Advanced Search
Article Contents

Parameterization of Sheared Entrainment in a Well-developed CBL. Part II: A Simple Model for Predicting the Growth Rate of the CBL


doi: 10.1007/s00376-016-5209-9

  • Following the parameterization of sheared entrainment obtained in the companion paper, Liu et al.(2016), the present study aims to further investigate the characteristics of entrainment, and develop a simple model for predicting the growth rate of a well-developed and sheared CBL. The relative stratification, defined as the ratio of the stratification in the free atmosphere to that in the entrainment zone, is found to be a function of entrainment flux ratio (A e). This leads to a simple expression of the entrainment rate, in which A e needs to be parameterized. According to the results in Liu et al.(2016), A e can be simply expressed as the ratio of the convective velocity scale in the sheared CBL to that in the shear-free CBL. The parameterization of the convective velocity scale in the sheared CBL is obtained by analytically solving the bulk model with several assumptions and approximations. Results indicate that the entrainment process is influenced by the dynamic effect, the interaction between mean shear and environmental stratification, and one other term that includes the Coriolis effect. These three parameterizations constitute a simple model for predicting the growth rate of a well-developed and sheared CBL. This model is validated by outputs of LESs, and the results show that it performs satisfactorily. Compared with bulk models, this model does not need to solve a set of equations for the CBL. It is more convenient to apply in numerical models.
  • 加载中
  • Ball F. K., 1960: Control of inversion height by surface heating. Quart. J. Roy. Meteor. Soc., 86, 483- 494.10.1002/qj.497086370057e07d96576d465f8440da8406fac3186http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49708637005%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/qj.49708637005/fullABSTRACT Viscous dissipation is inadequate to account for the destruction of all the thermal turbulence generated by upward transfer of heat in a deep convection layer. It is suggested that the surplus turbulent kinetic energy is destroyed by buoyancy forces in a region of downward transfer of heat in the upper part of the convection layer. This process is associated with mass transfer downwards through the convection (or subsidence) inversion and also with an upward movement of the inversion, though the latter may be overcome by sufficiently strong subsidence. A simple theory of the process is developed and the results deduced thereby agree well with observations of the diurnal variation of inversion height in Central Australia.
    Breuer H., F. Ács, Á. Horváth, P. Németh, and K. Rajkai, 2014: Diurnal course analysis of the WRF-simulated and observation-based planetary boundary layer height. Adv. Sci. Res., 11, 83- 88.10.5194/asr-11-83-20141d21567a82623bb86779338e922eb225http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2014AdSR...11...83Bhttp://adsabs.harvard.edu/abs/2014AdSR...11...83BWeather Research and Forecasting (WRF) single-column model simulations were performed in the late summer of 2012 in order to analyse the diurnal changes of the planetary boundary layer (PBL). Five PBL schemes were tested with the WRF. From the radiometer and wind-profiler measurements at one station, derived PBL heights were also compared to the simulations. The weather conditions during the measurement period proved to be dry; the soil moisture was below wilting point 85 percent of the time. Results show that (1) simulation-based PBL heights are overestimated by about 500-1000 m with respect to the observation-based PBL heights, and (2) PBL height deviations between different observation-based methods (around 700 m in the midday) are comparable with PBL height deviations between different model schemes used in the WRF single-column model. The causes of the deviations are also discussed. It is shown that in the estimation of the PBL height the relevance of the atmospheric profiles could be as important as the relevance of the estimation principles.
    Canut G., M. Lothon, F. Sad, and F. Lohou, 2010: Observation of entrainment at the interface between monsoon flow and the Saharan Air Layer. Quart. J. Roy. Meteor. Soc., 136, 34- 46.10.1002/qj.471c2561a268d1c003aa0bc2e6f055bf9d9http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.471%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1002/qj.471/pdfNot Available
    Conzemius R., 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 senses 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.
    Conzemius R. J., E. Fedorovich, 2006: 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/JAS3696.1a0ed3d2267609603cc6daf82ae61a4d1http%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D2216928http://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...
    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.
    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., R. Conzemius, and D. Mironov, 2004: 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.
    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
    Hong S.-Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an explicit treatment of entrainment processes. Mon. Wea. Rev., 134, 2318- 2341.10.1175/MWR3199.179f98ee85a3853a6bfee0ec84e90c901http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2006MWRv..134.2318Hhttp://adsabs.harvard.edu/abs/2006MWRv..134.2318HThis paper proposes a revised vertical diffusion package with a nonlocal turbulent mixing coefficient in the planetary boundary layer (PBL). Based on the study of Noh et al. and accumulated results of the behavior of the Hong and Pan algorithm, a revised vertical diffusion algorithm that is suitable for weather forecasting and climate prediction models is developed. The major ingredient of the revision is the inclusion of an explicit treatment of entrainment processes at the top of the PBL. The new diffusion package is called the Yonsei University PBL (YSU PBL). In a one-dimensional offline test framework, the revised scheme is found to improve several features compared with the Hong and Pan implementation. The YSU PBL increases boundary layer mixing in the thermally induced free convection regime and decreases it in the mechanically induced forced convection regime, which alleviates the well-known problems in the Medium-Range Forecast (MRF) PBL. Excessive mixing in the mixed layer in the presence of strong winds is resolved. Overly rapid growth of the PBL in the case of the Hong and Pan is also rectified. The scheme has been successfully implemented in the Weather Research and Forecast model producing a more realistic structure of the PBL and its development. In a case study of a frontal tornado outbreak, it is found that some systematic biases of the large-scale features such as an afternoon cold bias at 850 hPa in the MRF PBL are resolved. Consequently, the new scheme does a better job in reproducing the convective inhibition. Because the convective inhibition is accurately predicted, widespread light precipitation ahead of a front, in the case of the MRF PBL, is reduced. In the frontal region, the YSU PBL scheme improves some characteristics, such as a double line of intense convection. This is because the boundary layer from the YSU PBL scheme remains less diluted by entrainment leaving more fuel for severe convection when the front triggers it.
    Jonker H. J. J., M. Van Reeuwijk, P. P. Sullivan, and E. G. Patton, 2013: On the scaling of shear-driven entrainment: A DNS study. J. Fluid Mech., 732, 150- 165.10.1017/jfm.2013.39471a79c247de537a1f9a49aef0d27f83dhttp%3A%2F%2Fjournals.cambridge.org%2Farticle_S0022112013003947http://journals.cambridge.org/article_S0022112013003947The deepening of a shear-driven turbulent layer penetrating into a stably stratified quiescent layer is studied using direct numerical simulation (DNS). The simulation design mimics the classical laboratory experiments by Kato & Phillips (J. Fluid Mech., vol.0237, 1969, pp.02643–655) in that it starts with linear stratification and applies a constant shear stress at the lower boundary, but avoids sidewall and rotation effects inherent in the original experiment. It is found that the layers universally deepen as a function of the square root of time, independent of the initial stratification and the Reynolds number of the simulations, provided that the Reynolds number is large enough. Consistent with this finding, the dimensionless entrainment velocity varies with the bulk Richardson number as \$R{i}^{- 1/ 2} \$. In addition, it is observed that all cases evolve in a self-similar fashion. A self-similarity analysis of the conservation equations shows that only a square root growth law is consistent with self-similar behaviour.
    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.
    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.
    Lilly D. K., 1968: Models of cloud-topped mixed layers under a strong inversion. Quart. J. Roy. Meteor. Soc., 94, 292- 309.10.1002/qj.49709440106d42439f9eb80054e4fdaa078321acde6http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49709440106%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/qj.49709440106/abstractAbstract Theoretical models are constructed with the aim of relating, explaining and predicting features of a radiatively active turbulent cloud layer over the sea and under a strong subsidence inversion. Both dry aerosol clouds (no phase change) and wet clouds (with a phase change and latent heat exchanges) are considered. For the wet cloud case an important element of the theory is the requirement that the wet-bulb potential temperature must increase upwards in the inversion. For both cases entrainment of the upper warm air is hypothesized to lie between upper and lower limits determined from the turbulent energy budget. The dry cloud case is solved for both steady state and transient results, with only the transient behaviour depending on the entrainment hypothesis. Only steady state solutions are presented for the more complex wet cloud case and these differ somewhat for the maximum and minimum entrainment limits. Observational data from Oakland, California are used for comparison with those steady state solutions, with results indicating the essential validity of the approach. Detailed comparisons, especially for determination of the most correct entrainment rate, are hampered both by inadequate measurement of the inversion properties and by uncertainties in the net radiation flux leaving the cloud top. Computations of the latter suggest that several presently used radiation models are still in serious disagreement, at least for application to downward flux under an inversion. It is suggested that the present theory provides a partial explanation of the origin of the trade wind inversion.
