Next Article in Journal
A Finite Element Method for Compressible and Turbulent Multiphase Flow Instabilities with Heat Transfer
Next Article in Special Issue
Recent Advances in Experimental and Numerical Studies on Cloud and Erosion Behaviors in Cavitating Jets
Previous Article in Journal
Pulsation-Enhanced Transport in Pseudo-Periodic Porous Channels
Previous Article in Special Issue
Modelling and Simulation of the Entrapment of Non-Wetting Fluids Within Disordered Porous Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Atmospheric Gray-Zone (a.k.a. Terra Incognita) Problem: A Strategy Analysis from an Engineering Viewpoint

Department of Mathematics and Statistics, University of Wyoming, 1000 E. University Avenue, Laramie, WY 82071, USA
Fluids 2025, 10(11), 301; https://doi.org/10.3390/fluids10110301
Submission received: 1 October 2025 / Revised: 7 November 2025 / Accepted: 10 November 2025 / Published: 18 November 2025
(This article belongs to the Special Issue Feature Reviews for Fluids 2025–2026)

Abstract

The Terra Incognita (or gray-zone) problem seen in atmospheric flow simulations causes serious consequences: it implies, e.g., significantly incorrect flow predictions and results that often simply depend on flow simulation settings as the computational grid applied. There is definitely the need for a robust gray-zone modeling to ensure that research and technology decisions are based on reliable results. As a matter of fact, solution approaches to deal with this problem in atmospheric and engineering type simulations reveal remarkable differences. In contrast to atmospheric flow simulations, there exists a broad spectrum of solution concepts for engineering applications. Driven by these conceptual differences, the paper presents an analysis of the Terra Incognita problem and corresponding solution concepts. Specifically, the paper presents a modeling approach that overcomes the core problem of currently applied methods. A new method of providing a resolution-aware turbulence length scale (one of the major problems in atmospheric flow simulations) is presented. This approach is capable of seamlessly covering the full range of microscale to mesoscale simulations, and to appropriately deal with mesoscale to microscale couplings.

