgate.net/publication/40124905_Entrainment_into_sheared_convective_boundary_layers_as_predicted_by_different_large_eddy_simulation_codesEntrainment is a complex physical process that is a driving mechanism of convective boundary-layer (CBL) development. There is no consensus in the boundarylayer research community regarding the role played by wind shears in the entrainment dynamics.
    Flamant C., J. Pelon, B. Brashers, and R. Brown, 1999: Evidence of a mixed-layer dynamics contribution to the entrainment process. Bound.-Layer Meteor., 93, 47- 73.10.1023/A:1002083425811d4c51794fdb300ea7d1a8b1f54b3b816http%3A%2F%2Fwww.ingentaconnect.com%2Fcontent%2Fklu%2Fboun%2F1999%2F00000093%2F00000001%2F00237681http://www.ingentaconnect.com/content/klu/boun/1999/00000093/00000001/00237681The internal thermal boundary layer developing over the Mediterranean during a cold-air outbreak associated with a Tramontane event has been studied by means of airborne lidar, in situ sensors, and a modelling approach that consisted of nesting the University of Washington (UW) planetary boundary-layer (PBL) model in an advective zero-order jump model. This approach bypasses some of the deficiencies associated with each model: the absence of the dynamics in the mixed layer for the zero-order jump model and the lack of an inversion at the PBL top for the UW PBL model. Particular attention is given to the parameterization of the entrainment flux at the PBL top. Values of the entrainment closure parameter derived with the model when matching PBL structure observations are much lower than those derived with standard zero-order jump models. They also are in good agreement with values measured in different meteorological situations by other studies. This improvement is a result of the introduction of turbulent kinetic energy production in the mixed layer.
    Garcia J. R., J. P. Mellado, 2014: The two-layer structure of the entrainment zone in the convective boundary layer. J. Atmos. Sci., 71, 1935- 1955.10.1175/JAS-D-13-0148.18a6b7048-3a83-4da7-baac-e2c92fe64bb9feb0b137fb431b5383c0c35d3db837cchttp%3A%2F%2Fconnection.ebscohost.com%2Fc%2Farticles%2F96286975%2Ftwo-layer-structure-entrainment-zone-convective-boundary-layerrefpaperuri:(9fb41f6ea2ee89986da2641c687297c9)http://connection.ebscohost.com/c/articles/96286975/two-layer-structure-entrainment-zone-convective-boundary-layerAbstract The entrainment zone (EZ) of a dry, shear-free convective boundary layer growing into a linearly stratified fluid is studied by means of direct numerical simulation. The scale separation between the boundary layer thickness and the Kolmogorov length scale is shown to be sufficient to observe Reynolds number similarity in the statistics of interest during the equilibrium entrainment regime. Contrary to previous considerations, the vertical structure of the entrainment zone is found to be better described by the superposition of two sublayers: 1) an upper EZ sublayer that is dominated by overshooting thermals and is characterized by a penetration depth that scales with the ratio of the convective velocity and the buoyancy frequency of the free troposphere and 2) a lower EZ sublayer that is dominated by troughs of mixed fluid and is characterized by the integral length scale of the mixed layer. Correspondingly, different buoyancy scales are identified. The consequences of this multiplicity of scales on the entrainment rate parameters are evaluated directly, without resorting to any bulk model, through an exact relation among the mean entrainment rate, the local buoyancy increment, and both the turbulent and the finite-thickness contributions to the entrainment ratio A measured at the height of minimum buoyancy flux. The smaller turbulent contribution to A that is usually observed for relatively thick EZs is found to be compensated by the smaller local buoyancy increment instead of by the finite-thickness contribution. The two-layer structure of the entrainment zone is found to affect the exponent of the power-law relation between the normalized mean entrainment rate and the convective Richardson number such that the exponent deviates from 鈭1 for typical atmospheric conditions, although it asymptotically approaches 1 for higher Richardson numbers.
    Gentine P., G. Bellon, and C. C. van Heerwaarden, 2015: A closer look at boundary layer inversion in large-eddy simulations and bulk models: Buoyancy-driven case. J. Atmos. Sci., 72, 728- 749.10.1175/JAS-D-13-0377.112c584dede0768688d18df132f541dc6http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2015JAtS...72..728Ghttp://adsabs.harvard.edu/abs/2015JAtS...72..728GNot Available
    Heus T., Coauthors, 2010: Formulation of the dutch atmospheric large-eddy simulation (DALES) and overview of its applications. Geoscientific Model Development, 3, 415- 444.10.5194/gmd-3-415-20104a2f69a50533fe8fd1ae86e1f5b4a9bbhttp%3A%2F%2Fwww.oalib.com%2Fpaper%2F1377513http://www.oalib.com/paper/1377513The current version of the Dutch Atmospheric Large-Eddy Simulation (DALES) is presented. DALES is a large-eddy simulation code designed for studies of the physics of the atmospheric boundary layer, including convective and stable boundary layers as well as cloudy boundary layers. In addition, DALES can be used for studies of more specific cases, such as flow over sloping or heterogeneous terrain, and dispersion of inert and chemically active species. This paper contains an extensive description of the physical and numerical formulation of the code, and gives an overview of its applications and accomplishments in recent years.
    Hoxit L. R., 1974: Planetary boundary layer winds in baroclinic conditions. J. Atmos. Sci., 31, 1003- 1020.10.1175/1520-0469(1974)0312.0.CO;21c89de640d222522d10c40a13972ade9http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1974JAtS...31.1003Hhttp://adsabs.harvard.edu/abs/1974JAtS...31.1003HSystematic stratifications and analyses of low-level radiosonde data are performed for portions of the eastern half of the United States. The procedures are designed to specify changes in the planetary boundary layer wind profile resulting from variations in baroclinicity. The angle between the winds and isobars, the ageostrophic wind components, the surface stress, and the surface wind speeds are all shown to be functions of the orientation of the thermal wind vector relative to the surface geostrophic wind. These variations are consistent with a mixing-length model of the additional turbulent momentum transport initiated by the vertical shear of the geostrophic wind.
    Huang J. P., X. H. Lee, and E. G. Patton, 2011: Entrainment and budgets of heat, water vapor, and carbon dioxide in a convective boundary layer driven by time-varying forcing. J.Geophys. Res., 116, D06308.10.1029/2010JD01493860005e7008a7001328c7d1974219a48fhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010JD014938%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1029/2010JD014938/pdfA large-eddy simulation (LES) code is coupled with a land surface model to investigate the diurnal variation of the atmospheric boundary layer (ABL). The diurnal evolution of the ABL is driven by a time-varying incoming solar radiation. The results show that the domain average surface fluxes of sensible heat, water vapor, and carbon dioxide are smooth functions of time but the fluxes at any given surface grid point show random variations, especially the sensible heat flux. At the ABL top, the LES-resolved entrainment fluxes of these scalars also evolve with time and are not fixed fractions of their respective surface fluxes. Entrainment efficiency (the ratio of entrainment flux at zto w, where zis the ABL height, wis entrainment velocity, and is the jump of scalar across the entrainment zone) is highest for COand lowest for sensible heat. The first-order jump condition model is very good approximation to simulated entrainment fluxes which are largely controlled by the vertical gradients of the scalars across the capping inversion. Our results suggest that over the range of geostrophic winds considered (0-5 m s), neither the surface nor the entrainment flux reveals sensitivity to the geostrophic wind speed variations.
    Kim S.-W., S.-U. Park, and C.-H. Moeng, 2003: Entrainment processes in the convective boundary layer with varying wind shear. Bound.-Layer Meteor., 108, 221- 245.10.1023/A:102417022929319654e64e98c75ef5082c6ead81963e1http%3A%2F%2Fwww.ingentaconnect.com%2Fcontent%2Fklu%2Fboun%2F2003%2F00000108%2F00000002%2F05111563http://www.ingentaconnect.com/content/klu/boun/2003/00000108/00000002/05111563Large-eddy simulations (LES) are performed to investigate the entrainment and the structure of the inversion layer of the convective boundary layer (CBL) with varying wind shears. Three CBLs are generated with the constant surface kinematic heat flux of 0.05 K m sand varying geostrophic wind speeds from 5 to 15 m s. Heat flux profiles show that the maximum entrainment heat flux as a fraction of the surface heat flux increases from 0.13 to 0.30 in magnitude with increasing wind shear. The thickness of the entrainment layer, relative to the depth of the well-mixed layer, increases substantially from 0.36 to 0.73 with increasing wind shear. The identification of vortices and extensive flow visualizations near the entrainment layer show that concentrated vortices perpendicular to the mean boundary-layer wind direction are identified in the capping inversion layer for the case of strong wind shear. These vortices are found to develop along the mean wind directions over strong updrafts, which are generated by convective rolls and to appear as large-scale wavy motions similar to billows generated by the Kelvin Helmholtz instability. Quadrant analysis of the heat flux shows that in the case of strong wind shear, large fluctuations of temperature and vertical velocity generated by large amplitude wavy motions result in greater heat flux at each quadrant than that in the weak wind shear case.
    Kim S.-W., S.-U. Park, D. Pino, and J. V.-G. de Arellano, 2006: Parameterization of entrainment in a sheared convective boundary layer using a first-order jump model. Bound.-Layer Meteor., 120, 455- 475.10.1007/s10546-006-9067-31c4677c17df08cea388547149589eaffhttp%3A%2F%2Fwww.ingentaconnect.com%2Fcontent%2Fklu%2Fboun%2F2006%2F00000120%2F00000003%2F00009067http://www.ingentaconnect.com/content/klu/boun/2006/00000120/00000003/00009067Basic entrainment equations applicable to the sheared convective boundary layer (CBL) are derived by assuming an inversion layer with a finite depth, i.e., the first-order jump model. Large-eddy simulation data are used to determine the constants involved in the parameterizations of the entrainment equations. Based on the integrated turbulent kinetic energy budget from surface to the top of the CBL, the resulting entrainment heat flux normalized by surface heat flux is a function of the inversion layer depth, the velocity jumps across the inversion layer, the friction velocity, and the convection velocity. The developed first-order jump model is tested against large-eddy simulation data of two independent cases with different inversion strengths. In both cases, the model reproduces quite reasonably the evolution of the CBL height, virtual potential temperature, and velocity components in the mixed layer and in the inversion layer.
    Lemone M. A., M. Y. Zhou, C.-H. Moeng, D. H. Lenschow, L. J. Miller, and R. L. Grossman, 1999: An observational study of wind profiles in the baroclinic convective mixed layer. Bound.-Layer Meteor., 90, 47- 82.10.1023/A:10017033036971745ef0cf1760def501daf72a3e86786http%3A%2F%2Fwww.ingentaconnect.com%2Fcontent%2Fklu%2Fboun%2F1999%2F00000090%2F00000001%2F00187684http://www.ingentaconnect.com/content/klu/boun/1999/00000090/00000001/00187684A comprehensive planetary boundary-layer (PBL) and synoptic data set is used to isolate the mechanisms that determine the vertical shear of the horizontal wind in the convective mixed layer. To do this, we compare a fair-weather convective PBL with no vertical shear through the mixed layer (10 March 1992), with a day with substantial vertical shear in the north-south wind component (27 February). The approach involves evaluating the terms of the budget equations for the two components of the vertical shear of the horizontal wind; namely: the time-rate-of-change or time-tendency term, differential advection, the Coriolis terms (a thermal wind term and a shear term), and the second derivative of the vertical transport of horizontal momentum with respect to height (turbulent-transport term). The data, gathered during the 1992 STorm-scale Operational and Research Meteorology (STORM) Fronts Experiments Systems Test (FEST) field experiment, are from gust-probe aircraft horizontal legs and soundings, 915-MHz wind profilers, a 5-cm Doppler radar, radiosondes, and surface Portable Automated Mesonet (PAM) stations in a roughly 50 脳 50 km boundary-layer array in north-eastern Kansas, nested in a mesoscale-to-synoptic array of radiosondes and surface data. We present evidence that the shear on 27 February is related to the rapid growth of the convective boundary layer. Computing the shear budget over a fixed depth (the final depth of the mixed layer), we find that the time-tendency term dominates, reflecting entrainment of high-shear air from above the boundary layer. We suggest that shear within the mixed layer occurs when the time-tendency term is sufficiently large that the shear-reduction terms namely the turbulent-transport term and differential advection terms 鈥 cannot compensate. In contrast, the tendency term is small for the slowly-growing PBL of 10 March, resulting in a balance between the Coriolis terms and the turbulent-transport term. Thus, the thermal wind appears to influence mixed-layer shear only indirectly, through its role in determining the entrained shear.
    Lewellen D. C., W. S. Lewellen, 1998: Large-eddy boundary layer entrainment. J. Atmos. Sci., 55, 2645- 2665.10.1175/1520-0469(1998)0552.0.CO;2611229986f11e8dd13d1165d4e08a028http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1998JAtS...55.2645Lhttp://adsabs.harvard.edu/abs/1998JAtS...55.2645LA series of large-eddy simulations have been performed to explore boundary layer entrainment under conditions of a strongly capped inversion layer with the boundary layer dynamics driven dominantly by buoyant forcing. Different conditions explored include cloud-top cooling versus surface heating, smoke clouds versus water clouds, variations in cooling height and optical depth of longwave radiation, degree of cloud-top evaporative instability, and modest wind shear. Boundary layer entrainment involves transport and mixing over a full range of length scales, as warm fluid from the region of the capping inversion is first transported into the boundary layer and then mixed throughout. While entrainment is often viewed as the small-scale process of capturing warm fluid from the inversion into the top of the boundary layer, this need not be the physics that ultimately determines the entrainment rate. In these simulations the authors find instead that the entrainment rate is often limited by the boundary layercale eddy transport and is therefore surprisingly insensitive to the smaller scales of mixing near the inversion. The fraction of buoyant energy production available to drive large eddies that is lost to entrainment rather than dissipation was found to be nearly constant over a wide range of simulation conditions, with no apparent fundamental difference between top- versus bottom-driven or cloudy versus clear boundary layers. In addition, it is found that for quasi-steady boundary layers with dynamics driven by cloud-top cooling there is an effective upper limit on the entrainment rate for which the boundary layer dynamics just remains coupled, which is often approached when the cloud top is evaporatively unstable.