    Liu P., J. Sun, and L. Shen, 2016: Parameterization of sheared entrainment in the well-developed convective boundary layer. Part I: Evaluation of the scheme through large-eddy simulations. Adv. Atmos. Sci.,10.1007/s00376-016-5208-x.a4480e655d303e9ee34b28695cc3be93http%3A%2F%2F159.226.119.58%2Faas%2FEN%2F10.1007%2Fs00376-016-5208-xhttp://159.226.119.58/aas/EN/10.1007/s00376-016-5208-x
    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.
    Pino D., J. V.-G. 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. V.-G. 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. J. Appl. Meteor. Climatol., 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.
    Shin H., S.-Y. Hong, 2011: Intercomparison of planetary boundary-layer parametrizations in the WRF model for a single day from CASES-99. Bound.-Layer Meteor., 139, 261- 281.10.1007/s10546-010-9583-z7fbe68bb0104214bd28209ed9c944445http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FADS%3Fid%3D2011BoLMe.139..261Shttp://onlinelibrary.wiley.com/resolve/reference/ADS?id=2011BoLMe.139..261SThis study compares five planetary boundary-layer (PBL) parametrizations in the Weather Research and Forecasting (WRF) numerical model for a single day from the Cooperative Atmosphere-Surface Exchange Study (CASES-99) field program. The five schemes include two first-order closure schemes—the Yonsei University (YSU) PBL and Asymmetric Convective Model version 2 (ACM2), and three turbulent kinetic energy (TKE) closure schemes—the Mellor–Yamada–Janji04 (MYJ), quasi-normal scale elimination (QNSE), and Bougeault–Lacarrére (BouLac) PBL. The comparison results reveal that discrepancies among thermodynamic surface variables from different schemes are large at daytime, while the variables converge at nighttime with large deviations from those observed. On the other hand, wind components are more divergent at nighttime with significant biases. Regarding PBL structures, a non-local scheme with the entrainment flux proportional to the surface flux is favourable in unstable conditions. In stable conditions, the local TKE closure schemes show better performance. The sensitivity of simulated variables to surface-layer parametrizations is also investigated to assess relative contributions of the surface-layer parametrizations to typical features of each PBL scheme. In the surface layer, temperature and moisture are more strongly influenced by surface-layer formulations than by PBL mixing algorithms in both convective and stable regimes, while wind speed depends on vertical diffusion formulations in the convective regime. Regarding PBL structures, surface-layer formulations only contribute to near-surface variability and then PBL mean properties, whereas shapes of the profiles are determined by PBL mixing algorithms.
    Skamarock W.C., Coauthors, 2008: A description of the advanced research WRF version 3. NCAR Tech. Note TN-475+STR, 77 pp.
    Sühring M., B. Maronga, F. Herbort, S. Raasch, 2014: On the effect of surface heat-flux heterogeneities on the mixed-layer-top entrainment. Bound.-Layer Meteor., 151, 531- 556.10.1007/s10546-014-9913-7ba6410a7db0fd9a731bddba2320cf910http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2014BoLMe.151..531Shttp://adsabs.harvard.edu/abs/2014BoLMe.151..531SWe used a set of large-eddy simulations to investigate the effect of one-dimensional stripe-like surface heat-flux heterogeneities on mixed-layer top entrainment. The profiles of sensible heat flux and the temporal evolution of the boundary-layer depth revealed decreased entrainment for small heat-flux amplitudes and increased entrainment for large heat-flux amplitudes, compared to the homogeneously-heated mixed layer. For large heat-flux amplitudes the largest entrainment was observed for patch sizes in the order of the boundary-layer depth, while for significantly smaller or larger patch sizes entrainment was similar as in the homogeneous case. In order to understand the underlying physics of this impact, a new approach was developed to infer local information on entrainment by means of the local flux divergence. We found an entrainment maximum over the centre of the stronger heated surface patch, where thermal energy is accumulated by the secondary circulation (SC) that was induced by the surface heterogeneity. Furthermore, we observed an entrainment maximum over the less heated patch as well, which we suppose is to be linked to the SC-induced horizontal flow convergence at the top of the convective boundary layer (CBL). For small heat-flux amplitudes a counteracting effect dominates that decreases entrainment, which we suppose is the horizontal advection of cold air in the lower, and warm air in the upper, CBL by the SC, stabilizing the CBL and thus weakening thermal convection. Moreover, we found that a mean wind can reduce the heterogeneity-induced impact on entrainment. If the flow is aligned perpendicular to the border between the differentially-heated patches, the SC and thus its impact on entrainment vanishes due to increased horizontal mixing, even for moderate wind speeds. However, if the flow is directed parallel to the border between the differentially-heated patches, the SC and thus its impact on entrainment persists.
    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.
    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.
    Xie B., J. C. H. Fung, A. Chan, and A. Lau, 2012: Evaluation of nonlocal and local planetary boundary layer schemes in the WRF model. J. Geophys. Res., 117, D12103.10.1029/2011JD017080d0a74580a08442071f7f056c590c6212http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2011JD017080%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2011JD017080/fullAbstract Top of page Abstract 1.Introduction 2.Model Setup and Configurations 3.Brief Descriptions of PBL Schemes and Surface Layer Schemes 4.Results and Discussions 5.Conclusions Acknowledgments References Supporting Information [1] A realistic reproduction of planetary boundary layer (PBL) structure and its evolution is critical to numerical simulation of regional meteorology and air quality. Conversely, insufficient realism in the simulated physical properties often leads to degraded meteorological and air quality prognostic skills. This study employed the Weather Research and Forecasting model (WRF) to evaluate model performance and to quantify meteorological prediction differences produced by four widely used PBL schemes. Evaluated were two nonlocal PBL schemes, YSU and ACM2, and two local PBL schemes, MYJ and Boulac. The model grid comprised four nested domains at horizontal resolutions of 27km, 9km, 3km and 1km respectively. Simulated surface variables 2m temperature and 10m wind at 1km resolution were compared to measurements collected in Hong Kong. A detailed analysis of land-atmosphere energy balance explicates heat flux and temperature variability among the PBL schemes. Differences in vertical profiles of horizontal velocity, potential temperature, bulk Richardson number and water vapor mixing ratio were examined. Diagnosed PBL heights, estimated by scheme specific formulations, exhibited the large intrascheme variance. To eliminate formulation dependence in PBL height estimation, lidar measurements and a unified diagnosis were jointly used to reanalyze PBL heights. The diagnosis showed that local PBL schemes produced shallower PBL heights than those of nonlocal PBL schemes. It is reasonable to infer that WRF, coupled with the ACM2 PBL physics option can be a viable producer of meteorological forcing to regional air quality modeling in the Pearl River Delta (PRD) Region.
  • [1] Peng LIU, Jianning SUN, Lidu SHEN, 2016: Parameterization of Sheared Entrainment in a Well-Developed CBL. Part I: Evaluation of the Scheme through Large-Eddy Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 1171-1184.  doi: 10.1007/s00376-016-5208-x
    [2] 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
    [3] Shi LUO, Chunsong LU, Yangang LIU, Yaohui LI, Wenhua GAO, Yujun QIU, Xiaoqi XU, Junjun LI, Lei ZHU, Yuan WANG, Junjie WU, Xinlin YANG, 2022: Relationships between Cloud Droplet Spectral Relative Dispersion and Entrainment Rate and Their Impacting Factors, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 2087-2106.  doi: 10.1007/s00376-022-1419-5
    [4] 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
    [5] 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
    [6] SUN Jianning, JIANG Weimei, CHEN Ziyun, YUAN Renmin, 2005: Parameterization for the Depth of the Entrainment Zone above the Convectively Mixed Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 114-121.  doi: 10.1007/BF02930874
    [7] 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
    [8] 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
    [9] 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
    [10] 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
    [11] 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
    [12] Changhai LIU, 2005: A Numerical Investigation of a Slow-Moving Convective Line in a Weakly Sheared Environment, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 625-639.  doi: 10.1007/BF02918706
    [13] 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
    [14] ZENG Qingcun, CHENG Xueling, HU Fei, PENG Zhen, 2010: Gustiness and Coherent Structure of Strong Winds and Their Role in Dust Emission and Entrainment, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 1-13.  doi: 10.1007/s00376-009-8207-3
    [15] 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
    [16] Peng Jiayi, Wu Rongsheng, Wang Yuan, 2002: Initiation Mechanism of Meso-β Scale Convective Systems, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 870-884.  doi: 10.1007/s00376-002-0052-6
    [17] Fan Beifen, Ye Jiadong, William R. Cotton, Gregory J. Tripoli, 1990: Numerical Simulation of Microphysics in Meso-β-Scale Convective Cloud System Associated with a Mesoscale Convective Complex, ADVANCES IN ATMOSPHERIC SCIENCES, 7, 154-170.  doi: 10.1007/BF02919153
    [18] Zhizhen XU, Jing CHEN, Mu MU, Guokun DAI, Yanan MA, 2022: A Nonlinear Representation of Model Uncertainty in a Convective-Scale Ensemble Prediction System, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1432-1450.  doi: 10.1007/s00376-022-1341-x
    [19] Xiaoran ZHUANG, Jinzhong MIN, Liu ZHANG, Shizhang WANG, Naigeng WU, Haonan ZHU, 2020: Insights into Convective-scale Predictability in East China: Error Growth Dynamics and Associated Impact on Precipitation of Warm-Season Convective Events, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 893-911.  doi: 10.1007/s00376-020-9269-5
    [20] Yidan SI, Shenshen LI, Liangfu CHEN, Chao YU, Zifeng WANG, Yang WANG, Hongmei WANG, 2018: Validation and Spatiotemporal Distribution of GEOS-5-Based Planetary Boundary Layer Height and Relative Humidity in China, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 479-492.  doi: 10.1007/s00376-017-6275-3