1. Introduction

Accurate microscale (a resolution of meters to a few km) and mesoscale (a resolution of a few km to hundreds of km) simulations of atmospheric flows are essential for understanding, predicting, and managing weather, climate, and environmental impacts at local and regional scales. This concerns, e.g., high-resolution weather prediction, severe weather forecasting, urban and environmental applications (air pollution dispersion at street level, wind safety for buildings and infrastructure), wildfire and smoke behavior, climate and renewable energy applications, and disaster risk reduction (chemical/biological releases).
In particular, the coupling of mesoscale and microscale simulation methods is a major challenge of wind energy research [1,2,3,4,5,6]. Microscale simulation methods are capable of providing detailed information about terrain effects and interactions with structures such as wind turbines. The atmosphere drives changes at the wind plant scale; it is therefore essential to model the variability at the mesoscale. Consequently, the appropriate coupling of mesoscale and microscale simulation methods for atmospheric boundary layer (ABL) processes is highly relevant to the evaluation of topography effects and the evaluation of the suitability of wind turbine designs and wind farm effectiveness [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27]. Results of an extended study of strategies to deal with this problem (performed in the frame of the U.S. Department of Energy Mesoscale-to-Microscale-Coupling project) were presented recently [6,16,17,18,19,20,21,28,29]. There is definitely progress regarding the understanding of the nature of this problem and potential ways to address this issue. But simultaneously, there is so far no indication that a convincing solution to this problem can be expected in the near future. Doubrawa and Muñoz-Esparza [22] recently concluded “the need for a robust gray-zone modeling solution is urgent to ensure that research and technology decisions are based on reliable results”.
The technical nature of these issues is the following (an overview of terminology applied in this paper can be found in Table 1). Microscale and mesoscale core techniques are well developed. Microscale simulations are usually performed by using the large eddy simulation (LES) technique in conjunction with surface parameterizations of heat and momentum fluxes on relatively fine grids that resolve energetic, turbulent eddies aside from the close proximity to the surface. Such simulations are characterized by a relatively small turbulent viscosity (a relatively small turbulence length scale). Mesoscale simulations are performed on the basis of Reynolds-averaged Navier–Stokes (RANS) equation methods in conjunction with relatively coarse grids. Turbulent fluctuations are damped out, which is managed through a relatively large turbulent viscosity (a relatively large turbulence length scale). The need to couple mesoscale and microscale simulation methods (and the use of mesoscale models at a higher resolution to improve simulation results) requires the use of LES (RANS) methods at a rather coarse (fine) resolution, which disagrees with the underlying physical principles of these techniques. For example, there is no guarantee that resolved fluctuations produced by coarse-grid LES relate to real turbulence. There is also no guarantee that mechanisms that govern RANS equations are properly functional in the presence of chaotic impacts (fluctuations). Thus, there exists a gray-zone under such conditions, a region of (at the minimum) an unclear standard of simulations. In regard to atmospheric flow simulations, the gray-zone existence was identified by Wyngaard (2004) and called the Terra Incognita problem [30].
In regard to the gray-zone problem, it is of interest that there exist significant differences between atmospheric and engineering solution approaches. In contrast to such developments in regard to atmospheric flows, a huge variety of methods have been developed in respect to engineering applications. One very popular way is wall-modeled LES (WMLES) [31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46], where RANS components are involved close to solid walls. A variation in this approach is given by the development of Reynolds-stress-constrained LES (RSC-LES) [47,48,49,50,51,52,53,54,55,56,57,58]. Another very popular way is given by detached eddy simulation (DES) [59,60,61,62,63,64,65,66,67,68,69,70], where the performance of RANS models is improved by switching from the RANS turbulence length scale applied close to the wall to a much smaller LES-type length scale away from the wall. A variety of alternative methods were suggested [71,72,73,74], including unified RANS-LES (UNI-LES) [75,76,77,78,79,80,81], partially averaged Navier–Stokes (PANS) [82,83,84,85,86,87,88,89,90,91,92,93], partially integrated transport modeling (PITM) [94,95,96,97,98,99,100,101,102,103,104], and scale adaptive simulation (SAS) methods [70,105,106,107,108,109,110]. These developments were driven by needs to appropriately deal with separated turbulent flow [59,67,68] and many other flow simulations [111].
The motivation of this paper is to provide an analysis of the atmospheric gray-zone problem from an engineering viewpoint by simultaneously developing a solution concept for this problem. Section 2 deals with a closer look at the atmospheric gray-zone problem, its nature, its relevance, and currently applied solution concepts. The specific goal is twofold. First, the predominant trend of dealing with the gray-zone problem in regard to atmospheric flow simulations (to follow engineering DES concepts) is identified. Second, evidence obtained by engineering simulations is used to argue that this current trend implies a variety of issues. Section 3 presents a solution strategy to overcome the gray-zone problem in atmospheric flow simulations based on exact mathematical developments. Following CES concepts, a hybrid RANS-LES model is presented that is demonstrated to overcome the typical DES problems. This discussion includes computational evidence for conceptual CES advantages. Section 4 presents conclusions of this analysis.

2. The Atmospheric Gray-Zone Problem

Let us first have a closer look at the concrete problem considered, current solution concepts, and corresponding concepts used in engineering applications.