    Mahrt L., D. H. Lenschow, 1976: Growth dynamics of the convectively mixed layer. J. Atmos. Sci., 33, 41- 51.10.1175/1520-0469(1976)0332.0.CO;218401ed2aba14c08432c3296d4eaad19http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1976JAtS...33...41Mhttp://adsabs.harvard.edu/abs/1976JAtS...33...41MA model for the growth of a convectively mixed layer is derived by layer integrating the basic equations and parameterizing unknown terms in the mixed layer turbulence kinetic energy equation by means of free convection similarity theory. When shear generation of turbulence energy is neglected in the turbulent inversion layer capping the mixed layer, the model essentially reduces to that of Tennekes. This shear generation is found to be important only in cases of significant baroclinicity and shallow mixed layer depth or small free flow stratification.
    Moeng C.-H., P. P. Sullivan, 1994: A comparison of shear- and buoyancy-driven planetary boundary layer flows. J. Atmos. Sci., 51, 999- 1022.10.1175/1520-0469(1994)051<0999:ACOSAB>2.0.CO;2a945d6409e2832addc2a5902eabf8f89http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1994JAtS...51..999Mhttp://adsabs.harvard.edu/abs/1994JAtS...51..999MPlanetary boundary layer (PBL) flows are known to exhibit fundamental differences depending on the relative combination of wind shear and buoyancy forces. These differences are not unexpected in that shear instabilities occur locally, while buoyancy force sets up vigorous thermals, which result in nonlocal transport of heat and momentum. At the same time, these two forces can act together to modify the flow field. In this study, four large-eddy simulations (LESs) spanning the shear and buoyancy flow regimes were generated; two correspond to the extreme cases of shear and buoyancy-driven PBLs, while the other two represent intermediate PBLs where both forces are important. The extreme cases are used to highlight and quantify the basic differences between shear and convective PBLs in 1) flow structures, 2) overall statistics, and 3) turbulent kinetic energy (TKE) budget distributions. Results from the two intermediate LES cases are used to develop and verify a velocity scaling and a TKE budget model, which are proposed for the intermediate PBL. The velocity variances and the variance fluxes (i.e., third moments) normalized by this velocity scaling are shown to become quantities on the order of one, and to lie mostly between those of the two extreme PBL cases. The proposed TKE budget model is shown to adequately reproduce the profiles of the TKE budget terms and the TKE.
    Nieuwstadt F. T. M., R. A. Brost, 1986: The decay of convective turbulence. J. Atmos. Sci., 43, 532- 546.10.1175/1520-0469(1986)0432.0.CO;2fbf50b319f50c812ee9c599a8adae4fehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1986JAtS...43..532Nhttp://adsabs.harvard.edu/abs/1986JAtS...43..532NAbstract Using simulations with a large-eddy model we have studied the decay of convective turbulence in the atmospheric boundary layer when the upward surface sensible heat flux is suddenly stopped. The decay of turbulent kinetic energy and temperature variance scales with the dimensionless time tw * / h . The temperature fluctuations start to decrease almost immediately after the forcing has been removed, whereas the turbulent kinetic energy stays constant for a time t 鈮 h / w * . Vertical velocity fluctuations decay faster than horizontal fluctuations. Entrainment persists well into the decay process and may explain departures from similarity. Some evidence suggests a decoupling of large and small scales during the decay.
    Otte M. J., J. C. Wyngaard, 2001: Stably stratified interfacial-layer turbulence from large-eddy simulation. J. Atmos. Sci., 58, 3424- 3442.10.1175/1520-0469(2001)0582.0.CO;238321ecbd1227f8b6186b1e8a7a85cbbhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FADS%3Fid%3D2001JAtS...58.3424Ohttp://onlinelibrary.wiley.com/resolve/reference/ADS?id=2001JAtS...58.3424OThe structure of the interfacial layer capping the atmospheric boundary layer is not well understood. The dominant influence on turbulence within the interfacial layer is the stable stratification induced by the capping inversion. A series of 26 high-resolution large eddy simulation runs ranging from neutral, inversion-capped to free-convection cases are used to study interfacial layer turbulence. The interfacial layer is found to be similar in many aspects to a classic stable boundary layer. For example, the shapes of interfacial layer spectra and cospectra, including the locations of the spectral peaks, agree with previous observations from nocturnal PBLs. The eddy diffusivities, variances, structure-function parameters, and dissipation rates within the interfacial layer, suitably nondimensionalized using local scaling, also agree with observations from nocturnal PBLs. These results may lead to improved models of the interfacial layer and entrainment, and may also have implications for remote sensing of the interfacial layer.
    Pino D., J. Vilà-Guerau De Arellano, 2008: Effects of shear in the convective boundary layer: analysis of the turbulent kinetic energy budget. Acta Geophysica, 56, 167- 193.10.2478/s11600-007-0037-zc39b39c1caf994827e9c06a3f877266chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2008AcGeo..56..167Phttp://adsabs.harvard.edu/abs/2008AcGeo..56..167PEffects of convective and mechanical turbulence at the entrainment zone are studied through the use of systematic Large-Eddy Simulation (LES) experiments. Five LES experiments with different shear characteristics in the quasi-steady barotropic boundary layer were conducted by increasing the value of the constant geostrophic wind by 5 m s-1 until the geostrophic wind was equal to 20 m s-1. The main result of this sensitivity analysis is that the convective boundary layer deepens with increasing wind speed due to the enhancement of the entrainment heat flux by the presence of shear. Regarding the evolution of the turbulence kinetic energy (TKE) budget for the studied cases, the following conclusions are drawn: (i) dissipation increases with shear, (ii) the transport and pressure terms decrease with increasing shear and can become a destruction term at the entrainment zone, and (iii) the time tendency of TKE remains small in all analyzed cases. Convective and local scaling arguments are applied to parameterize the TKE budget terms. Depending on the physical properties of each TKE budget contribution, two types of scaling parameters have been identified. For the processes influenced by mixed-layer properties, boundary layer depth and convective velocity have been used as scaling variables. On the contrary, if the physical processes are restricted to the entrainment zone, the inversion layer depth, the modulus of the horizontal velocity jump and the momentum fluxes at the inversion appear to be the natural choices for scaling these processes. A good fit of the TKE budget terms is obtained with the scaling, especially for shear contribution.
    Pino D., J. Vilà-Guerau de Arellano, and P. G. Duynkerke, 2003: The contribution of shear to the evolution of a convective boundary layer. J. Atmos. Sci., 60, 1913- 1926.10.1175/1520-0469(2003)0602.0.CO;2f9ad67c5412a6698d8b062a40a6a36fdhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FADS%3Fid%3D2003JAtS...60.1913Phttp://onlinelibrary.wiley.com/resolve/reference/ADS?id=2003JAtS...60.1913PStudies the role of shear in the development and maintenance of a convective boundary layer. Results of applying large eddy simulations (LES); Emphasis given to the growth of the boundary layer; Analysis of the processes that drive the mechanism.
    Pino D., J. Vilà-Guerau de Arellano, and S.-W. Kim, 2006: Representing sheared convective boundary layer by zeroth- and first-order-jump mixed-layer models: Large-eddy simulation verification. Journal of Applied Meteorology and Climatology, 45, 1224- 1243.10.1175/JAM2396.1b2517681b9dc4f23f2ca4524e246d307http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2006JApMC..45.1224Phttp://adsabs.harvard.edu/abs/2006JApMC..45.1224PDry convective boundary layers characterized by a significant wind shear on the surface and at the inversion are studied by means of the mixed-layer theory. Two different representations of the entrainment zone, each of which has a different closure of the entrainment heat flux, are considered. The simpler of the two is based on a sharp discontinuity at the inversion (zeroth-order jump), whereas the second one prescribes a finite depth of the inversion zone (first-order jump). Large-eddy simulation data are used to provide the initial conditions for the mixed-layer models, and to verify their results. Two different atmospheric boundary layers with different stratification in the free atmosphere are analyzed. It is shown that, despite the simplicity of the zeroth-order-jump model, it provides similar results to the first-order-jump model and can reproduce the evolution of the mixed-layer variables obtained by the large-eddy simulations in sheared convective boundary layers. The mixed-layer model with both closures compares better with the large-eddy simulation results in the atmospheric boundary layer characterized by a moderate wind shear and a weak temperature inversion. These results can be used to represent the flux of momentum, heat, and other scalars at the entrainment zone in general circulation or chemistry transport models.
    Rand all, D. A., 1984: Buoyant production and consumption of turbulence kinetic energy in cloud-topped mixed layers. J. Atmos. Sci., 41, 402- 413.10.1175/1520-0469(1984)0412.0.CO;2ab20bc3447559bd3fb02305590d90590http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1984JAtS...41..402Rhttp://adsabs.harvard.edu/abs/1984JAtS...41..402RNot Available
    Sorbjan Z., 1996a: Numerical study of penetrative and "solid lid" nonpenetrative convective boundary layers. J. Atmos. Sci., 53, 101- 112.10.1175/1520-0469(1996)0532.0.CO;2ef7e57b4408044b367364bbbe865b991http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1996JAtS...53..101Shttp://adsabs.harvard.edu/abs/1996JAtS...53..101SA large eddy simulation model was used to generate and compare statistics of turbulence during nonpenetrative and penetrative dry convection. In penetrative convection dimensionless vertical velocities in updrafts were found to have almost the same values as in the nonpenetrative case. The countergradient transport of heat and moisture was found to be present during nonpenetrative convection at /> 0.6. For penetrative convection the countergradient transport of heat occurred only in a layer 0.5 < /< 0.75, while the countergradient transport of humidity was not present. During nonpenetrative convection, temperature and humidity were perfectly correlated. In penetrative convection the correlation coefficient was found to be less than unity, varying from about 0.9 near the surface to about 0.7 at the top of the mixed layer.
    Sorbjan Z., 1996b: Effects caused by varying the strength of the capping inversion based on a large eddy simulation model of the shear-free convective boundary layer. J. Atmos. Sci., 53, 2015- 2024.10.1175/1520-0469(1996)0532.0.CO;27d830c87022924a07e018810456a5559http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1996JAtS...53.2015Shttp://adsabs.harvard.edu/abs/1996JAtS...53.2015SEffects caused by variation of the potential temperature lapse rate Γ in the free atmosphere are examined based on a “large eddy simulation” model of the shear-free convective atmospheric boundary layer. The obtained results show that only near the top of the boundary layer are the statistical moments involving temperature strongly sensitive to changes of the parameter Γ. Furthermore, the moments involving only the vertical velocity are practically independent of Γ. The ratio of the heat fluxes at the top and the bottom of the mixed layer increases when Γ increases. For the values of Γ from 1 to 10 K/km, typically observed in the atmosphere, the heat flux ratio varies in the range 610.2 to 610.3. When Γ increases by an order of magnitude to 100 K/km, increases only slightly to about 610.4. When Γ decreases to zero, the heat flux , at the top of the mixed layer also decreases to zero. In this case, the thermal structure of the atmospheric boundary layer is found to be similar to nonpenetrative “solid lid” convection in a tank.
    Sorbjan Z., 2004: Large-eddy simulations of the baroclinic mixed layer. Bound.-Layer Meteor., 112, 57- 80.10.1023/B:BOUN.0000020161.99887.b3134cfd0dc15c9cd6b29d3c0b7251e76bhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004BoLMe.112...57Shttp://adsabs.harvard.edu/abs/2004BoLMe.112...57SThe effects of baroclinicity, imposed on the dry mixed layer by the presenceof large-scale, horizontal temperature gradients, have been investigated basedon a large-eddy simulation model. The purpose of this investigation is to examinesimultaneous impacts of thermal stratification and shear in the atmospheric boundarylayer. For this purpose, five cases are considered - one barotropic, and four baroclinic.Based on the performed simulations, a new parametrization of the interfacial layer isproposed. The parameterization employs new interfacial scaling, which is valid at thetop of the mixed layer. In terms of new scales, dimensionless moments characterizingturbulence at the top of the shearless mixed layer are universal constants. In the shearedcase, dimensionless statistics of turbulence are shown to be functions of the interfacialRichardson number.
    Stull R. B., 1973: Inversion rise model based on penetrative convection. J. Atmos. Sci., 30, 1092- 1099.10.1175/1520-0469(1973)0302.0.CO;2696749287e25aaae500acbc2ac322284http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1973JAtS...30.1092Shttp://adsabs.harvard.edu/abs/1973JAtS...30.1092SA mathematical model to describe the height changes and other characteristics of an inversion base under the influence of surface convection and general subsidence is developed. Inversion interface dynamics and entrainment rates are formulated based on an unstable boundary layer environment of well-organized, plume-like, penetrative convection. The use of unstable boundary layer scaling velocities in describing the convection leads to a natural inclusion of the relevant parameters associated with inversions into this model. It is found that the model does accurately predict realistic rates of inversion rise and of temperature changes for conditions where organized free convection is prevalent.
    Stull R. B., 1976: The energetics of entrainment across a density interface. J. Atmos. Sci., 33, 1260- 1267.10.1175/1520-0469(1976)0332.0.CO;28e96c8c924d9f9c04481cecb8ca3e04ehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1976JAtS...33.1260Shttp://adsabs.harvard.edu/abs/1976JAtS...33.1260Sversion interface; and 4) energy losses due to internal gravity waves. It is shown that most previously published theories are just special cases of this more general energetics theory.
    Sullivan P. P., E. G. Patton, 2011: The effect of mesh resolution on convective boundary layer statistics and structures generated by large-eddy simulation. J. Atmos. Sci., 68, 2395- 2415.10.1175/JAS-D-10-05010.11d827eb869815f83ac35cb669fdf0e44http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2011JAtS...68.2395Shttp://adsabs.harvard.edu/abs/2011JAtS...68.2395SNot Available