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 10 October 2015
Manuscript revised: 07 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 II: A Simple Model for Predicting the Growth Rate of the CBL

  • 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: Following the parameterization of sheared entrainment obtained in the companion paper, Liu et al.(2016), the present study aims to further investigate the characteristics of entrainment, and develop a simple model for predicting the growth rate of a well-developed and sheared CBL. The relative stratification, defined as the ratio of the stratification in the free atmosphere to that in the entrainment zone, is found to be a function of entrainment flux ratio (A e). This leads to a simple expression of the entrainment rate, in which A e needs to be parameterized. According to the results in Liu et al.(2016), A e can be simply expressed as the ratio of the convective velocity scale in the sheared CBL to that in the shear-free CBL. The parameterization of the convective velocity scale in the sheared CBL is obtained by analytically solving the bulk model with several assumptions and approximations. Results indicate that the entrainment process is influenced by the dynamic effect, the interaction between mean shear and environmental stratification, and one other term that includes the Coriolis effect. These three parameterizations constitute a simple model for predicting the growth rate of a well-developed and sheared CBL. This model is validated by outputs of LESs, and the results show that it performs satisfactorily. Compared with bulk models, this model does not need to solve a set of equations for the CBL. It is more convenient to apply in numerical models.

1. Introduction
  • The depth of the CBL is an important parameter in air pollution and NWP models. In these models, the CBL height is often diagnosed from the critical bulk Richardson number or TKE profile (Skamarock et al., 2008). These diagnosed CBL heights exhibit large intra-scheme variances (Shin and Hong, 2011; Xie et al., 2012; Breuer et al., 2014). Thus, a proper method to estimate the CBL height is important for numerical models. The growth rate of a CBL is actually the entrainment rate when there is no background vertical velocity. In the 1960s, (Ball, 1960) and (Lilly, 1968) proposed a bulk model framework to describe the evolution of the CBL. Since then, the bulk model approach has been widely used to predict the CBL entrainment rate. For a sheared CBL, the results of LESs in (Kim et al., 2003) showed that the turbulence in the entrainment zone enhances due to the break of Kelvin-Helmholtz billows at the upper edge of the entrainment zone. As a result, the entrainment process accelerates and the IL deepens. In order to adequately capture the entrainment process in a sheared CBL, at least the first-order model (FOM) is needed (Kim et al., 2006; Conzemius and Fedorovich, 2007), and the set of equations for the CBL should be solved. Unfortunately, the bulk model is complex and difficult to apply in numerical models since it includes too many unknown variables. Therefore, it is imperative to develop a simple model for predicting the growth rate of a sheared CBL.

    The entrainment rate is associated with the entrainment flux ratio A e, which is defined as the ratio of heat flux at the CBL top to that at the surface. (Kim et al., 2006) proposed a parameterization of A e for the sheared CBL in the FOM framework. They only considered the sheared CBL under the condition of height-constant geostrophic velocity (GC). However, the LES results in (Conzemius and Fedorovich, 2006) indicated that the entrainment process has different characteristics under the condition of sheared geostrophic velocity with a zero value at the surface (GS). Following the derivation in (Kim et al., 2006), (Liu et al., 2016) developed a parameterization of A e for a well-developed and sheared CBL. This scheme takes into account the buoyancy effect and the shear effect in the surface layer, the mixed layer and the IL. The shear effect in the IL is represented by local momentum fluxes and velocity jumps at the CBL top, which is similar to that in (Pino and De Arellano, 2008). It still includes many variables and thereby cannot be applied directly. However, if the relations between the entrainment variables (e.g., the entrainment momentum fluxes, the potential temperature jump, and the velocity jumps) can be described by external parameters (such as the background stratification, the geostrophic wind gradient, and the surface friction velocity), the parameterizations of A e and the entrainment rate can be simplified and become applicable. For this reason, the present study aims to develop a simple model appropriate for predicting the growth rate of the sheared CBL by using external parameters.

    The parameterization of the entrainment rate suggested by (Sun and Xu, 2009) and the A e scheme proposed in (Liu et al., 2016) are simplified with some assumptions in this study. The LES outputs obtained in (Liu et al., 2016) are used for analysis and evaluation. The paper is organized as follows: Section 2 discusses the theoretical considerations in the FOM framework, and a simplified parameterization for the entrainment rate is proposed. In section 3, the characteristics of the sheared entrainment are analyzed based on the LES results, and the proposed simple parameterization is verified using the LES outputs. Conclusions and discussion are given in the final section.