2.1. The Problem Considered

The atmospheric gray-zone problem is illustrated in Figure 1: Figure 1a shows a mesoscale RANS grid, Figure 1c shows an LES grid, and the grid in Figure 1b is in between the RANS and LES grids. Specifically, mesoscale RANS simulations do not attempt to resolve turbulence structures, whereas microscale LES simulations are focused on a resolution of turbulence structures. It is worth noting that mesoscale and microscale simulations are usually performed using very different equation structures (involving, e.g., one-dimensional (1D) and 3D turbulence schemes), and, first of all, by using very different turbulence length scales (a large-scale turbulence length scale L is used in mesoscale simulations, whereas an LES filter width Δ is applied in LES).
Significant questions arise in regard to simulations using intermediate grids as shown in Figure 1b. These are conditions where the grid spacings ( Δ ) are comparable to the size of energy-containing turbulent eddies L. What can be wrong with such simulations? Such simulations are significantly affected by fluctuations (produced resolved motion), but these fluctuations cannot represent real turbulence because there is insufficient flow resolution. At the same time, there is no guarantee that mechanisms that govern RANS equations are properly functional in the presence of random disturbances (resolved fluctuations).
Explicit shortcomings of gray-zone simulations are described in Table 2. Some of the key issues are the following ones (the horizontal blocks in Table 2 are ordered according to the four following issues):
1.
Inappropriate parametrizations: (i) Turbulence schemes reveal drawbacks: the resolved part is too large (or too weak) if the turbulence scheme is used without (or with) the mass-flux scheme [112,113], (ii) schemes need to account for 3D turbulence in the gray-zone [113,114,115], (iii) there can be excessive mixing due to a large turbulence length scale at coarse resolutions, which results in the cut-off of resolved turbulence in the gray-zone [116].
2.
Spin-up need and its sensitivity: (i) Without perturbations (initiation of resolved turbulence called spin-up), it is possible that resolved turbulence does not become established at all [117]. (ii) Models require an enhancement of diffusion to give a collapse of turbulence at Δ x h (h is the boundary-layer depth), whereas spin up of turbulence within the gray-zone requires a reduced diffusion [118].
3.
Grid and scheme dependence of results: When the grid spacing is similar to the dominant physical scales, then the grid influences what motions get allowed or forced, leading to artificial structures, and spurious triggering of convection or turbulence. For example, (i) mesoscale structures (rolls vs. cells, turbulence flux profiles, etc.) depend strongly on the horizontal grid resolution (and anisotropy between horizontal and vertical motions) [22,119], (ii) with respect to a tropical cyclone structure and intensity (eyewall slope, radius of maximum wind, latent heating, microphysics) simulation results vary significantly with the resolution; the results depend on convective parametrization schemes [120].
4.
Incorrect simulation predictions: Errors in regard to precipitation extremes, boundary layer depth, diurnal heating, and cloud fraction often increase in gray-zones. For example, (i) we see overestimations and erroneous precipitation locations or an erroneous simulation of heavy precipitation at high resolution [120,121,122], or (ii) the neutral ABL can be ill-represented (erroneous ABL height and wind-direction rotation, strong effect of resolution) [123].
Table 2. Failures in gray-zone atmospheric simulations. GZC refers to gray-zone coverage (usually through varying horizontal resolution). The horizontal blocks follow the summary paragraph in Section 2.1.
Table 2. Failures in gray-zone atmospheric simulations. GZC refers to gray-zone coverage (usually through varying horizontal resolution). The horizontal blocks follow the summary paragraph in Section 2.1.
Cases and ResolutionGray-Zone Performance and Requirements
Honnert et al. (2011) [112]; Honnert (2016) [113]: CBLs (five dry and cloudy cases); GZC; several mesoscale parametrizationsTurbulence schemes reveal drawbacks: resolved part is too large (or too weak) if the turbulence scheme is used without (or with) the mass-flux scheme; the turbulence scheme does not mix the boundary layer well enough, schemes need to account for 3D turbulence in gray-zone
Honnert (2018) [123]: neutral ABL; 3.2 × 3.2 × 1.5 km3; GZC; different turbulence and length scale modelsNeutral ABL is ill-represented (erroneous ABL height and wind-direction rotation, strong effect of resolution); most of the turbulence schemes are incorrect
Honnert et al. (2021) [124]: three boundary layer cases (a free convective case, a neutral case and a cold air outbreak case); GZC; mesoscale parametrizations; use of DES-type turbulence length scaleWith new DES-type mixing length, the turbulence scheme produces improved proportions between subgrid and resolved turbulent exchanges in LES, in the gray zone and at the mesoscale
Juliano et al. (2022) [114]: idealized cases (growing CBL over homogeneous land surface, sea-breeze front, mountain–valley circulation); GZC; 3D ABL schemeNew 3D ABL scheme is very promising; turbulent length scale model (and master length scale concept) need improvements, in particular for convective and stable cases
Efstathiou and Beare (2015) [116]: 3D dry CBL simulations; domain of at least 9.6 × 9.6 km2; ranging from the LES ( Δ x = 25 m) to the mesoscale limit ( Δ x = 3200 m)Excessive mixing due to the large values of mixing length at coarse resolutions results in the cut-off of resolved turbulence in the gray zone; the damping of resolved motions comes earlier for wind shear runs
Kealy et al. (2019) [117]: 3D LES of day 33 of the Wangara experiment (a widely studied CBL case); domain of 9.6 × 9.6 × 2.5 km3; ranging from the LES ( Δ x = 20 m) to gray-zone runs ( Δ x = 200 800 m)Without perturbations (initiation of resolved turbulence called spin-up), resolved turbulence does not become established at all; (1) applying structured perturbations, and (2) using a dynamically evolving Smagorinsky coefficient are shown to encourage faster spin-up
Beare and Efstathiou (2025) [118]: 3D CBL LES; case of Sullivan and Patton [125]; domain of 5120 × 5120 × 2048 m3, ranging from the LES ( Δ x = 50 m) to gray-zone runs ( Δ x = 500 1000 m)Models require an enhancement of diffusion to give a collapse of turbulence at Δ x h (h is the boundary-layer depth), whereas spin up of turbulence within the gray zone requires a reduced diffusion
De Roode et al. (2022) [119]: clear CBL, 12.8 × 12.8 × 1.594 km3; GZC; LESMesoscale structures (rolls vs. cells, turbulence flux profiles, etc.) depend strongly on horizontal grid resolution (and anisotropy between horizontal and vertical motions)
Sun et al. (2013) [120]: sensitivity of Typhoon Shanshan (2006) simulations to changes in horizontal grid spacing and to choices of convective parametrization schemes at gray-zone resolutions (1–7.5 km); WRF model is appliedTropical cyclone structure and intensity (eyewall slope, radius of maximum wind, latent heating, microphysics) vary significantly with resolution; results depend on convective parametrization schemes
Park et al. (2022) [121]: convective cell-related heavy rainfall in South Korea; effect of a scale-aware convective parametrization scheme (timescale and entrainment rates are adjusted based on the horizontal grid spacing) is compared with simpler schemes; gray-zone simulationsSimpler schemes imply overestimations and erroneous precipitation locations or an erroneous simulation of heavy precipitation at high resolution; scale-aware schemes can improve simulations although model parameter settings play a crucial role
Tomassini et al. (2023) [122]: several global cases (e.g., convection–circulation over Africa, Hurricane Dorian and Typhoon Goni, Darwin mesoscale convective systems case, Atlantic Extratropics) using 5 km resolutionVery similar observations as reported by Park et al. (2022) [121]: a scale-aware turbulence scheme and a carefully reduced and simplified mass-flux convection scheme outperform simpler schemes