    Sullivan P. P., C.-H. Moeng, B. Stevens, D. H. Lenschow, and S. D. Mayor, 1998: Structure of the entrainment zone capping the convective atmospheric boundary layer. J. Atmos. Sci., 55, 3042- 3064.10.1175/1520-0469(1998)0552.0.CO;20512918a15664cbeb9e3b0d7f365e95fhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1998JAtS...55.3042Shttp://adsabs.harvard.edu/abs/1998JAtS...55.3042SAbstract The authors use large-eddy simulation (LES) to investigate entrainment and structure of the inversion layer of a clear convectively driven planetary boundary layer (PBL) over a range of bulk Richardson numbers, Ri. The LES code uses a nested grid technique to achieve fine resolution in all three directions in the inversion layer. Extensive flow visualization is used to examine the structure of the inversion layer and to illustrate the temporal and spatial interaction of a thermal plume and the overlying inversion. It is found that coherent structures in the convective PBL, that is, thermal plumes, are primary instigators of entrainment in the Ri range 13.6 81 Ri 81 43.8. At Ri = 13.6, strong horizontal and downward velocities are generated near the inversion layer because of the plume–interface interaction. This leads to folding of the interface and hence entrainment of warm inversion air at the plume’s edge. At Ri = 34.5, the inversion’s strong stability prevents folding of the interface but stron...
    Sun J. N., Y. Wang, 2008: Effect of the entrainment flux ratio on the relationship between entrainment rate and convective Richardson number. Bound.-Layer Meteor., 126, 237- 247.10.1007/s10546-007-9231-4d93eceb40e9fe4602180137214e7aab4http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2008BoLMe.126..237Shttp://adsabs.harvard.edu/abs/2008BoLMe.126..237SThe parameterization of the dimensionless entrainment rate ( w / w ) versus the convective Richardson number ( Ri ) is discussed in the framework of a first-order jump model (FOM). A theoretical estimation for the proportionality coefficient in this parameterization, namely, the total entrainment flux ratio, is derived. This states that the total entrainment flux ratio in FOM can be estimated as the ratio of the entrainment zone thickness to the mixed-layer depth, a relationship that is supported by earlier tank experiments, and suggesting that the total entrainment flux ratio should be treated as a variable. Analyses show that the variability of the total entrainment flux ratio is actually the effect of stratification in the free atmosphere on the entrainment process, which should be taken into account in the parameterization. Further examination of data from tank experiments and large-eddy simulations demonstrate that the different power laws for w / w versus Ri can be interpreted as the variability of the total entrainment flux ratio. These results indicate that the dimensionless entrainment rate depends not only on the convective Richardson number but also upon the total entrainment flux ratio.
    Sun J. N., Q. J. Xu, 2009: Parameterization of sheared convective entrainment in the first-order jump model: Evaluation through large-eddy simulation. Bound.-Layer Meteor., 132, 279- 288.10.1007/s10546-009-9394-2f1d08ffc61b9b800ffe263dc42d01c1chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009BoLMe.132..279Shttp://adsabs.harvard.edu/abs/2009BoLMe.132..279SIn this note, two different approaches are used to estimate the entrainment-flux to surface-flux ratio for a sheared convective boundary layer (CBL); both are derived under the framework of the first-order jump model (FOM). That suggested by Sun and Wang (SW approach) has the advantage that there is no empirical constant included, though the dynamics are described in an implicit manner. The second, which was proposed by Kim et al. and Pino et al. (KP approach), explicitly characterizes the dynamics of the sheared entrainment, but uncertainties are induced through the empirical constants. Their performances in parameterizing the CBL growth rate are compared and discussed, and a new value of the parameter A in the KP approach is suggested. Large-eddy simulation (LES) data are employed to test both approaches: simulations are conducted for the CBL growing under varying conditions of surface roughness, free-atmospheric stratification, and wind shear, and data used when the turbulence is in steady state. The predicted entrainment rates in each case are tested against the LES data. The results show that the SW approach describes the evolution of the sheared CBL quite well, and the KP approach also reproduces the growth of the CBL reasonably, so long as the value of A is modified to 0.6.
    Sun J. N., W. M. Jiang, Z. Y. Chen, and R. M. Yuan, 2005: A laboratory study of the turbulent velocity characteristics in the convective boundary layer. Adv. Atmos. Sci.,22, 770-780, doi: 10.1007/BF02918721.10.1007/BF0291872190ebd1c2b7ffcd3757fe75ce4271a554http%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-DQJZ200505016.htmhttp://d.wanfangdata.com.cn/Periodical_dqkxjz-e200505017.aspxBased on the measurement of the velocity field in the convective boundary layer (CBL) in a convection water tank with the particle image velocimetry (PIV) technique, this paper studies the characteristics of the CBL turbulent velocity in a modified convection tank. The experiment results show that the velocity distribution in the mixed layer clearly possesses the characteristics of the CBL thermals, and the turbulent eddies can be seen obviously. The comparison of the vertical distribution of the turbulent velocity variables indicates that the modeling in the new tank is better than in the old one. The experiment data show that the thermal鈥檚 motion in the entrainment zone sometimes fluctuates obviously due to the intermittence of turbulence. Analyses show that this fluctuation can influence the agreement of the measurement data with the parameterization scheme, in which the convective Richardson number is used to characterize the entrainment zone depth. The normalized square velocity w i 2 / w * 2 at the top of the mixed layer seems to be time-dependent, and has a decreasing trend during the experiments. This implies that the vertical turbulent velocity at the top of the mixed layer may not be proportional to the convective velocity ( w * ).
    Tennekes H., 1973: A model for the dynamics of the inversion above a convective boundary layer. J. Atmos. Sci., 30, 558- 567.10.1175/1520-0469(1973)0302.0.CO;2a376200825a73affbad785ca0da5bea8http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1973JAtS...30..558Thttp://adsabs.harvard.edu/abs/1973JAtS...30..558TThe differential equations governing the strength Δ (a potential temperature difference) and the height of inversions associated with dry penetrative convection are considered. No assumptions on the magnitude of the downward heat flux at the inversion base are needed to obtain an algebraic equation that relates and Δ to the heating history of the boundary layer and to the initial conditions. After the nocturnal inversion has been filled in by heating, the inversion base generally grows linearly with time in the morning, but is proportional to the square root of time in the afternoon. The variation of Δ with time differs greatly from case to case.
    Tennekes H., A. G. M. Driedonks, 1981: Basic entrainment equations for the atmospheric boundary layer. Bound.-Layer Meteor., 20, 515- 531.10.1007/BF0012229982b7f52ef9fb22b14aae508e9f58bf5chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1981BoLMe..20..515Thttp://adsabs.harvard.edu/abs/1981BoLMe..20..515TThe parameterization of penetrative convection and other cases of turbulent entrainment by the atmospheric boundary layer is reviewed in this paper. The conservation equations for a one-layer model of entrainment are straightforward; all modeling problems arise in the context of the parameterization of various terms in the budget of turbulent kinetic energy. There is no consensus in the literature on the parameterization of shear production and of dissipation. Unfortunately, field experiments are not sufficiently accurate to guide the selection of suitable hypotheses. Carefully designed laboratory experiments are needed to settle the problems that remain.
    vanZanten, M. C., P. G. Duynkerke, J. W. M. Cuijpers, 1999: Entrainment parameterization in convective boundary layers. J. Atmos. Sci., 56, 813- 828.10.1175/1520-0469(1999)0562.0.CO;2fe9cc0b716154193f79ef27ea1304a58http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1999JAtS...56..813Vhttp://adsabs.harvard.edu/abs/1999JAtS...56..813VVarious runs were performed with a large eddy simulation (LES) model to evaluate different types of entrainment parametrizations. For this evaluation, three types of boundary layers were simulated: a clear convective boundary layer (CBL), a boundary layer containing a smoke concentration, and a cloud-topped boundary layer. It is shown that the assumption that the entrainment flux equals the product of the entrainment rate and the jump over a discontinuous inversion is not valid in CBLs simulated by an LES model. A finite inversion thickness (i.e., a first-order jump model) is needed to define an entrainment flux for which this approximation of the flux is valid. This entrainment flux includes not only the buoyancy flux at the inversion, but also the surface heat flux. The parameterization of the buoyancy flux at the inversion is evaluated for different closures, as suggested in the literature (i.e., Eulerian partitioning, process partitioning, and a closure developed by Deardorff), and where needed is extended for use in a first-order jump model. The closure based on process partitioning is found to yield consistent results in all types of convective boundary layers and shows the best agreement with the limit found in LES results if the longwave radiative flux divergence takes place in a much shallower layer than the mixed layer.
    Zeman O., H. Tennekes, 1977: Parameterization of the turbulent energy budget at the top of the daytime atmospheric boundary layer. J. Atmos. Sci., 34, 111- 123.10.1175/1520-0469(1977)034<0111:POTTEB>2.0.CO;29f35074020f33de9586710888b97b82chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1977jats...34..111zhttp://adsabs.harvard.edu/abs/1977jats...34..111zAbstract The budget of turbulent kinetic energy at the base of the inversion which caps the daytime atmospheric boundary layer depends on the lapse rate of potential temperature in the air aloft. The principal gain term in the energy budget is turbulent transport of kinetic energy, the principal loss term is buoyant conversion of kinetic energy into potential energy. The contributions made by these and other terms in the energy budget need to be parameterized for applications to inversion-rise prediction schemes. This paper contains a detailed analysis of the effects of dissipation near the inversion base, which leads to reduced entrainment if the air aloft is very stable. The parameterized energy budget also includes the Zilitinkevich correction, the influence of mechanical energy production near the inversion base, and modifications needed to incorporate cases in which the surface heat flux is negligible. Extensive comparisons of the theoretical model with experimental data indicate that a simplified treatment of the energy budget is adequate for forecasts of the development of convective mixed layers. The parameterization scheme is also applicable to thermocline erosion in the ocean; in that case, however, some of the minor terms in the energy budget often play a major role.
  • [1] Peng LIU, Jianning SUN, Lidu SHEN, 2016: Parameterization of Sheared Entrainment in a Well-developed CBL. Part II: A Simple Model for Predicting the Growth Rate of the CBL, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 1185-1198.  doi: 10.1007/s00376-016-5209-9
    [2] Lei ZHU, Chunsong LU, Xiaoqi XU, Xin HE, Junjun LI, Shi LUO, Yuan WANG, Fan WANG, 2024: The Probability Density Function Related to Shallow Cumulus Entrainment Rate and Its Influencing Factors in a Large-Eddy Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 173-187.  doi: 10.1007/s00376-023-2357-6
    [3] HAN Bo, LU Shihua, AO Yinhuan, 2012: Development of the Convective Boundary Layer Capping with a Thick Neutral Layer in Badanjilin: Observations and Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 177-192.  doi: 10.1007/s00376-011-0207-4
    [4] Yongqiang LIU, 2005: Land Breeze and Thermals: A Scale Threshold to Distinguish Their Effects, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 889-902.  doi: 10.1007/BF02918688
    [5] Yan ZHENG, Liguang WU, Haikun ZHAO, Xingyang ZHOU, Qingyuan LIU, 2020: Simulation of Extreme Updrafts in the Tropical Cyclone Eyewall, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 781-792.  doi: 10.1007/s00376-020-9197-4
    [6] LIU Huizhi, Sang Jianguo, 2011: Numerical Simulation of Roll Vortices in the Convective Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 477-482.  doi: 10.1007/s00376-010-9229-6
    [7] MIAO Shiguang, JIANG Weimei, 2004: Large Eddy Simulation and Study of the Urban Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 650-661.  doi: 10.1007/BF02915732
    [8] SUN Jianning, JIANG Weimei, CHEN Ziyun, YUAN Renmin, 2005: A Laboratory Study of the Turbulent Velocity Characteristics in the Convective Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 770-780.  doi: 10.1007/BF02918721
    [9] Shaofeng LIU, Michael HINTZ, Xiaolong LI, 2016: Evaluation of Atmosphere-Land Interactions in an LES from the Perspective of Heterogeneity Propagation, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 571-578.  doi: 10.1007/s00376-015-5212-6
    [10] GUO Xiaofeng, CAI Xuhui, 2005: Footprint Characteristics of Scalar Concentration in the Convective Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 821-830.  doi: 10.1007/BF02918682
    [11] LI Pingyang, JIANG Weimei, SUN Jianning, YUAN Renmin, 2003: A Laboratory Modeling of the Velocity Field in the Convective Boundary Layer with the Particle Image Velocimetry Technique, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 631-637.  doi: 10.1007/BF02915506
    [12] Bangjun Cao, Xianyu Yang, JUN WEN, Qin Hu, Ziyuan Zhu, 2023: Large eddy simulation of vertical structure and size distribution of deep layer clouds, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-023-3134-2
    [13] Surendra S. Parasnis, Savita B. Morwal, K. G. Vernekar, 1991: Convective Boundary Layer in the Region of the Monsoon Trough-A Case Study, ADVANCES IN ATMOSPHERIC SCIENCES, 8, 505-509.  doi: 10.1007/BF02919273
    [14] BIAN Jianchun, CHEN Hongbin, 2008: Statistics of the Tropopause Inversion Layer over Beijing, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 381-386.  doi: 10.1007/s00376-008-0381-1
    [15] Dongxiao WANG, Guojing LI, Lian SHEN, Yeqiang SHU, 2022: Influence of Coriolis Parameter Variation on Langmuir Turbulence in the Ocean Upper Mixed Layer with Large Eddy Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1487-1500.  doi: 10.1007/s00376-021-1390-6
    [16] HAN Bo, ZHAO Cailing, LÜ Shihua, WANG Xin, 2015: A Diagnostic Analysis on the Effect of the Residual Layer in Convective Boundary Layer Development near Mongolia Using 20th Century Reanalysis Data, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 807-820.  doi: 10.1007/s00376-014-4164-6
    [17] Zhao Ming, 1987: ON THE PARAMETERIZATION OF THE VERTICAL VELOCITY AT THE TOP OF PLANETARY BOUNDARY LAYER, ADVANCES IN ATMOSPHERIC SCIENCES, 4, 233-239.  doi: 10.1007/BF02677070
    [18] Zhong Zhong, Wang Hanjie, 2000: A Study of the Relationship between Low-level Jet and Inversion Layer over an Agroforest Ecosystem in East China Plain?, ADVANCES IN ATMOSPHERIC SCIENCES, 17, 299-310.  doi: 10.1007/s00376-000-0011-z
    [19] Li Xin, Hu Fei, Liu Gang, Hong Zhongxiang, 2001: Multi-scale Fractal Characteristics of Atmospheric Boundary-Layer Turbulence, ADVANCES IN ATMOSPHERIC SCIENCES, 18, 787-792.
    [20] SUN Jianning, 2009: On the Parameterization of Convective Entrainment: Inherent Relationships among Entrainment Parameters in Bulk Models, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 1005-1014.  doi: 10.1007/s00376-009-7222-8