2. Simplification of the parameterization of entrainment rate
  • In this study, Θ, U and V represent horizontally averaged potential temperature and velocity components, while θ, u and v represent the fluctuation parts of potential temperature and the velocity components. \(\overline{w\theta}\), \(\overline{uw}\) and \(\overline{vw}\) represent horizontally averaged vertical fluxes of potential temperature and velocity components. γθ and γu represent the vertical gradients of the initial potential temperature and the geostrophic velocity in the x-direction. γθ and γu are external parameters, which are assumed to remain unchanged during CBL development. Figure 1 depicts the idealized profiles of Θ, U and V, and their fluxes, in a well-developed sheared CBL. The CBL height h1 is defined as the level at which \(\overline{w\theta}\) reaches its minimum. h0 is the first zero-crossing height of the \(\overline{w\theta}\) profile. h2 is defined as the level toward which \(\overline{w\theta}\) becomes larger than 10% of its minimum. The layer between h1 and h2 is the inversion layer (IL), in which the idealized Θ increases with height. The IL thickness is ∆ h21=h2-h1. The layer between h0 and h2 is the entrainment zone, in which \(\overline{w\theta}\) is negative. The entrainment zone thickness is ∆ h20=h2-h0. Θ1, U1 and V1 are values of Θ, U and V at h1. Θ2, U2 and V2 are values of Θ, U and V at h2. The potential temperature jump is ∆Θ=Θ21, and the two components of the velocity jump are ∆ U=U2-U1 and ∆ V=V2-V1. The A e is defined as \(A_e=-\overline{w\theta}_1/\overline{w\theta}_s =h_1/h_0-1\). \(\overline{uw}_1\) and \(\overline{vw}_1\) are the momentum fluxes at h1, which are obtained by integrating the momentum equations from the surface to h2. For further details, please refer to (Liu et al., 2016).

    Figure 1.  Idealized profiles of horizontally averaged potential temperature $\Theta$, velocity components $U$ and $V$, and their vertical fluxes $\overline{w\theta}$, $\overline{uw}$ and $\overline{vw}$ in the GS cases. Dash-dot lines represent $h_0$, $h_1$ and $h_2$, and dotted lines represent zero.

  • In the FOM framework and under the GC condition, the parameterization of the entrainment rate (w e) and A e proposed by (Kim et al., 2006) and evaluated by (Pino et al., 2006) (hereafter KP06) can be expressed as \begin{equation} \label{eq1} w_{{e,KP}}=\dfrac{\partial h_1}{\partial t}=\dfrac{\Delta h_{21}+(2h_1+\Delta h_{21})A_{{e,KP}}}{h_1(2\Delta\Theta-\gamma_\theta\Delta h_{21})} \overline{w\theta}_{s} (1)\end{equation} and \begin{eqnarray} \label{eq2} A_{{e,KP}}&=&\dfrac{-\overline{w\theta}_1}{\overline{w\theta}_{s}}=A_{1{,KP}}\dfrac{1}{1+\frac{\Delta h_{21}}{h_1}}+ A_{2,{KP}}\dfrac{1}{1+\frac{\Delta h_{21}}{h_1}}\dfrac{u_\ast^3}{w_\ast^3}+\nonumber\\ &&A_{3{,KP}}\dfrac{-\frac{1}{2}\overline{uw}_1\Delta U-\frac{1}{2}\overline{vw}_1\Delta V}{\left(1+\frac{\Delta h_{21}}{h_1}\right)w_\ast^3} , (2)\end{eqnarray} where \(\overline{w\theta}_s\) is the surface heat flux; w* is the convective velocity scale in the shear-free CBL, defined as \(w_\ast^3=(g/\Theta_0)\overline{w\theta}_sh_1\); u* is the surface friction velocity; and A1 ,KP=0.20, A 2,KP=0.26, and A 3,KP=1.44. In order to conveniently compare with the parameterization scheme proposed in (Liu et al., 2016), here we do not express Eq. (3) in the same form as in KP06, in which \(-\overline{uw}_1\Delta U/2-\overline{vw}_1\Delta V/2\) is replaced by a derived relationship [(i.e., Kim et al., 2006, Eq. (5)].

    Following the work of (vanZanten et al., 1999), (Sun and Wang, 2008) derived a parameterization of the entrainment rate for a shear-free CBL, in which w e is normalized by w* and is proportional to the inverse of the convective Richardson number. They argued that their scheme is still valid for the sheared entrainment process because the scheme is derived from the profiles of potential temperature and its flux, which have the same shape in both sheared and shear-free CBLs. This argument is supported by the LES outputs for a sheared CBL (Sun and Xu, 2009). The alternative form of (Sun and Wang, 2008) given in (Sun and Xu, 2009) is expressed as \begin{equation} \label{eq3} w_{{e,SW}}=\dfrac{\Delta h_{21}+(h_1+\Delta h_{21})A_{{e,SW}}}{h_1\Delta\Theta}\overline{w\theta}_{s} , (3)\end{equation} and \begin{equation} \label{eq4} A_{{e,SW}}=\dfrac{-\overline{w\theta}_1}{\overline{w\theta}_{s}}=\dfrac{\Delta h_{10}}{h_0} , (4)\end{equation} where ∆ h10=h1-h0 is the thickness of the lower part of the entrainment zone. As pointed out in (Sun and Xu, 2009), KP06 and (Sun and Wang, 2008) are equivalent in describing sheared entrainment, although the expressions are different. KP06 explicitly includes the effects of shear-produced turbulence, whereas (Sun and Wang, 2008) uses a geometric relation to implicitly represent the wind shear effects.

    In both KP06 and (Sun and Wang, 2008), the effect of wind shear on the entrainment rate is represented by the A e. It should be noted that Eq. (3) is derived for GC CBLs. A parameterization of A e in a well-developed and sheared CBL is derived in (Liu et al., 2016). The expression reads: $$ A_e=A_1 \dfrac{1}{1+\frac{\Delta h_{21}}{h_1}}+A_2 \dfrac{C^{-1/2}_D}{1+\frac{\Delta h_{21}}{h_1}} \frac{u^3_*}{w^3_*}+$$ $$ Term I \ \ \ \ \ \ \ \ \ \ Term II$$ $$\ \ \ \ A_3\dfrac{(-\frac{1}{2} \overline{uw}_1 \Delta U-\frac{1}{2} \overline{vw}_1 \Delta V)}{(1+\frac{\Delta h_{21}}{h_1})w^3_*}+$$ $$\ \ \ \ \ \ \ \ \ \ Term III$$ $$\ \ \ \ \ \ \ \ \ \ A_4 \dfrac{(V_1-V_S)(-frac{1}{2} \overline{vw}_S-\frac{1}{2} \overline{vw}_1+\frac{1}{12}f\gamma_u h_1^2}{(1+\frac{\Delta h_{21}}{h_1})w^3_*}, (5)$$ $$\ \ \ \ \ \ \ \ \ \ Term IV$$ where V s is the velocity in the y-direction at 0.1h1 (the top of the surface layer), and C D=u*2/(U s2+V s2) is the surface drag coefficient. Note that the geostrophic velocity in the y-direction is zero in this study. The coefficients have been determined in (Liu et al., 2016), i.e., A1=0.21, A2=0.01, A3=0.86, and A4=0.70.

    The surface drag coefficient in Eq. (5) is not a constant. It is worth noting that an additional term that represents the effect of wind shear in the mixed layer is included in Eq. (5). In the GC case, Eq. (5) reduces to Eq. (3). On the other hand, Eq. (5) is the explicit form of Eq. (5) for a sheared CBL. The combination of Eq. (4) and Eq. (5) can be used to predict the evolution of the sheared CBL depth. However, with so many variables in the two equations, they are inconvenient for application. In this study, we attempt to develop a simplified scheme based on characteristics of sheared entrainment obtained from the LES outputs.