2.2. Solution Concepts

Thorough reviews of ongoing activities to overcome the obvious shortcomings of dealing with gray-zone treatments in atmospheric simulations can be found elsewhere [22,115]; see also the references reported in Table 2. Specifically, there is work along the following lines:
1.
Incorporation of turbulence anisotropy: When the horizontal grid spacing is relatively coarse, horizontal (momentum, heat, and scalar) gradients are relatively small compared to vertical gradients, such that 1D (vertical) turbulence models are appropriate. But within the gray-zone (using finer grids), such horizontal gradients become nonnegligible and 3D effects of turbulence should be considered [114,123]. Thus, the inclusion of anisotropy is clearly an opportunity to substantially improve simulation results. A valuable extension of previously applied simulation methods has been presented recently by Juliano et al. [114]. Specifically, the authors presented a 3D ABL parametrization based on the level 2.5 model of Mellor and Yamada [126] and demonstrated the advantages in comparison with simpler methods. Additional attempts are described elsewhere [115].
2.
Turbulence length scale transport models: More specifically in regard to the inclusion of an appropriate turbulence length scale, it is worth noting the following. The predominant present approach to address this problem is the definition of appropriate algebraic length scales. An approach that can provide more generally applicable turbulence length scales is the use of two-equation turbulence models which enable (in one or the other way) the inclusion of length scales subject to local turbulence production and dissipation processes. Only a few attempts exist to address the problem in this way, see, e.g., Refs. [127,128,129,130,131,132,133,134].
3.
Turbulence length scale merging: Mesoscale and microscale regimes differ essentially through the inclusion of relatively large and small length scales, respectively. As is well-known, the inclusion of a length scale that appropriately describes the structure of the turbulence in the gray-zone represents a highly relevant challenge: the simple use of usually applied RANS or LES length scales is proven to be inappropriate. Based on existing concepts [135], a valuable extension of existing methods has been presented recently by Honnert et al. [124]. The basic approach is an appropriate merging of RANS and LES length scales, which enables significant simulation performance advantages. A discussion of a promising alternative way to address this question, which is based on a continuous length scale blending suggested by Senocak et al. [136], can be found elsewhere [27].
4.
Scale-aware parametrizations: A substantial disadvantage of using usually applied mesoscale parametrizations in the gray-zone is their scale independence. The latter assumes turbulent kinetic energies and length scales that are often way too large to properly represent gray-zone turbulence variables. It is now commonly accepted that scale-aware turbulence schemes (which include reference to the grid scale Δ ) are very efficient to overcome this problem. Several works in this regard are cited in Table 2, see Refs. [120,121,122]. We note that varying scale-awareness concepts are currently in use, e.g., McWilliams et al. require the following: “a scale-aware parameterization needs to set the proper energy or scalar transfer between the resolved and subgrid fields” [137].
5.
Initiation of resolved turbulence: In gray-zone simulations, the development of resolved flow structures is certainly not trivial. As is well known, the presence of excessive modeled motion can block (suppress) the development of resolved flow, which can cause significant shortcomings in applications [73]. Valuable attempts to overcome this issue have been reported, for example, in Refs. [117,118], where spin-up techniques have been presented to support the initiation of resolved turbulence. It has to be noted, however, that the success of these techniques turned out to be sensitive to the treatment of diffusion [118].