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 10 October 2015
Manuscript revised: 02 February 2016
Manuscript accepted: 23 May 2016
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Parameterization of Sheared Entrainment in a Well-Developed CBL. Part I: Evaluation of the Scheme through Large-Eddy Simulations

  • 1. School of Atmospheric Sciences & Institute for Climate and Global Change, Nanjing University, Nanjing 210023, China
  • 2. Jiangsu Provincial Collaborative Innovation Center of Climate Change, Nanjing 210023, China

Abstract: The entrainment flux ratio A e and the inversion layer (IL) thickness are two key parameters in a mixed layer model. A e is defined as the ratio of the entrainment heat flux at the mixed layer top to the surface heat flux. The IL is the layer between the mixed layer and the free atmosphere. In this study, a parameterization of A e is derived from the TKE budget in the first-order model for a well-developed CBL under the condition of linearly sheared geostrophic velocity with a zero value at the surface. It is also appropriate for a CBL under the condition of geostrophic velocity remaining constant with height. LESs are conducted under the above two conditions to determine the coefficients in the parameterization scheme. Results suggest that about 43% of the shear-produced TKE in the IL is available for entrainment, while the shear-produced TKE in the mixed layer and surface layer have little effect on entrainment. Based on this scheme, a new scale of convective turbulence velocity is proposed and applied to parameterize the IL thickness. The LES outputs for the CBLs under the condition of linearly sheared geostrophic velocity with a non-zero surface value are used to verify the performance of the parameterization scheme. It is found that the parameterized A e and IL thickness agree well with the LES outputs.

1. Introduction
  • The development of the ABL over land is dominated by surface heating during the daytime. Convective activities and turbulent mixing are common within the daytime ABL, i.e., the CBL. The top of the CBL is capped by an interface layer, where the stratified free atmospheric air aloft is entrained down into the mixed layer by overshooting thermals. The entrainment process can significantly influence CBL evolution and profiles of mean variables within the CBL (Hoxit, 1974; Arya and Wyngaard, 1975; Lemone et al., 1999). It is important in NWP and air pollution models.