  • The relation between ∆Θ/∆ h20 and γθ is often used as a parameter to characterize the thermal structure in the entrainment zone (Deardorff, 1979; Fedorovich, 1995; Gentine et al., 2015). It is called the relative stratification parameter and is defined as \begin{equation} \label{eq6} G=\dfrac{\gamma_\theta}{\Delta\Theta/\Delta h_{20}} . (6)\end{equation}

    The LES results in (Fedorovich et al., 2004) indicate that the dimensionless parameter G is a constant of around 1.2 for shear-free entrainment. Whether it is a constant for sheared entrainment has not been discussed in previous studies. A relationship between G and A e is derived from the equation of temperature by order analysis using our LES results (details of the derivation given in Appendix A). It is expressed as \begin{equation} \label{eq7} G=\frac{1+K_1A_{e}}{1+A_{e}} , (7a)\end{equation} where the parameter K1 is defined as K1=2-∆ h21/2∆ h20. In order to further simplify the problem, we assume ∆ h21≈ ∆ h20/2 (details about G and the validation of the approximation are discussed in the next section). Thus, K1 can be treated as a constant of 7/4, and Eq. (8) can be written as \begin{equation} \label{eq8} G=\frac{1+\frac{7}{4}A_{e}}{1+A_{e}} . (7b)\end{equation} Equation (7b) indicates that G is a function of A e rather than a constant. Substituting Eqs. (6) and (7b) into Eq. (3) and applying the relation ∆ h20=∆ h21+∆ h10, we get \begin{equation} \label{eq9} w_{e}=\left(1+\dfrac{7}{4}A_{e}\right)\frac{\overline{w\theta}_{s}}{\gamma_\theta h_1} . (8)\end{equation} In (Liu et al., 2016), a new convective velocity scale (w m), which includes the contributions of shear-produced TKE in the whole CBL, is proposed. The results in (Liu et al., 2016) indicate that w m is suitable for characterizing the vertical turbulent motion at the mixed layer top. It is expressed as \begin{eqnarray} \label{eq10} w_{m}^3&=&w_\ast^3+\frac{A_2}{A_1}C_{D}^{-1/2}u_\ast^3+\frac{A_3}{A_1}\left(-\dfrac{1}{2}\overline{uw}_1\Delta U-\dfrac{1}{2} \overline{vw}_1\Delta V\right)+\nonumber\\ &&\dfrac{A_4}{A_1}(V_1-V_{s})\left(-\dfrac{1}{2}\overline{vw}_{s}-\dfrac{1}{2}\overline{vw}_1+\dfrac{1}{12}f\gamma_uh_1^2\right) . (9)\end{eqnarray} Then, Eq. (5) can be rewritten as \begin{equation} \label{eq11} A_{e}=\dfrac{A_1}{1+\Delta h_{21}/h_1}\dfrac{w_{m}^3}{w_\ast^3} . (10)\end{equation} The variation of ∆ h21/h1 is relatively large during the development of a CBL. However, the variation of 1+∆ h21/h1 is so small (Table 1) that it can be approximately treated as a constant. The LES outputs show that the average value of 1+∆ h21/h1 is 1.19. Since the value of A1 is set to 0.21 (Liu et al., 2016), Eq. (10) is further simplified to be \begin{equation} \label{eq12} A_{e}=0.18\frac{w_{m}^3}{w_\ast^3} . (11)\end{equation} It is worth noting that Eq. (11) is not the scheme in the zeroth-order jump model (ZOM), although the expression has the same form as that in the ZOM. As shown in Eq. (9), the parameterization is still complex and includes many unknown variables. Thus, how to simplify Eq. (9) is the problem we need to solve.

    (Liu et al., 2016) shows that the effects of wind shear in the surface layer and mixed layer [the second and fourth terms on the right-hand side of Eq. (5)] on the entrainment flux are quite small. Since the terms on the right-hand side of Eq. (9) are proportional to the terms on the right-hand side of Eq. (5), it can be concluded that the second and fourth terms on the right-hand side of Eq. (9) are small when compared with the first and third terms. The ratio of the sum of the second term and the fourth term to w m3 is smaller than 5% in most simulated cases (Table 1). For the purpose of simplicity, the two small terms are neglected in the simplified parameterization. Therefore, the simplification of Eq. (9) begins with the third term, hereafter expressed as \(X_3=-\overline{uw}_1\Delta U/2-\overline{vw}_1\Delta V/2\). In order to derive ∆ U, ∆ V, \(\overline{uw}_1\) and \(\overline{vw}_1\), the velocity equation is vertically integrated from the surface to the level just above the entrainment zone top. Rearrangement of the integrations yields (see details of derivation in Appendix B) \begin{eqnarray} \label{eq13} h_1\dfrac{\partial U_1}{\partial t}&=&\overline{uw}_{s}-\overline{uw}_1\!+\!\frac{1}{2}fh_1[(V_{s}+V_1)-(V_{{g,s}}\!+\!V_{g,1})] ,(12)\nonumber\\\\ \left(1+\dfrac{\Delta h_{21}}{2h_1}\right)\overline{uw}_1&=&\dfrac{\Delta h_{21}}{2h_1}\overline{uw}_{s}- \left(\Delta U-\dfrac{1}{2}\gamma_u\Delta h_{21}\right)\dfrac{\partial h_1}{\partial t}+\nonumber\\ &&\dfrac{1}{4}f\Delta h_{21}[(V_{s}-V_1)-(V_{{g,s}}-V_{{g},1})] ,(13)\\ \dfrac{\partial \Delta U}{\partial t}&=&\gamma_u\dfrac{\partial h_1}{\partial t}-\dfrac{\partial U_1}{\partial t} . (14)\end{eqnarray} By using the differential transition, such as \(\partial/\partial t=\partial/\partial h_1\cdot\partial h_1/\partial t=(S/h_1)\partial/\partial h_1\), where \(S=\overline{w\theta}_s[1+(7/4)A_e]/\gamma_\theta\) comes from Eq. (8), Eq. (14) becomes \begin{equation} \label{eq14} \frac{S}{h_1}\frac{\partial \Delta U}{\partial h_1}=\gamma_u\frac{\partial h_1}{\partial t}-\frac{\partial U_1}{\partial t} . (15)\end{equation} Substituting Eqs. (12) and (13) into Eq. (15), we rearrange the equation to be: \begin{eqnarray} \frac{\partial \Delta U}{\partial h_1}&=&-\frac{1}{2h_1+\Delta h_{21}}(2\Delta U-2\gamma_u\Delta h_{21}-2\gamma_uh_1)-\nonumber\\ &&\frac{2h_1}{2h_1+\Delta h_{21}}\frac{\overline{uw}_{s}}{S}-\frac{1}{S}\frac{h_1}{2h_1+\Delta h_{21}}f\Delta h_{21}(V_1-V_{{g},1})-\nonumber\\ &&\frac{1}{S}\frac{h_1}{2h_1+\Delta h_{21}}fh_1(V_{s}-V_{{g,s}}+V_1-V_{{g,1}}) . \label{eq15} (16)\end{eqnarray} The solution of Eq. (16) can be obtained, in which the coefficients are substituted in for the purpose of simplicity (see details in Appendix C). It reads: \begin{eqnarray} \label{eq16} \Delta U&=&0.57\gamma_uh_1-0.48\frac{h_1}{S}\overline{uw}_{s}-\nonumber\\ &&\frac{fh_1^2(0.19V_1-0.19V_{{g},1}+0.16V_{s}-0.16V_{{g,s}})}{S} .\qquad (17)\end{eqnarray} To solve Eq. (17), we assume that ∆ U depends only on h1 and 1+∆ h21/h1=1.19. Substituting Eq. (17) into Eq. (13) gives a simple expression of \(\overline{uw}_1\). It is written as \begin{eqnarray} \label{eq17} \overline{uw}_1&=&0.57\overline{uw}_{s}-0.47\gamma_uS+\nonumber\\ &&fh_1(0.14V_1-0.14V_{{g},1}+0.20V_{s}-0.20V_{{g,s}}) ,\quad (18)\end{eqnarray} where V g,s and V g,1 are the geostrophic velocity at the surface and CBL top, respectively.