2.3. Solution Concepts: An Engineering Perspective

The orientation of current attempts to improve atmospheric gray-zone modeling by accounting for anisotropy of turbulence can be seen as being specific for the ABL. Such anisotropy modeling can be certainly involved in engineering simulations, but the usual focus is on the generation of meaningful resolved motion in separated turbulent flow simulations, having in mind that the use of simple models can be compensated by the inclusion of a sufficient amount of resolved motion. Apart from the inclusion of anisotropy, the current attempts to deal with the atmospheric gray-zone problem, basically, follow the engineering development of DES methods, usually without explicitly mentioning this fact. More specifically, by considering points 2–5 presented in Section 2.2, it is worth noting the following. The calculation of a turbulence length via a scale transport equation (via k ϵ or k ω models) is standard in engineering simulations, and the merging of RANS and LES length scales is the core concept of DES methods. Scale-aware parametrizations are applied depending on the concrete problem considered. The problem of dealing with blocked resolved motion or inappropriate contamination of RANS regions via resolved motion is a standard issue of DES methods.
What is the problem with such following of engineering DES concepts? The advantage of DES is its conceptual simplicity. However, the DES concept faces significant issues, as explained in the following:
P1.
Core problem: The core idea of DES methods is to enable a seamless transition between modeled (RANS) and resolved (LES) regimes based on a switch of RANS and LES length scales. This idea sounds perfectly fine as long as such a seamless transition is actually realized in simulations. Unfortunately, this is not always the case: the switch of length scales does often not imply a corresponding switch of simulation regimes. Specifically, a desired flow resolution imposed by the LES length scale does often not translate into a corresponding real flow resolution [73]. A usual practical problem is the blocking of resolved motion through modeled motion or an inappropriate contamination of RANS regions via resolved motion.
P2.
Uncertainty of proper approach and predictions: The DES concept can be applied with respect to most regularly implied turbulence models. It can be applied in a variety of versions including delayed DES (DDES), improved delayed DES (IDDES), in zonal and in non-zonal set-ups. Usually, the results significantly depend on the choice of these modeling options. DES is known to sometimes require schemes for the initialization of resolved motions at case-dependent locations. DES is known to be rather sensitive to different mesh distributions (using the same number of grid points) and model parameter settings. This modeling uncertainty often translates into a relatively large uncertainty range of simulation results [73,138].
P3.
Balanced simulation performance and computational cost: Usually, the uncertainty range of DES is used to set up simulations such that specific flow characteristics (like pressure or skin-friction distributions) are well characterized at the cost of the simulation performance in regard to other flow characteristics (like velocity and stress profiles). The fact that the real flow resolution accomplished by DES methods can be significantly below expectations [139,140] also has implications for the computational cost of DES methods. It is common practice to compensate simulation shortcomings by an increase in computational cost (finer grids), such that DES simulations can be very expensive [139,140,141].

3. Toward a Gray-Zone Problem Solution

Perspectives to overcome the gray-zone problem, which were developed in regard to engineering applications, will be presented next. First, we will address turbulence and turbulence length scale modeling. Then, an explicit solution concept for the gray-zone problem will be presented in conjunction with related computational evidence.