    There have been many studies on the entrainment process. (Kim et al., 2006) pointed out that much work had been done for the free CBL, whereas studies about the uncertainties in the entrainment process were quite limited compared to studies of the sheared CBL. In the past decade, research on the entrainment process has mainly focused on how to understand and parameterize the effect of wind shear (Kim et al., 2003, 2006; Pino et al., 2003, 2006; Conzemius and Fedorovich, 2006a, 2006b, 2007; Sun and Xu, 2009). Bulk models are often employed to describe CBL evolution. Two commonly used types of bulk models are the zeroth-order model (ZOM), which represents the inversion layer (IL) as an infinitesimally thin interface (Tennekes, 1973), and the first-order model (FOM), which assumes a certain IL thickness (Betts, 1974). The IL structure is important for CBL dynamics (Sorbjan, 1996a, 1996b, 2004; Lewellen and Lewellen, 1998; Sullivan et al., 1998; vanZanten et al., 1999; Otte and Wyngaard, 2001; Kim et al., 2003, 2006; Sun and Wang, 2008). (Conzemius and Fedorovich, 2007) reviewed previous studies of bulk models and suggested that at least the FOM is needed in order to adequately capture the entrainment process in a sheared CBL. (Huang et al., 2011) demonstrated that the FOM can adequately describe not only the entrainment heat flux but also the entrainment fluxes of water vapor and other conservative scalars such as carbon dioxide.

    The deficiencies of FOMs were reviewed in (Gentine et al., 2015). In FOMs based on (Betts, 1974), both the potential temperature and heat flux profiles are assumed linear in the IL, and the mixed-layer top is located with the minimum heat flux height. (Deardorff, 1979) argued that the representation of the IL in such models is oversimplified. Firstly, the observed maximum vertical gradient of potential temperature is generally much higher in observations than in the FOM. Secondly the minimum heat flux level is located above the mixed-layer top. Thirdly, the assumption that the mixed-layer height is equal to the minimum heat flux height generates a singularity for the IL growth rate equation under strong inversions. Thereafter, the IL thickness has generally been parameterized (e.g., defined as a function of the convective Richardson number) to avoid this singularity. On the other hand, the linear profiles are incompatible: parabolic flux profiles should correspond to linear profiles of conserved variables. (Deardorff, 1979) proposed a more realistic representation of the IL——the so-called general structure model (GSM). However, the structure of the IL in the GSM needs to be parameterized (Fedorovich and Mironov, 1995; Fedorovich et al., 2004a), which limits its applicability. To overcome the limitations in previous FOMs, (Gentine et al., 2015) proposed a new IL model based on a second-order polynomial for the potential temperature profile, and a third-order polynomial for the heat flux profile. This model can accurately prognosticate the growth rate of the IL, and of the mixed layer, under purely convective conditions. However, our study focuses on the entrainment process under shear conditions. It is not clear how wind shear impacts the profile shapes of velocity and potential temperature and their fluxes. In order to simplify the derivations, we use an FOM with linear profiles in the IL. The relative errors between the linear and curving flux profiles are discussed in this study.

    The bulk model consists of a set equations for the CBL. Parameterizations of the entrainment flux ratio A e (defined as the ratio of entrainment heat flux at the top of the mixed layer to the surface heat flux) and IL thickness are needed for closure of the CBL equations (e.g., Kim et al., 2006; Pino et al., 2006). (Kim et al., 2006) developed a parameterization of A e for the CBL under the condition of height-constant geostrophic velocity (GC case). (Conzemius and Fedorovich, 2007) developed a bulk CBL model under the condition of linearly sheared geostrophic velocity with a zero value at the surface (GS case), in which the A e is not explicitly expressed and the IL thickness is parameterized assuming a constant gradient Richardson number. Note that large uncertainties exist in the parameterization of sheared entrainment. (Pino et al., 2006) suggested that about 70% of TKE produced by wind shear across the IL is available for entrainment, whereas (Conzemius and Fedorovich, 2006a) proposed a value of 40%. (Sun and Xu, 2009) argued that the fraction should be 30%. Such a large discrepancy among different studies indicates that further investigation of the entrainment process is necessary for a better understanding of the CBL.

    The IL thickness is another key parameter in bulk models. (Pino and Vilà-Guerau De Arellano, 2008) suggested that the IL thickness is a natural length scale that characterizes the shear-produced turbulence in the TKE budget at the CBL top. (Kim et al., 2006) proposed three schemes to estimate the IL thickness based on different empirical considerations of the effect of wind shear. (Conzemius and Fedorovich, 2007) developed a scheme under the assumption that the flux Richardson number remains at 0.25 in the entrainment zone (the layer where vertical potential temperature flux is negative). However, the LES results of (Pino and Vilà-Guerau De Arellano, 2008) showed the flux Richardson number to be larger than 0.25. Therefore, the scheme of (Conzemius and Fedorovich, 2007) needs further validation and the parameterization of the IL thickness should be modified.

    In this paper, a parameterization scheme of A e for a well-developed CBL is developed in an FOM framework. As in (Conzemius and Fedorovich, 2007), the CBL is assumed to develop under the condition of the GS case. The impacts of different factors on A e are discussed. A new convective velocity scale for both buoyancy and shear effects is proposed to parameterize the IL thickness. LESs are conducted to evaluate the parameterization of A e and the performance of parameterization for the IL thickness. In a companion paper, Part II, these parameterizations are further simplified according to the characteristics of entrainment derived from the LES output, and a simple model for predicting the growth rate of the well-developed CBL is proposed and evaluated.

2. LES experiments and output
  • Twenty-six CBL cases are simulated using an LES model to provide sufficient basic data in this study. The model used is DALES (the Dutch Atmospheric Large-Eddy Simulation model), which is based on the LES code of (Nieuwstadt and Brost, 1986) and developed by researchers from Delft University, the Royal Netherlands Meteorological Institute, Wageningen University and the Max Planck Institute for Meteorology (Heus et al., 2010). The domain size in this study is 10.0× 10.0× 4.0 km3, in the x, y and z directions respectively, with grid dimensions of 256× 256× 400. (Sullivan and Patton, 2011) pointed out that the lower-order moment statistics (means, variations and fluxes) become grid-independent when the ratio of CBL height to LES filter width is larger than 56. This value corresponds to their case D, in which the mesh resolution was thought to be fine enough to characterize the entrainment process. In the present study, the ratio of CBL height to LES filter width is about 31, which is slightly larger than that of case C in (Sullivan and Patton, 2011). Their results showed that the lower-order moment statistics change slightly while the entrainment rate is obviously overestimated in case C. However, their sensitivity experiments indicated that the finer vertical resolution can improve the LES estimates of entrainment rate efficiently. Our vertical mesh resolution is closer to that in case D than case C in (Sullivan and Patton, 2011). It is expected that the mesh resolution in this study will be able to describe the turbulence statistics and entrainment process reasonably. A third-order Runge-Kutta scheme with self-adaptive time stepping is used for time integration. The surface is treated as a semi-slip boundary at the bottom, and Monin-Obukhov similarity theory is applied at the lowest model level to calculate surface momentum flux. The top 1 km of the domain is a sponge layer and periodic boundary conditions are applied at the lateral boundaries. The closure scheme for the calculation of subgrid-scale fluxes is based on the TKE method (Deardorff, 1980).

    For all cases in this study, the surface potential temperature flux is prescribed to be 0.1 K m s-1, and the potential temperature at the surface is initially set to be 300 K. The Coriolis parameter f is set to be a constant value of 10-4 s-1. Half of the cases are conducted with a large gradient of potential temperature [γθ=0.006 K m-1 (denoted as 6)], and the other half are conducted with a small gradient [γθ=0.003 K m-1 (denoted as 3)]. Two cases are free-convection cases (denoted as NS00), and the others are divided into two groups, i.e., the GC and GS groups, with the geostrophic velocity along the x-direction. In the GC group, the geostrophic velocity is prescribed with three different values: 10 m s-1, 15 m s-1 and 20 m s-1 (denoted as GC10, GC15 and GC20, respectively). In the GS group, the geostrophic velocity is zero at the surface and linearly increases with height at three different vertical gradients: 10 m s-1, 15 m s-1 and 20 m s-1 per 2 km (denoted as GS10, GS15 and GS20, respectively). Two values of surface roughness lengths, z0=0.1 m and z0=0.01 m, are used to represent the rough surface (denoted as R) and the smooth surface (denoted as S). The case name GC15R3 means that the simulation is conducted under the conditions of a constant geostrophic velocity of 15 m s-1, over a rough surface, with z0=0.1 m, and an initial potential temperature gradient of 3 K km-1. Results from the 26 cases are used to determine the empirical constants in the parameterization schemes.

    The present study is based on linear equations of potential temperature and momentum for a horizontally homogeneous CBL. With Galilean transformation, CS cases can be easily transformed to GS cases. It is expected that the parameterizations derived from GS cases should be suitable for CS cases. However, this is not true for a nonlinear system such as the three-dimensional CBL. Furthermore, non-zero surface geostrophic velocity leads to changes in surface friction velocity and mixed-layer velocity, and consequently affects velocity and fluxes at the CBL top. The changes in these variables are not linear. For this reason, four additional CS CBL experiments (C5S10S3, C5S15S3, C5S15S6 and C5S15R3) are conducted to validate the parameterizations derived from GS cases. The case name C5S15S3 means that the surface value of geostrophic velocity is 5 m s-1, the gradient of geostrophic velocity is 15 m s-1 per 2 km, the surface is smooth (z0=0.01 m), and the gradient of potential temperature is 3 K km-1.

  • The integration for each simulation covers 28 800 s, and the results for the period from 4800 s to 28 800 s are output at a time interval of 100 s. The horizontally averaged idealized profiles of potential temperature, velocity and their fluxes can be obtained from the LES output. Figures 1 and 2 depict these LES cases and idealized profiles. In this paper, the mean and the fluctuating parts of a turbulent variable are denoted with uppercase and lowercase letters; for example, Θ and θ represent the domain averaged and fluctuating parts of potential temperature. A horizontally averaged turbulent flux is denoted by an overbar; for example, \(\overline{w\theta}\) means horizontally averaged kinematic heat flux. The idealized profile of Θ is assumed to vary linearly with height in all cases. (Fedorovich, 1995) gave the idealized profile of \(\overline{w\theta}\) in the mixed layer as \begin{equation} \label{eq1} \overline{w\theta}=\left(1-\dfrac{z}{h_1}\right)\overline{w\theta}_{\rm s}+\dfrac{z}{h_1}\overline{w\theta}_1 , (1)\end{equation}

    where z is height and h1 is the CBL height, which is defined as the height at which \(\overline{w\theta}\) from the LES output reaches its minimum; \(\overline{w\theta}_\rm s\) and \(\overline{w\theta}_1\) are the kinematic heat flux at the surface and at h1 respectively. Based on this equation, A e is expressed as \begin{equation} \label{eq2} A_{\rm e}=-\frac{\overline{w\theta}_1}{\overline{w\theta}_{\rm s}}=\frac{h_1-h_0}{h_0} , (2)\end{equation} where h0 is the first zero-crossing height of the \(\overline{w\theta}\) profile. h2 is defined as the level at which \(\overline{w\theta}\) is larger than 10% of its minimum value, which is the same as in (Pino et al., 2006) and (Conzemius and Fedorovich, 2006a). The layer from h1 to h2 is the so-called IL, and its thickness is ∆ h21=h2-h1. The layer between h0 and h2 is the so-called entrainment zone, and its thickness is ∆ h20=h2 -h0. It should be noted that the definition of the IL is based on the idealized profile of Θ , while the entrainment zone is based on the profile of \(\overline{w\theta}\). Following (Pino et al., 2006), Θ1 (the potential temperature in the mixed layer) is determined from the LES Θ profile at the center of the CBL (h1/2), Θ2 is determined from the LES Θ profile towards h2, and the potential temperature jump across the IL is ∆Θ=Θ21. Following the approach of (Fedorovich, 1995), the profile of \(\overline{w\theta}\) in the IL is defined as a quadratic function of z if Θ linearly increases with height. Calculations show that the time-averaged relative error of the integral \(\overline{w\theta}\) in the IL has a maximum value of 7.8% in GC20S6. Because the errors introduced by the linear assumption are small in all of the cases, we prefer to use the linear \(\overline{w\theta}\) profile in this study. (Kim et al., 2006) pointed out that the linear profile of \(\overline{w\theta}\) gives a larger A e than the curving profile. However, the errors of A e are associated with the empirical constants in the parameterization scheme. They are obtained by a least squares fit to the LES outputs. The results show that the derived A e parameterization can perform very well (details given in section 3).