    The expressions of ∆ V and \(\overline{vw}_1\) are derived in the same way as ∆ U and \(\overline{uw}_1\). For the CBL under the GC condition, V is constant while \(\overline{vw}\) varies linearly with height in the mixed layer. Their profiles present the same characteristics as that of U and \(\overline{uw}\). Thus, the expressions of ∆ V and \(\overline{vw}_1\) have the same form as ∆ U and \(\overline{uw}_1\). According to Eq. (17) and Eq. (18), they can be written as (the terms associated with the vertical gradient of the geostrophic velocity in the y-direction γv diminish because γv=0) \begin{equation} \label{eq18} \Delta V\!\!=\!-0.48\frac{h_1}{S}\overline{vw}_{s}\!+\!\frac{fh_1^2(0.19U_1\!\!-\!0.19U_{{g},1}\!\!+\!0.16U_{s}\!-\!0.16U_{{g,s}})}{S} (19)\end{equation} and \begin{equation} \label{eq19} \overline{vw}_1=0.57\overline{vw}_{s}-fh_1(0.14U_1-0.14U_{{g},1}+0.20U_{s}-0.20U_{{g,s}}). (20)\end{equation} Therefore, in the GC case, the third term in Eq. (9) can be written as (the terms associated with γu diminish because γu=0 in the GC case) \begin{equation} \label{eq20} X_{3,{GC}}=-\frac{1}{2}(\overline{uw}_1\Delta U+\overline{vw}_1\Delta V)=0.14\frac{h_1}{S}u_\ast^4+F_{f,{GC}} , (21a)\end{equation} where Ff, GC represents the sum of the terms, including the Coriolis parameter (expression not shown). In order to estimate X3, GC, Eq. (9) with the exact coefficients is written here (the second and fourth terms have been omitted): \begin{equation} \label{eq21} w_{m}^3=w_\ast^3+4.10X_3 . (9')\end{equation} The LES results show that w m3=w*3+5u*3, the formula proposed by (Moeng and Sullivan, 1994), is a good approximation for describing w m under the GC condition. Comparing Eq. (9') with Eq. (21a), it is found that 0.14(h1/S)u*4+Ff, GC≈ u*3, which can lead to w m3≈ w*3 +5u*3.

    However, for a CBL under the GS condition, V is not constant in the mixed layer, while the shape of the \(\overline{vw}\) profile is different from that in the GC case. The expressions of ∆ V and \(\overline{vw}_1\) should be different to those in the GC case. The LES results show that, in the GS case, \(\overline{vw}_1\Delta V\) is far smaller than \(\overline{uw}_1\Delta U\), and thus can be neglected (Table 1). That is, \(\overline{uw}_1\Delta U+\overline{vw}_1\Delta V\approx\overline{uw}_1\Delta U\). The LES results also show that, in the GS case, \(\overline{vw}_s\approx 0\) and \(\overline{uw}_s\approx-u_\ast^2\). Thereby, X3, GS can be obtained from Eq. (17) and Eq. (18). It is written as \begin{eqnarray} \label{eq22} X_{3,{GS}}&=&-\frac{1}{2}\overline{uw}_1\Delta U\nonumber\\ &=&0.28u_\ast^2\gamma_u h_1+0.14\gamma_u^2Sh_1+0.14\frac{h_1}{S}u_\ast^4+F_{f,{GS}} ,\qquad\ (21b)\end{eqnarray} where Ff, GS represents the sum of the terms, including the Coriolis parameter (expression not shown). When the geostrophic velocity gradient becomes zero (i.e., γu=0), X3, GS should reduce to X3, GC. Based on this consideration, we assume that the approximation 0.14(h1/S)u*4+Ff, GS≈ u*3 is also appropriate for the GS case. Thus, the expression of X3 for both the GC and GS cases is \begin{equation} \label{eq23} X_3=\eta u_\ast^3+0.28\gamma_u u_\ast^2h_1+0.14\gamma_u^2Sh_1 , (22)\end{equation} where η is an empirical constant introduced by the above approximations. Replacing S in Eq. (22) by the relation \(S=\overline w\theta_s[1+(7/4)A_e]/\gamma_\theta\), and further replacing A e by using Eq. (11), then substituting Eq. (22) into Eq. (9'), we get the following equation: \begin{equation} \label{eq24} \left(1-0.37a_3\frac{\Theta_0\gamma_u^2}{g\gamma_\theta}\right)w_m^3=\left(1+a_3\frac{\Theta_0\gamma_u^2}{g\gamma_\theta}\right)w_\ast^3+a_1 u_\ast^3+a_2u_\ast^2\gamma_u h_1 , (23)\end{equation} where a1=4.10η, a2=1.15 and a3=0.57. Since some approximations and assumptions are introduced in the derivation, these coefficients should be adjusted. Multiple linear regression of the LES outputs in the GC and GS cases gives a1 =6.02, a2 =0.24 and a3=0.86.

    (Pino and De Arellano, 2008) used local momentum fluxes and velocity jumps to represent the shear production rate of TKE. Equations (17) and (18) indicate that these local quantities are related to surface fluxes, entrainment rate, CBL height and geostrophic velocity gradient. Equation (22) also indicates that the net shear production rate of TKE in the IL comprises the dynamic effect (second term), the interaction between mean shear and environmental stratification (third term), and one other term that includes the Coriolis effects (first term).

    Equations (8), (11) and (23) constitute a simple model to predict the height of a sheared CBL. Compared with the KP06 and Sun and Wang (2008) methods, this model does not need those variables at the CBL top, such as ∆Θ , ∆ h21, and so on. Owing to the coarse resolution of NWP and air pollution models, these variables cannot be resolved explicitly (Hong et al., 2006). Thereby, they must be obtained from a bulk model. However, the input parameters in our simplified model can be easily derived in NWP and air pollution models. \(\overline{w\theta}_s\) and u* can be estimated by a land surface model. γθ can be derived as the mean gradient of potential temperature within a certain thickness (for example, 1000 m) above h2. γu can be treated as the mean gradient of velocity above h2 and determined by the same method for the calculation of γθ. In the real atmosphere, the treatment of γu may introduce some errors. However, the CBL's growth is mainly controlled by surface heating, which explains 70%-90% of this process (Canut et al., 2010; Sühring et al., 2014). This implies that the contribution of entrainment to CBL growth is about 10%-30%. It is expected that the errors induced by γu cannot significantly influence the prediction accuracy. Therefore, this model is more convenient to apply in numerical models.

3. Evaluation and discussion
  • The outputs of the 30 LES runs, described in (Liu et al., 2016), are used to verify the parameterizations under the conditions of varying geostrophic wind or wind shear, external stratification, and surface roughness length. The main features of these simulations are described in detail in (Liu et al., 2016). There are two simulations for the shear-free CBL (NS00S3 and NS00S6), while all other simulations are for the sheared CBL and divided into three groups. One group is for CBLs under the GC condition, with vertically uniform geostrophic velocities of 10 m s-1, 15 m s-1 and 20 m s-1, respectively. The second group is for CBLs under the GS condition, with geostrophic wind gradients of 10 m s-1 (2 km)-1, 15 m s-1 (2 km)-1 and 20 m s-1 (2 km)-1, respectively, and zero surface geostrophic velocity. In each group under the GC and GS conditions, the simulations are conducted with external temperature gradients of 3 K km-1 and 6 K km-1, and surface roughness length values of 0.01 m and 0.1 m, respectively. The name of a simulation case is given according to the simulation conditions. For example, GC20R6 means that the simulation is conducted under conditions of U g=20 m s-1, a rough surface with z0=0.1 m, and ∂Θ/∂ z=6 K km-1. In section 2, the derivation of Eq. (23) is based on assumptions and approximations that are obtained from results of the GC and GS cases, and the coefficients are fitted from simulations of these 24 cases. In order to confirm their validity, a third group of simulations are conducted under the CS condition (C5S10S3, C5S15S3, C5S15S6 and C5S15R3). The CS condition can be regarded as a combination of the GC and GS conditions, in which the geostrophic wind shear exists while the surface geostrophic velocity is not zero. In the four CS cases, the surface geostrophic velocity is 5 m s-1 (denoted as C5), while its vertical gradient is 10 m s-1 (denoted as S10) or 15 m s-1 (denoted as S15) per 2 km, and the meanings of the last two letters in the case names are the same as for the GC and GS cases. The integration covers 28 800 s, and the results from 4800 s to 28 800 s are output at an interval of 100 s for further calculations and analyses. The methods to determine the variables used in calculations and analyses are introduced in (Liu et al., 2016).