3.1. Turbulence Model

For clarity purposes, the following development will focus on two-equation, eddy viscosity-based turbulence models. However, this is not at all a required limitation. A model that includes the anisotropy effect (which generalizes the well-known Mellor–Yamada hierarchy [126]) is presented in Appendix A. Relevant facts to note in this regard are the following ones. First, the latter model is based on a realizable stochastic turbulence model. Second, the model implies significant implications for the master length scale concept suggested by Mellor and Yamada. Third, this model is fully compatible with length scale modeling as presented in Section 3.2 and Section 3.3.
The typical structure of a microscale model for the stratified ABL [13,142,143,144,145] is the following one. The i-th component of the spatially filtered velocity is denoted by U ˜ i , and Θ ˜ refers to a corresponding scalar (the potential temperature). We have the incompressible continuity equation U ˜ i / x i = 0 , and the momentum and potential temperature equations read
D U ˜ i D t = ( p ˜ / ρ + 2 k / 3 ) x i + 2 ( ν + ν t ) S ˜ i k x k ϵ i k n f k U ˜ n β θ Δ θ g i ,
D Θ ˜ D t = x k ν P r + ν t P r t Θ ˜ x k + S θ .
Here, D / D t = / t + U ˜ k / x k denotes the filtered Lagrangian time derivative and the sum convention is used throughout this paper. In addition, we have here the filtered pressure p ˜ , ρ is the constant mass density, k is the modeled energy, ν is the constant kinematic viscosity, and S ˜ i j = ( U ˜ i / x j + U ˜ j / x i ) / 2 is the rate-of-strain tensor. The modeled viscosity is given by ν t = C μ k 1 / 2 L . Here, C μ is a model parameter with standard value C μ = 0.09 , and L is a characteristic turbulence length scale. The latter can be related to the dissipation rate ϵ = k / τ of modeled kinetic energy, or the dissipation time scale τ , or turbulence frequency ω = 1 / τ by L = k 3 / 2 / ϵ = k 1 / 2 τ = k 1 / 2 / ω , respectively. On top of that we involve the gravity vector g i , β θ is the thermal expansion coefficient (the inverse reference temperature), Δ θ is the difference between a reference temperature and Θ ˜ , f k is the Coriolis vector, and ϵ i j k is the Levi-Civita symbol. In the Θ ˜ equation we have the molecular and turbulent heat diffusivities ν / P r and ν t / P r t , where P r and P r t , are the Prandtl number and turbulent Prandtl number, respectively. The effect of radiation can be taken into account via the source term S θ .
A turbulence model requires two main ingredients: a model for the kinetic energy of turbulence, and a model for the scale of turbulence. The first ingredient is usually provided by a transport equation for the modeled energy k. The second ingredient can be provided, e.g., on the basis of a transport equation for the length scale L (the dissipation rate ϵ , or the turbulence frequency ω ). Compared to corresponding algebraic models, the use of a transport equation for turbulence scale variables offers an essential advantage: L can be a realistically calculated subject to a potentially significant nonequilibrium between production and dissipation of turbulence, i.e., the calculation of L is not constrained by idealistic equilibrium ideas. Compared to k ϵ or k ω model developments, however, the development of k L models [107,108,126,146,147,148,149,150] is far behind; generally accepted suggestions do not exist so far. One reason for that is the large variety of suggested model formulations, which sometimes neglect the influence of production on L [149], require empirical extra terms to be consistent with the log-law [108,151], and often involve rather different diffusion contributions.