    Figure 1.  Idealized profiles (solid lines) of a CBL with constant geostrophic wind (GC). Top: horizontally averaged potential temperature $\Theta$ and its vertical flux $\overline{w\theta}$; middle: horizontally averaged $x$-component velocity $U$ and its vertical flux $\overline{uw}$; bottom: horizontally averaged $y$-component velocity $V$ and its vertical flux $\overline{vw}$. Thick dashed lines represent LES profiles,dash-dot lines represent $h_0$, $h_1$ and $h_2$. The vertical axis represents height above the surface.

    Figure 2.  As in Fig. 1, but with linearly increasing geostrophic wind (GS).

    The idealized profiles of the velocity components U and V are also assumed to be a linear function of z (Figs. 1 and 2). For the GC cases, U and V are constant in the mixed layer; thereby, U1 and V1 (velocity componets at h1) are determined from the LES velocity profiles at the center of the CBL (h1/2). For the GS and CS cases, the idealized U is constant in the mixed layer, whereas the idealized V linearly increases with height in the mixed layer. Thus, the determination of U1 in the GS and CS cases is the same as in the GC cases. Values of V at the surface and in the middle of the CBL (h1/2), which are denoted as V s and V1/2, are determined from the LES profile of V at 0.1h1 and 0.5h1, respectively. Then, V1 is given as 2V1/2-V s. For all sheared cases, the idealized U and V are assumed to be linear functions of height in the IL. U2 and V2 are determined from the LES profiles of U and V towards h2. The wind jumps across the IL are ∆ U=U2-U1 and ∆ V=V2-V1.

    Looking at the idealized profiles of momentum fluxes, it is found that \(\overline{uw}\) and \(\overline{vw}\) below h1 vary linearly with height in the GC cases. In the GS and CS cases, \(\overline{uw}\) below h1 is a linear function of height, whereas \(\overline{vw}\) below h1 is a quadratic function of z (Fedorovich, 1995). According to the idealized profiles, the momentum fluxes at h1 are written as \begin{eqnarray} \overline{uw}_1&=&\dfrac{1}{1+{\Delta h_{21}}/{h_1}}\Bigg[\dfrac{\Delta h_{21}}{2h_1}\overline{uw}_{\rm s}-\left(\Delta U- \frac{1}{2}\gamma_u\Delta h_{21}\right)w_{\rm e}+ \frac{1}{4}f\Delta h_{21}(V_{\rm s}-V_1)\Bigg]\ ({\rm for\ all\ cases}) ,\nonumber\\[1.2mm] \overline{vw}_1&=&\dfrac{1}{1+{\Delta h_{21}}/{h_1}}\left(\dfrac{\Delta h_{21}}{2h_1}\overline{vw}_{\rm s}-w_{\rm e}\Delta V\right)\ ({\rm for\ GC\ cases}) ,\quad (3)\\[-2mm]\nonumber \end{eqnarray} and \begin{eqnarray} \overline{vw}_1&=&\dfrac{1}{1+{\Delta h_{21}}/{h_1}}\Bigg\{ \dfrac{\Delta h_{21}}{h_1}\overline{vw}_{\rm s}-\dfrac{1}{2}\Delta h_{21}\dfrac{\partial V_{\rm s}}{\partial t}-\nonumber\\[1.2mm] &&\left(\Delta V+\dfrac{V_{\rm s}-V_1}{2h_1}\Delta h_{21}\right)w_{\rm e}- \dfrac{1}{2}f\Delta h_{21}(U_{\rm s}-U_{{\rm g,s}}) \Bigg\}\ ({\rm for\ GS\ and\ CS\ cases}) . (4)\end{eqnarray} The above equations are derived by vertically integrating the momentum equations (the derivations are given in Appendix A), where U g is geostrophic velocity; U g,s is the surface geostrophic velocity; γu is the gradient of U g, w e is the entrainment rate, which is defined as \(w_\rm e=\partial\langle h_1\rangle/\partial t\), \(\langle h_1\rangle\) is the least squares fit of h1, according to the relation \(h_1\propto \sqrt t\), to avoid a negative value of w e. (Fedorovich et al., 2004a) showed that h1 follows this relation in the NS case. Our LES outputs show that this relation is also effective in the sheared CBL cases (figures presented in the companion paper, Part II). Similarly, the departure of the idealized \(\overline{uw}\) and \(\overline{vw}\) profiles from the curving ones in the IL also introduces some errors. Calculations show that the maximum relative error of \(\overline{uw}\) is 8% in the sheared CBL cases. The maximum relative errors of \(\overline{vw}\) are 10.6%, 658% and 371% in the GC, GS and CS cases, respectively. However, the LES outputs show that in the GS and CS cases, the contribution of IL wind shear to the entrainment is negligibly small in the y-direction when compared with that in the x-direction (results reported in the companion paper, Part II). Therefore, the idealized profiles can characterize the IL shear effect on the entrainment reasonably.

3. Parameterization of A e and evaluation
  • The derivation begins with the TKE (E) budget in Boussinesq approximation. It is expressed as \begin{eqnarray} \label{eq3} \dfrac{\partial E}{\partial t}&=&\dfrac{g}{\Theta_0}\overline{w\theta} -\left(\overline{uw}\dfrac{\partial U}{\partial z}+\overline{vw}\dfrac{\partial V}{\partial z}\right)- \left(\dfrac{\partial\overline{wE}}{\partial z}+\dfrac{1}{\rho_0}\dfrac{\partial\overline{wp}}{\partial z}\right)-\varepsilon ,\quad\ (5)\end{eqnarray} where ρ0 is the air density (Moeng and Sullivan, 1994), p is the fluctuating part of pressure, g represents the acceleration of gravity, ε is the viscous dissipation rate of TKE. The TKE storage term on the left-hand side is small compared to other terms, except in the early stage of CBL evolution (Driedonks, 1982; Randall, 1984). The LES outputs indicate that after 4800s of integration, when the CBL is well-developed, this term is small and can be neglected. The first and second terms on the right-hand side are the production rates of TKE by buoyancy and wind shear, respectively. The third term is the vertical transport rate of TKE. This is a divergence term, and thus its integration from the surface to h2 should be zero (Moeng and Sullivan, 1994). ε is usually assumed to be proportional to its production rate (Flamant et al., 1999; Conzemius and Fedorovich, 2006b; Kim et al., 2006). Applying idealized profiles of potential temperature, velocity and their fluxes, as shown in Fig. 2, the vertical integration of the TKE budget can be written as (see derivation in Appendix B) \begin{eqnarray} &&-\dfrac{1}{2}\dfrac{g}{\Theta_0}\overline{w\theta}_1(h_1+\Delta h_{21})\nonumber\\ &&=\dfrac{1}{2}(1\!-\!\alpha_1)\dfrac{g}{\Theta_0}\overline{w\theta}_{\rm s}h_1+(1-\alpha_2)C_{\rm D}^{-1/2}u_\ast^3+\nonumber\\ &&\quad(1-\alpha_3)\left(-\dfrac{1}{2}\overline{uw}_1\Delta U-\dfrac{1}{2}\overline{vw}_1\Delta V\right)+\nonumber\\ &&\quad(1-\alpha_4)(V_1-V_{\rm s})\Bigg(-\dfrac{1}{2}\overline{vw}_{\rm s}-\dfrac{1}{2}\overline{vw}_1+\dfrac{1}{12}f\gamma_uh_1^2\Bigg) .\qquad \label{eq4} (6)\end{eqnarray} where the coefficients α123, and α4 are the proportions of the dissipation rate to the corresponding production rate and do not vary with height, $(u_\ast=\sqrt[4]{\overline{uw}_\rm s^2+\overline{vw}_\rm s^2})$ is the friction velocity, and C D=u*2/(U s2+V s2) is the surface drag coefficient. Together with the definition of the convective velocity scale, i.e., \(w_\ast^3=(g/\Theta_0)\overline{w\theta}_\rm sh_1\), the above equation yields the parameterization of A e, which is expressed as $$ A_e=A_1 \dfrac{1}{1+\frac{\Delta h_{21}}{h_1}}+A_2 \dfrac{C^{-1/2}_D}{1+\frac{\Delta h_{21}}{h_1}} \frac{u^3_*}{w^3_*}+$$ $$ Term I \ \ \ \ \ \ \ \ \ \ Term II$$ $$\ \ \ \ A_3\dfrac{(-\frac{1}{2} \overline{uw}_1 \Delta U-\frac{1}{2} \overline{vw}_1 \Delta V)}{(1+\frac{\Delta h_{21}}{h_1})w^3_*}+$$ $$\ \ \ \ \ \ \ \ \ \ Term III$$ $$\ \ \ \ \ \ \ \ \ \ A_4 \dfrac{(V_1-V_S)(-frac{1}{2} \overline{vw}_S-\frac{1}{2} \overline{vw}_1+\frac{1}{12}f\gamma_u h_1^2}{(1+\frac{\Delta h_{21}}{h_1})w^3_*}, (7)$$ $$\ \ \ \ \ \ \ \ \ \ Term IV$$ where A1=1-α1, A2=2(1-α2), A3=2(1-α3), and A4=2(1-α4). The terms on the right-hand side represent the contributions of the buoyancy (Term I), the surface layer shear (Term II), the IL shear (Term III) and the mixed layer shear (Term IV), respectively. Compared with the parameterization scheme for a GC case in (Kim et al., 2006), the only difference is that Eq. (7) has an additional term, i.e., Term IV. If the geostrophic velocity gradient vanishes, the GS case will become a GC case (that is, V s=V1), and Eq. (7) will turn out to be the same as the parameterization scheme for the GC case described in (Kim et al., 2006) [see their Eq. (22); the term \(-\overline{uw}_1\Delta U/2-\overline{vw}_1\Delta V/2\) is equivalent to that expressed in their Eq. (5B)].