  • According to the definition expressed in Eq. (6), the LES outputs are used to calculate the relative stratification parameter G, and the results are shown in Fig. 2 (blue dots). In the GC cases, G has a slightly decreasing trend during CBL development. However, the decrease is negligibly small (only about 0.1 in a long period of 24 000 s). In the GS cases, G almost does not vary with time. As shown in Table 1, the average value of G varies slightly in different cases. It increases with increasing geostrophic velocity in the GC cases and increasing geostrophic velocity gradient in the GS cases. However, the difference in G among the simulated cases is very small. The mean value of G in all of the simulated cases is 1.17 (Table 1), which is very close to the result for shear-free CBLs in Fedorovich et al. (2004, therein G≈ 1.2).

    Figure 2.  Comparison of $G$ (the relative stratification parameter) calculated from the definition and from the parameterization scheme in the GC and GS cases. The blue dots represent the results from Eq. (6) and the red dots represent the results from Eq. (7b).

    The parameterization of the relative stratification parameter, i.e., Eq. (7a), is derived in the FOM framework. As shown in Appendix A, an approximation is applied to simplify Eq. (A6). The LES outputs are used to calculate every term on the left-hand side of Eq. (A6), and results indicate that the second and third terms are one order smaller than the other two terms. The LES outputs provide a solid basis for Eq. (7a). However, Eq. (7a) still includes variables at the CBL top (∆ h21 and ∆ h20). A further simplified parameterization of G, Eq. (7b), is obtained by using the assumption of ∆ h21≈∆ h20/2, which may not be true, but the LES outputs show that K1 is almost a constant and approximately equals 7/4 (Table 1). In order to evaluate the performance of Eq. (9), the LES outputs are used to calculate G according to Eq. (7b). The results are also plotted in Fig. 2 (red dots) to compare to those calculated according to Eq. (6) [Fig. S1 in electronic supplementary material (ESM) for CS cases]. The spread of LES outputs (especially for GC20 cases when the integration time is longer than 20 000 s) is mainly attributed to the determination method of the upper edge of the entrainment zone, h2, where the instantaneous potential temperature flux profile is not equal to zero but fluctuates around zero. Significant fluctuations of ∆ h21 can lead to large fluctuations of ∆Θ. However, the fluctuation of ∆ h20=∆ h21 +∆ h10 is relatively small. A small value of ∆Θ always corresponds to a large value of G calculated from Eq. (6). Figure 2 shows that the results from Eq. (6) and Eq. (7b) agree very well in all of the simulation cases, indicating that Eq. (7b) can accurately describe the behavior of G.

  • The parameterization of entrainment rate, Eq. (8), is derived from Eq. (3) and Eq. (7b). (Sun and Xu, 2009) demonstrated that Eq. (3) works in sheared CBLs. Equation (7b) is validated in the previous section. Thus, we expect that Eq. (8) can predict the CBL height correctly. In order to prove this point, Eq. (10) is solved numerically using the Euler predictor-corrector method over the period from 5700 s to 28 800 s under the CS condition. The time step is 100 s, the same as the time interval of LES outputs. The parameterization schemes of A e and the convective velocity scale, Eqs. (11) and (23), are used to close Eq. (8). The LES outputs for the period from 4800 s to 6600 s are averaged to provide the initial condition at 5700 s. Results are illustrated in Figs. 3-5.

    Figure 3.  Convective velocity scales calculated from the CBL height prediction model [red line, Eq. (23)] and from LES outputs [blue dots, Eq. (9)] in CS cases.

    Figure 4.  $A_e$ values calculated from the CBL height prediction model [red line, Eq. (11)] and from LES outputs [blue dots, Eq. (4)] in CS cases.

    Figure 5.  Values of CBL height predicted by the model (red line) and obtained from LES outputs (blue dots) in CS cases. Green squares show the nonlinear fit of the result of LES CBL height based on the relation that the CBL height is proportional to the square root of time.

    Figure 3 shows that Eq. (23) slightly underestimates and overestimates w m at the beginning and end, respectively. This difference is mainly due to the approximation of 0.14(h1/S)u*4+Ff ≈ u*3, and a larger u* gives a larger error. However, the biases of estimated w m are very small, making it reasonable to conclude that the simplified form of the convective velocity scale Eq. (23) agrees well with its original form, Eq. (9). Figure 4 shows that the A e estimated by Eq. (11) is in good agreement with that derived from the LESs. As presented in previous studies (e.g. Conzemius and Fedorovich, 2006; Pino et al., 2006), the LES A e spreads widely because it is determined from 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 values of the parameterized A e are contained within the fluctuations of the LES outputs. Figure 5 indicates that Eq. (8) can correctly predict the CBL height.

    (Fedorovich et al., 2004) suggested that, in a shear-free CBL, the CBL height is proportional to the square root of time, i.e., \(h_1\propto\sqrt t\). The direct numerical simulations of the CBL driven by a constant momentum flux in (Jonker et al., 2013) yielded the same result, i.e., \(h_1 \propto \sqrt t\). Our LES results show that this relation also exists in the GC, GS and CS cases (Fig. 5 only displays the results in the CS case; see Fig. S2 in ESM for the GC and GS cases).

  • The entrainment rate is often parameterized as w e/w*=ARi*-1, where Ri*=(g/Θ0)h1∆Θ/w*2 is the bulk convective Richardson number. In the ZOM, the coefficient A is just the A e. In the FOM, this expression still applies, but the convective Richardson number is different to that in the ZOM because of the different definition of ∆Θ. Determination of the coefficient A has been explored (e.g., Lewellen and Lewellen, 1998; Sullivan et al., 1998; vanZanten et al., 1999; Sun and Wang, 2008). (Sun and Wang, 2008) gave a relation of A=(1+A e)∆ h20/h1, indicating that A is associated with not only the A e but also the entrainment zone thickness and CBL depth. The parameterization scheme for the entrainment rate in (Sun and Wang, 2008) can be written as Eq. (3). KP06 provided an equivalent scheme, expressed as Eq. (1). Both Eq. (1) and Eq. (3) are valid for a sheared CBL, and both imply that the effect of wind shear can be represented by the A e. On the other hand, the two schemes both include the variable ∆Θ, which is usually unknown. This problem is solved by introducing G. A simple relationship between G and A e, expressed as Eq. (7b), is derived in this study; and the parameterization of the entrainment rate turns out to be Eq. (8).

    For a sheared CBL, the influence of wind shear is included in the A e. Equation (11) further indicates that the effects of wind shear can be represented by the convective turbulent velocity scale in a sheared CBL, which is always larger than that in a shear-free CBL. Equation (23) gives a parameterization of the convective velocity scale, in which the CBL bulk variables, such as ∆ U, ∆ V, ∆ Θ and ∆ h21, which cannot be resolved explicitly in most numerical models, are not needed. This is the fundamental difference to the parameterization of KP06.