    By applying stepwise regression to outputs from all of the NS, GC and GS cases, the coefficients in Eq. (7) are given as A1=0.21, A2=0.01, A3=0.86 and A4=0.70. The value of A1 is very close to the classical value of 0.2 (Stull, 1976; Fedorovich et al., 2004a, 2004b). From the definition of A2, it can be easily shown that α2=99.5%, which means that surface shear-produced TKE dissipates locally (Conzemius and Fedorovich, 2006a; Pino and Vilà-Guerau De Arellano, 2008). A3=0.86 means that the fraction of the shear-generated TKE used for the entrainment process is 43%, which is approximately the same as in (Conzemius and Fedorovich, 2006a) and supports the argument of (Sun and Xu, 2009) that the value of 1.44 for A3 proposed in (Pino et al., 2006) is overestimated. The stepwise regression also shows relatively large uncertainties in the determination of A4. However, the fourth term is too small to significantly influence the accuracy of Eq. (7) (see the results in the next section).

    Figure 3.  Horizontally and 30-min averaged vertical profiles of the total (upper panels) and the subgrid (lower panels) TKE budget for the GC15S3, GS15S3 and C5S10S3 cases. S, B, T $\varepsilon $ and $\partial E/\partial t$ represent shear production, buoyancy production, transport and dissipation rates of TKE, respectively. $\varepsilon_\rm p$ represents a linear combination of the shear and buoyancy production. Thin grey lines from bottom to top in each panel represent $h_0/h_1$, 1 and $h_2/h_1$, respectively.

    In the above derivations, ε is treated as the sum of the dissipation rates of buoyancy-and shear-produced TKE. The regression of Eq. (6) to the LES outputs shows that the dissipation rate of shear-produced TKE varies in different parts of the CBL. That is, the parameterized dissipation rate ε p can be calculated as

    \begin{eqnarray} \label{eq6} \varepsilon_{\rm p}&=&-\alpha_1\dfrac{g}{\Theta_0}\overline{w\theta}_{\rm s}\left(1-\dfrac{z}{h_1}\right)-\alpha_xS ,(8)\\ \label{eq7} \alpha_1&=&0.79 ,\nonumber\\ \alpha_x&=&\left\{ \begin{array}{l@{\quad}l} \alpha_2=0.995 & (z\le 0.1h_1)\\[1mm] \alpha_3=0.57 & (z\ge h_1)\\[1mm] \alpha_4=0.65 & (0.1h_1 \le z\le h_1) \end{array} \right., (9)\end{eqnarray} where S is the shear production rate of TKE. Figure 3 depicts the profiles of ε p and the forcing terms on the right-hand side of the TKE budget from the LES cases with weak inversion. ε p is very close to the dissipation rate ε calculated from the LES outputs, suggesting that the coefficients in Eq. (7) are reasonable. The subgrid TKE budgets in these cases are also illustrated in Fig. 3. The results indicate that the subgrid TKE is negligibly small in the IL. The cases with strong inversion have the same situation [Fig. S1 in electronic supplementary material (ESM)]. Therefore, the resolved motions dominate the TKE budget in the IL and the derived parameterizations based on the LES outputs are reasonable. Figure 4 shows that the A e estimated by Eq. (7) agrees well with that derived from the LES outputs. As presented in previous studies, the value of A e calculated from LES outputs fluctuates significantly because of the fluctuation of instantaneous LES profiles (calculations show that the spread of LES A e is reduced significantly when the LES heat flux profiles are averaged over 500 s). It is satisfactory that the value of the parameterized A e is contained within the fluctuations of the LES outputs. Therefore, the parameterization expressed as Eq. (7) can capture the characteristics of entrainment flux for a well-developed sheared CBL.

    Figure 4.  Entrainment heat flux ratios in CS cases from LES outputs (blue dots) and calculated by Eq. (6) (red dots).

    Figure 5.  Each term on the right-hand side of Eq. (7) for the parameterization of $A_\rm e$ in the GC cases. The blue dots represent Term I, the green dots represent Term II, and the red dots represent Term III.

    Figure 6.  As in Fig. 5, but for the results in the GS cases. The cyan dots represent Term IV. Note that green dots and cyan dots overlap in some cases.

  • The evolution of each term in Eq. (7) during CBL development is illustrated in Fig. 5 for the GC case, and in Fig. 6 for the GS case. It is clear that Term I and Term III are the dominant terms for CBL development. The stratification and wind shear have little influence on Term I. Its value is about 0.18 and remains almost unchanged throughout CBL development in all of the simulation cases. Term II is always very small, albeit its value differs among cases. This result suggests that Term II has little influence on entrainment. The behavior of Term III is quite different in the GC and GS cases. Term III decreases with time and increases with stratification in the GC case, whereas in the GS case it remains almost constant throughout CBL development and decreases with stratification. Term IV exists in both the GS and CS cases. Figure 6 shows that this term is negligibly small in the early stage of a developing CBL. However, it increases during CBL development. Under the condition of a weak geostrophic velocity gradient, such as in GS10, Term IV can be neglected, since it remains very small throughout the entire CBL development process. On the other hand, if the geostrophic velocity gradient and h1 are sufficiently large, Term IV becomes relatively large. For example, in GS20R3, at the end of the simulation, when h1 is about 2000 m, the contribution of Term IV to the A e is about 17%. However, this situation seldom happens in the real atmosphere because it is difficult for a geostrophic velocity gradient as large as that shown in case GS20R3 to occur.

    In the GC case, the velocity jump across the IL increases slightly with time, but the momentum flux at h1 decreases with time. Thus, their total shear production rate of TKE has a slight decreasing trend (Fig. S2 in ESM). Meanwhile, the denominator of Term III (i.e., w*3) increases remarkably during this process. This is why Term III decreases with increasing CBL depth. The LES outputs indicate that a larger gradient of potential temperature can significantly enlarge the velocity jump across the IL and slightly decrease the momentum flux at h1 (Fig. S3 in ESM). Thereby, the shear-produced TKE and A e enlarge under stronger stratification. However, this does not mean that the growth rate of the CBL under stronger background stratification increases, since the capping inversion strength also enhances, which suppresses the CBL's development (Sun and Xu, 2009). The effect of the rough surface is to enlarge the value of Term III. This is because, under such a condition, the velocity in the mixed layer is smaller and the velocity jump at the CBL top is larger, as compared to under a smooth surface condition.

    In the GS case, the value of the velocity jump across the IL increases while the momentum fluxes remain almost constant with time; thus, the shear production of TKE in the IL enhances during the CBL's development (Fig. S4 in ESM). Meanwhile, the denominator of Term III increases steadily during this process. Thus, the value of Term III does not change significantly. This implies that the shear production of TKE (i.e. \(-(\overline{uw}_1\Delta U+\overline{vw}_1\Delta V)/2\)) is approximately proportional to w*3. The reduction effect of strong stratification on Term III means that the shear production of TKE may be proportional to the inverse of γθ. A larger geostrophic velocity gradient leads to a larger momentum flux at h1 and velocity jump across the IL, and consequently a larger value of Term III. Figure 6 also shows that Term III is almost not influenced by surface roughness.

    Based on the above results, it can be deduced that the shear production rate of TKE at the CBL top can be divided into two parts. One part is proportional to w*3, γu and 1/γθ; the other part is insensitive to CBL development and γθ but sensitive to surface roughness. The former part dominates in the GS case, whereas the latter works only in the GC case. In the CS case, these two parts cooperate, but the former still dominates. Thus, Term III shows a slight decreasing trend and becomes weak with larger γθ (Fig. S5 in ESM). The expressions and meaning of these two parts are discussed in detail in the companion paper, Part II.

  • Equation (7) can be rewritten as \begin{equation} \label{eq8} A_{\rm e}=\dfrac{A_1}{1+{\Delta h_{21}}/{h_1}}\frac{w_{\rm m}^3}{w_\ast^3} ,(10) \end{equation} and \begin{eqnarray} w_{\rm m}^3&=&w_\ast^3+\frac{A_2}{A_1}C_{\rm D}^{-1/2}u_\ast^3+\dfrac{A_3}{A_1}\left(-\dfrac{1}{2}\overline{uw}_1\Delta U-\dfrac{1}{2}\overline{vw}_1\Delta V\right)+\nonumber \end{eqnarray} \begin{eqnarray} &&\dfrac{A_4}{A_1}(V_1-V_{\rm s})\left(-\dfrac{1}{2}\overline{vw}_{\rm s}-\dfrac{1}{2}\overline{vw}_1+\dfrac{1}{12}f\gamma_uh_1^2\right) , \label{eq9} (11)\end{eqnarray} where w m can be interpreted as a new characteristic convective velocity scale that includes the contributions from both the buoyancy and the wind shears in a CBL. The results in Figs. 5 and 6 suggest that the characteristic convective velocity is mainly enhanced by the IL shear. In the ZOM, Eq. (9) reduces to A e=A1w m3/w*3, which agrees with the result of (Tennekes, 1973), (Driedonks, 1982), and (Moeng and Sullivan, 1994) that the A e can be expressed as 0.2w m3/w*3.

    For the GC case, the simplified form of Eq. (11) is often used to characterize the convective velocity scale in a sheared CBL, which includes only w* and u* on the right-hand side of the equation (Tennekes, 1973; Zeman and Tennekes, 1977; Tennekes and Driedonks, 1981; Driedonks, 1982; Boers et al., 1984; Batchvarova and Gryning, 1994; Moeng and Sullivan, 1994; Pino et al., 2003). For example, (Tennekes, 1973) suggested that w m3=w*3 +12.5u*3, while (Moeng and Sullivan, 1994) proposed that w m3=w*3+5u*3. Note that the equation w m3=w*3+Bu*3 only includes the contribution of shear-produced TKE in the surface layer. Actually, it can be regarded as the simplified form of Eq. (11) by assuming that \(-\overline{uw}_1\Delta U/2-\overline{vw}_1\Delta V/2\) is approximately proportional to u*3 [the last term of Eq. (11) is zero under the GC condition). As mentioned in the previous section, this term is insensitive to CBL development and stratification strength but sensitive to surface roughness. Our LES outputs also show that the result of (Moeng and Sullivan, 1994) is a good estimate of w m. However, for the GS and CS cases, the last term on the right-hand side of Eq. (11) is relatively small and can be neglected (although it is not zero), but the third term on the right-hand side of Eq. (11) is closely related to w*3 , γu and 1/γθ. In this situation, the simplified form w m3=w*3+Bu*3 is not a good approximation of Eq. (11).

4. Parameterization of the IL thickness
  • In the FOM, the IL thickness (∆ h21=h2-h1) is a key parameter that is often used in the mixed-layer model, as described in (Pino et al., 2006) and (Conzemius and Fedorovich, 2007). According to parcel theory, after the overshooting thermal rises across the IL, its kinetic energy is transformed to potential energy. That is to say, w m2∝(g/Θ0)∆Θ∆ h21. Based on this assumption, (Kim et al., 2006) gave the parameterization of the IL thickness in the form of \begin{equation} \label{eq10} \frac{\Delta h_{21}}{h_1}=aRi^{-1}+b , (12)\end{equation} and \begin{equation} \label{eq11} Ri=\dfrac{g}{\Theta_0}\Delta\Theta h_1/{w_{\rm m}^2} , (13)\end{equation} where a and b are empirical constants, and Ri is the convective Richardson number. (Kim et al., 2006) proposed an empirical formula to characterize the turbulence velocity scale under the GC condition, expressed as \begin{equation} \label{eq12} w_{\rm m}^2=w_\ast^2+cu_\ast^2+d(\Delta U^2+\Delta V^2) , (14)\end{equation} where c and d are empirical constants. (Kim et al., 2006) provided three groups of these empirical constants. (Pino et al., 2006) used this scheme in a mixed-layer model to evaluate their parameterization of A e. By applying stepwise regression to the outputs of all the NS, GC and GS cases, cu*2 in Eq. (14) is excluded (because the existence of this term makes the significance of regression reduced), and the constants are a=0.37, b=0.13 and d=0.19. When the constant b is constrained to be zero, the regression also excludes cu*2 in Eq. (14), and the constants are a=2.46 and d=0.056. The exclusion of cu*2 suggests that the wind shear in the surface layer has little effect on entrainment, which agrees with the result in section 3.1 that the surface shear-produced TKE dissipates locally. This scheme with different constants is denoted as KP1 and KP2, respectively (Table 1).

    In the previous section, a new convective velocity scale is proposed. We expect that it is appropriate for the estimation of IL thickness, and thus we use Eq. (12) but replace Eq. (14) with Eq. (11). By a least squares fit to the LES outputs of NS, GC and GS cases, the empirical constants a and b are determined to be 0.70 and 0.14, respectively. This scheme is denoted as LS1 (Table 1).

    The parameterization of the IL thickness is usually evaluated by field observations (e.g., Boers et al., 1984), experiments (e.g., Deardorff et al., 1980) and LESs (e.g., Fedorovich et al., 2004a). Uncertainties in h2 and Θ2 determined from LES outputs will subsequently result in biases in the calculation of ∆ h21 and ∆Θ. A positive bias of ∆ h21 (as well as ∆Θ) causes a negative bias of Ri-1, resulting in low correlation between ∆ h21/h1 and Ri-1 (Sun et al., 2005). For this reason, another Richardson number, RiN, which is based on the buoyancy frequency \(\sqrt{(g/\Theta_0)\gamma_\theta}\) in the free atmosphere, is used in (Fedorovich et al., 2004a). As proposed by (Stull, 1973), the time taken for the rising thermal to penetrate into the free atmosphere should be related to the buoyancy frequency. That is, \(\Delta h_21/w_\rm m\propto 1/\sqrt{(g/\Theta_0)\gamma_\theta}\). The parameterization scheme can be written as \begin{equation} \label{eq13} \frac{\Delta h_{21}}{h_1}=a_N Ri_N^{-\frac{1}{2}}+b_N , (15)\end{equation} and \begin{equation} \label{eq14} Ri_{N}=\frac{g}{\Theta_0}\frac{\gamma_\theta h_1^2}{w_{\rm m}^2} , (16)\end{equation} where aN and bN are empirical constants. For comparison purposes, we use Eq. (15) to parameterize the IL thickness and employ Eq. (11) to characterize the turbulent velocity in Eq. (16). The least squares fit to our LES outputs of NS, GC and GS cases yields aN=1.30 and bN=0.00. This scheme is denoted by LS2 (Table 1). It should be pointed out that if ∆Θ∝γθ∆ h21, Eqs. (12) and (15) should be identical. The LES outputs indicate that ∆Θ≈1.88γθ∆ h21 (Fig. S6 in ESM). The LS1 and LS2 schemes are actually equivalent; we denote them as LS1 and LS2 simply because they use different variables and have different constants.

    (Mahrt and Lenschow, 1976) and (Conzemius and Fedorovich, 2006a) suggested that a balance exists between the shear production and buoyancy destruction of TKE in the entrainment zone, which can be described by the flux Richardson number or gradient Richardson number. (Conzemius and Fedorovich, 2007) set the bulk gradient Richardson number (Ri b) in the IL to be a critical value of 0.15. The IL thickness is then given by \begin{equation} \label{eq15} \Delta h_{21}=Ri_{\rm b}\frac{\Delta U^2+\Delta V^2}{\frac{g}{\Theta _0}\Delta \Theta} . (17)\end{equation} The constant Ri b is found to be 0.23 by a least squares fit to our LES outputs of GC and GS cases. This scheme is denoted as CF (Table 1).

    In order to evaluate the performance of the above five parameterization schemes for the IL thickness, the relative errors are calculated and illustrated in Fig. 7. The relative error (Err) is defined as \begin{equation} \label{eq16} {\rm Err}=\dfrac{1}{n}\sum\left|\dfrac{\Delta h_{21{\rm ,p}}}{\langle{\Delta h_{{\rm 21,LES}}}\rangle}-1\right| , (18)\end{equation} where ∆ h21 ,p is the inversion layer thickness predicted by each parameterization scheme. As mentioned above, ∆ h21 is determined from the instantaneous LES profile of \(\overline{w\theta}\). This method can result in large errors that completely conceal differences between different parameterizations. In order to reduce errors, an equal weighted nine-point moving average operator is applied to ∆ h21, and the result is denoted by \(\langle\Delta h_\rm 21,LES\rangle\). The CF scheme has the largest error in most cases. Further analysis shows that the bulk gradient Richardson number varies from 0.17 to 2.31 in different cases, whereas only in a few cases is the bulk gradient Richardson number very close to 0.23. This is why the CF scheme has relatively large errors in most cases. The two KP schemes apply well, although KP2 has slightly lager errors than KP1. The new empirical constants significantly improve the performance of the KP schemes (Fig. S7 in ESM), implying that the original ones proposed by (Kim et al., 2006) and (Pino et al., 2006) are not very representative because they were derived from only a few LES cases. LS1 is not improved in comparison with the two KP schemes; and the reason is probably that they all use the convective Richardson number (Ri), as discussed previously. However, LS2 has the best performance, and the errors are less than 20% in all cases, suggesting that the RiN scheme is more suitable for characterizing the IL thickness.

    Figure 7.  Relative errors of the predicted capping IL thickness against the LES results. The parameterization schemes \small \parbox[t]14cmare listed in Table 1.

    The most recent study on shear-free entrainment by means of direct numerical simulation (Garcia and Mellado, 2014) suggests a two-layer model might be appropriate for studying the entrainment zone. The upper sub-layer thickness of the entrainment zone (δ) is defined based on the maximum potential temperature gradient in (Garcia and Mellado, 2014) [see Fig. 5 and Eq. (20) in their paper]. It is different to the IL thickness (∆ h21) defined in this study. The bottom of δ is located at the level of the maximum gradient of potential temperature that is higher than the bottom of ∆ h21, while the top of δ is lower than the top of ∆ h21. Thus, the upper sub-layer of the entrainment zone defined in (Garcia and Mellado, 2014) is part of the IL defined in this study. Their results show that the upper sub-layer thickness (δ) of the entrainment zone is actually the mean penetration depth of an overshooting thermal, which is directly affected by the background stratification N2 \(N=\sqrt{(g/\Theta_0)\gamma_\theta}\). The following relation is obtained [their Eq. (24)]: \begin{equation} \label{eq17} \delta=c_\delta({w_\ast}/N) , (19)\end{equation} where cδ=0.55 is the coefficient obtained from the direct numerical simulation results. In fact, the LS2 scheme is equivalent to ∆ h21=aN(w m/N). Following the method in (Garcia and Mellado, 2014), δ is determined from our LES Θ profile. The LES results show that ∆ h21/δ=2.44, which is quite close to 2.36, the ratio of aN to cδ. This implies that the LS2 scheme is similar to Eq. (18) in a shear-free CBL, because both ∆ h21 and δ are the overshooting distances of thermals rising in the stably stratified environment. The difference between the LS2 scheme and Eq. (18) is attributed to different definations of ∆ h21 and δ. Thus, in a shear-free CBL, the IL thickness is dominated by overshooting thermals. However, in a sheared CBL, the effect of wind shear on the IL thickness is also important. Our results suggest that w m is suitable for characterizing the joint effects of thermal overshooting and wind shear on IL thickness.

5. Conclusion and discussion
  • In an FOM framework, the parameterization of A e at the top of a well-developed CBL under the GS condition is derived by vertically integrating the TKE budget. Compared to the parameterization scheme under the GC condition proposed by (Kim et al., 2006) and (Pino et al., 2006), our scheme includes an additional term that represents the contribution of shear-produced TKE in the mixed layer. When the geostrophic velocity gradient becomes zero, the parameterization scheme turns out to be the one under the GC condition. This scheme is also valid for the CS case. Thus, the new parameterization developed in the present study is appropriate for entrainment approximation in a well-developed CBL under different linearly sheared geostrophic velocity conditions.

    The new parameterization contains four terms representing the effects of the buoyancy, surface layer shear, IL shear and mixed layer shear, respectively. The buoyancy and IL shear are the dominant terms among these four terms. The LES outputs indicate that the shear-produced TKE in the surface layer dissipates locally, and 43% of the shear-produced TKE at the CBL top contributes to the entrainment, which is approximately the same as the results in (Conzemius and Fedorovich, 2006a).

    A new convective velocity scale in the sheared CBL is proposed. It includes the contributions of buoyancy and wind shears. In the GC cases, the convective velocity scale is equivalent to the simplified form proposed by (Moeng and Sullivan, 1994), in which the effect of wind shear in the entire CBL can be approximately represented by the friction velocity. LES outputs show that the direct contribution of surface shear to the entrainment is relatively small. However, as pointed out by (Conzemius and Fedorovich, 2006a), the surface shear has an indirect effect on entrainment by slowing the flow in the CBL interior and inducing shear at the CBL top, and it is the IL shear that enhances the entrainment. Apparently, the contribution of IL shear to the entrainment process has been considered in the simplified formula by the friction velocity. However, note that in the GS and CS cases, the simplified form of the convective velocity scale is not valid because the shear-produced TKE in the IL is mainly related to w*, the geostrophic velocity gradient and stratification strength, rather than the friction velocity.

    The parameterization schemes of the IL thickness proposed in previous studies are evaluated by the LES outputs. The schemes suggested by (Kim et al., 2006) apply well when the new empirical constants are used. The empirical constants are derived by the stepwise regression to our LES outputs, which excludes the term representing the surface shear. This result supports that buoyancy and IL shear are the dominant factors of sheared entrainment. The parameterization scheme proposed by (Conzemius and Fedorovich, 2007) can only perform well in a few cases, because the bulk Richardson number varies widely in different cases. However, the buoyancy Richardson number approach (the RiN scheme), combined with the new convective velocity scale, can characterize the IL thickness well in all cases.

    The A e and IL thickness are important parameters in the mixed layer model. Our aim is to obtain a simplified scheme that can predict the developing process of a sheared CBL well. The parameterization scheme developed in this study represents our initial efforts to achieve this goal. The simplified model is further explored and discussed in the companion paper, Part II.

    Finally, it is worth noting that the parameterizations proposed in this study may only be applicable for CBLs under special conditions. In the derivations we neglect the storage term in the TKE budget and only consider the linearly sheared geostrophic velocity and stable background stratification. However, when the CBL is in its early developing stage, the storage term in the TKE budget is not negligibly small, and the entrainment process may exhibit different characteristics. There often exists a residual layer in the real atmosphere. When the CBL is growing through this layer, the entrainment process is different to that of the stratified free atmosphere above the CBL. In addition, in the real atmosphere the geostrophic wind may not vary linearly with height. The applicability of the parameterization schemes under these conditions is not well investigated and needs further evaluation. These problems will be investigated in future work.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return