    Actually, Eq. (3) is not the final form of the parameterization of A e in KP06. The final form is expressed as

    \begin{equation} \label{eq25} A_{{e,KP}}=\dfrac{A_{1,{KP}}\frac{h_1}{h_2}+A_{2,{KP}}\frac{u_\ast^3}{(w'_\ast)^3} +A_{3,{KP}}\frac{\Delta h_{21}}{2h_1+2h_2}\left[\frac{u_\ast^2 \Delta V_{e}}{(w'_\ast)^3}+ \frac{\Theta_{m}(\Delta V_{e})^2}{gh_2(\Delta\Theta-0.5\gamma_\theta\Delta h_{21})}\right]} {\left[1-A_{3,{KP}}\frac{\Theta_{m}(\Delta V_{e})^2}{2gh_2(\Delta\Theta-0.5\gamma_\theta\Delta h_{21})}\right]} , (24)\end{equation}

    where Θ m is the mean potential temperature in the mixed layer, \((w'_\ast)^3=g\overline{w\theta_s}h_2/\Theta_m\), and ∆ V e is the velocity jump across the IL. (Liu et al., 2016) shows that the value of 1.44 for A3, KP overestimates the contribution of shear-produced TKE in the IL to entrainment. Our LES outputs are used to optimize the parameters in Eq. (26). In (Kim et al., 2006), ∆ V e=0.5(|∆ U|+|∆ V|), and the linear regression yields A1, KP=0.22, A2, KP=-0.54 and A3, KP=2.39. In (Pino et al., 2006), $(\Delta V_e=\sqrt{(\Delta U)^2+(\Delta V)^2})$, and the linear regression yields A1, KP=0.22, A2, KP=0.23 and A3, KP=0.74. Because A2, KP=2C D-1/2(1-α2), A3, KP=2(1-α3), where α2 and α3 represent the proportions of the dissipation rate to the production rate, and C D is the surface drag coefficient; A2, KP and A3, KP must be positive and A3, KP must be less than 2. The negative value of A2, KP and the larger than 2 value of A3, KP imply that the treatment of ∆ V e in (Kim et al., 2006) is not reasonable. In the following calculation, only the ∆ V e in (Pino et al., 2006) is adopted. In order to identify the differences among (Sun and Wang, 2008), LS (LS is the scheme in this study) and the optimized KP06 schemes, the LES outputs are used to calculate w e in the simulated cases, and the relative errors are compared in Fig. 6. The relative error is defined as \begin{equation} \label{eq26} {Err}=\frac{1}{n}\sum\left|\dfrac{w_{{e,p}}}{w_{{e,LES}}}-1\right| , (25)\end{equation} where w e,p is the predicted entrainment rate, \(w_e,LES=\partial\langle h_1\rangle/\partial t\), and \(\langle h_1\rangle\) is the least squares nonlinear fit of LES h1 based upon the relation \(\langle h_1\rangle\propto \sqrt t\) (Fig. 5). In order to understand whether the error of KP06 comes from the parameterization scheme of the entrainment rate [namely, Eq. (2)] or the parameterization scheme of the A e [i.e., Eq. (26)], the errors of the entrainment rate predicted by Eq. (2) and the simplified A e parameterization scheme [i.e., Eq. (13)] (denoted as KP LS) are also calculated and shown in Fig. 6. Among all the three schemes, the LS scheme performs best, implying that the approximations applied in the derivation of Eq. (25) are reasonable and the derived parameterization can successfully capture the characteristics of the entrainment rate in a sheared CBL. Note that the LS scheme performs better than the (Sun and Wang, 2008) scheme, although the former is developed on the basis of the latter by using some approximations. The reason is because the spread of instantaneous LES variables (∆Θ, ∆ h21, and so on) used in the (Sun and Wang, 2008) scheme is larger than that of the instantaneous LES variable (namely, h1) used in the LS scheme. KP LS performs better than the KP06 scheme with optimized parameters but slightly worse than the (Sun and Wang, 2008) scheme. This suggests that Eq. (2) performs slightly worse than Eq. (4) and the A e estimated by Eq. (26) has large errors in some cases, even though the parameters have been optimized. The errors of Eq. (26) are partially due to the wide spread of instantaneous LES variables used in Eq. (26). Based on the above discussion, it is concluded that the assumptions used in the derivations are reasonable and the simplified parameterizations proposed in this study can correctly predict the A e and entrainment rate. Meanwhile, the simplified parameterizations proposed in this study do not include entrainment variables, which may introduce large calculation uncertainties.

    Figure 6.  Relative errors of the entrainment rate predicted by different schemes against the LES outputs in the simulated cases.

4. Concluding remarks
  • This study aims to simplify the parameterization of the entrainment rate so that it can be conveniently applied to predict the growth rate of a well-developed and sheared CBL. To achieve this goal, G (the relative stratification parameter) is introduced into the parameterization scheme, and the A e and convective velocity scale in a sheared CBL are simplified according to the characteristics of sheared entrainment, which are derived from the LES outputs. The major findings can be summarized as follows:

    (1) G, which is defined in Eq. (6), can be used to characterize the thermal structure in the entrainment zone. (Fedorovich et al., 2004) suggested that it is a constant of around 1.2. Theoretical analysis indicates it is a function of the A e as expressed in Eq. (7b), rather than a constant. This result is supported by the LES outputs.

    (2) When the relationship between G and the A e is introduced, the parameterization scheme of the entrainment rate proposed by (Sun and Wang, 2008) can be rewritten as a simple function of A e, surface heat flux, and the potential temperature gradient in the free atmosphere and the CBL height, as expressed in Eq. (8).

    (3) Shear-produced TKE at the CBL top enhances entrainment. (Pino and De Arellano, 2008) used the local momentum fluxes and velocity jumps to parameterize the shear production rate of TKE. Our results show that these local quantities are related to surface momentum fluxes, the entrainment rate, the geostrophic velocity gradient, the stratification in the free atmosphere, and the CBL height. The net shear production rate of TKE in the IL is expressed by Eq. (24), which comprises the dynamic effect, the interaction between mean shear and environmental stratification, and one other term that includes the Coriolis effect, and can be approximately characterized by u*3. Based on Eq. (22), the convective velocity scale proposed in (Liu et al., 2016) can be further expressed by Eq. (23).

    (4) In the framework of FOM, the IL thickness affects the A e. The LES outputs show that this effect can be described by a constant, and thus the parameterization of A e proposed in (Liu et al., 2016) is simplified by Eq. (11).

    (5) The parameterizations of the entrainment rate, A e and convective velocity scale constitute a prediction model for CBL height. The LES outputs show that it is an appropriate model for a well-developed and sheared CBL. Compared with bulk models, the parameters needed by this model can be easily derived in an NWP or air pollution model. Therefore, this model is more convenient for application. However, the performance in NWP and air pollution simulations needs further validation.

    As pointed out in (Liu et al., 2016), the parameterizations obtained in this study may only be suitable for well-developed CBLs under idealized conditions. The storage term in the TKE budget is ignored in the derivations. However, this term is not negligibly small in the early stage of CBL development. Under this condition, the performance of the simple model needs to be verified. In this study, the potential temperature gradient in the free atmosphere and the geostrophic velocity gradient are assumed to be constant. For a CBL that is growing through a pre-existing inversion or a residual layer, application of Eq. (8) becomes problematic, especially when there is a residual layer, because the relative stratification parameter may not exist. How to parameterize the entrainment rate under these conditions needs further investigation. In the real atmosphere, the geostrophic wind may not vary linearly with height. Thereby, the applicability of the simplified parameterizations under this condition needs more evaluation. As discussed in section 2.3, the simple model developed in this study is convenient to apply in numerical models. We will incorporate this scheme into numerical models and evaluate its performance under real conditions in future work.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return