3.2. Hybrid Length Scale Model

According to the author’s knowledge, the implementation of any transport equation for a scale variable in atmospheric flow codes represents a major exception, and only k ϵ models were considered so far [133]. In line with that, personal communication indicates that the implementation of a k ϵ model in regularly applied atmospheric flow codes is far from trivial. At least in regard to atmospheric flow applications, the consideration of a transport equation for L offers essential advantages: (i) there are homogeneous wall boundary conditions since both k and L vanish at solid surfaces, and (ii) there is direct access to the turbulence length-scale, which is revealing with respect to turbulent flow analysis. Given the potential benefits, k L models will be considered in the following with the understanding that the related analysis will enable it to overcome the problems reported in the last paragraph of the preceding subsection.
For reasons to be discussed in the following paragraph, we consider here the following k k 1 / 2 L model as basis for the further development,
D k D t = P + P b ϵ + D k , D k L D t = C 1 L k P + P b α ϵ + D k L .
We note that k 1 / 2 L is proportional to the turbulent viscosity, i.e., the boundary conditions for this model are as simple as for the k L model. The diffusion terms read D k = [ ν t k / x j ] / x j , and D k L = [ ( ν t / σ ) ( k L ) / x j ] / x j , where σ is a model parameter. The shear production of kinetic energy is given by P = ν t S 2 , where S = ( 2 S ˜ m n S ˜ n m ) 1 / 2 is the characteristic shear rate, and the buoyancy production is given by P b = R i P / P r t where R i = β θ g Θ ˜ / x 3 S 2 is the gradient Richardson number [152,153,154,155]. The buoyancy term in the k L equation is sometimes multiplied with a coefficient C b . This buoyancy coefficient C b is characterized by a large uncertainty: its values range from 1.4 to + 1.45 [156,157] including C b = 1 [126]. For simplicity, we assume the buoyancy coefficient C b is equal to unity in the following in line with the recommendation of Mellor and Yamada [126]. It is worth noting that the turbulence model considered is consistent with the u i u j ˜ Equation (A3) presented in Appendix A based on k = u i u i ˜ / 2 . In regard to deriving the k equation from Equation (A3), the effect of the Coriolis term is neglected; the Coriolis effect is known to possibly affect the anisotropy of turbulence but not the kinetic energy [126,158]. The derivation of the k L equation from a k ϵ model reveals the relationships [159]
C 1 = 2 C ϵ 1 = 0.56 , C 1 α = 2 C ϵ 2 = 0.08 ,
which follows from using the k ϵ model standard values C ϵ 1 = 1.44 and C ϵ 2 = 1.92 .
Let us come back to the question of why the consideration of an k 1 / 2 L equation is beneficial compared to the direct consideration of a L equation. The exact k L model implied by Equation (3) is presented in Appendix B. It is shown there that this approach faces technical questions given by the switch of dissipation terms into production terms, and the need to provide dissipation through quadratic energy gradient terms. These problems are avoided by defining the transport of L via an equation for k 1 / 2 L .
Equation (3) is in line with Rotta’s derivation of a turbulence length scale equation [146,150], see Refs. [107,108]. These equations are also in line with the consideration of a generic turbulence model [159]. But there are two major problems related to Equation (3).
  • The first problem is that the equation structure is not completely specified as long as the concrete production–dissipation structure is unknown. The latter requires the specification of model parameters like α as functions of L or other variables. There are certainly questions about this. Mellor and Yamada [108,126,151] applied a version of this model which includes an empirical quadratic length scale ratio in α to ensure consistency with the log-law. A discussion of possibilities to actually specify this quadratic length scale ratio can be found elsewhere [160,161]. Rotta suggested a modification of the production term by involving the third-order velocity derivative [146,150]. Menter and Egorov [108] criticized Rotta’s approach. They suggested the implementation of another empirical quadratic length scale ratio (involving the von Kármán length scale) in the production coefficient C 1 . But there are different lines to argue regarding the concrete model formulation: a linear dependence on the corresponding second velocity derivative was proposed first [107], whereas a quadratic formulation has been used later [108].
  • The second significant problem related to Equation (3) is the model’s scale and resolution independence. The model as is is incapable of seamlessly covering LES and RANS regimes, i.e., the model cannot properly cover the gray-zone regime.
A way to overcome these issues related to existing gray-zone solution approaches in regard to atmospheric flow simulations will be presented next on the basis of continuous eddy simulation (CES) methods developed recently [73,138,139,140,144,145,159,162,163,164,165,166,167,168,169,170]. A rudimentary expectation for a model capable of providing a transition between resolved LES and modeled RANS regimes is its applicability to a wide ranges of scales. The latter can be accomplished by a modification of Equation (1), where the constant α is replaced by a variable model parameter α ,
D k D t = P + P b ϵ + D k , D k L D t = C 1 L k P + P b α ϵ + D k L .
Using variational analysis, it is possible to determine an exact expression for α given by
α = 1 + L 2 ( α 0 1 ) ,
where L = L / L 0 refers to the ratio of a model length scale (L) to a reference length scale L 0 . The details of this derivation can be found elsewhere [159]. No assumption regarding the reference length scale L 0 has to be made at this point, which can be chosen, e.g., according to DES concepts or as von Kármán length scale L v K = κ | S / S | (where κ is the von Kármán constant, S = ( 2 S ˜ m n S ˜ n m ) 1 / 2 is the characteristic shear rate and S = [ 2 U ˜ i / x k 2 × 2 U ˜ i / x j 2 ] 1 / 2 ). Equation (6) is of a purely technical nature, and a corresponding expression can be obtained by hybridizing the k-equation by introducing a variable model parameter in front of ϵ [159].
In regard to the two major problems described above, we observe the following:
  • As discussed above, current turbulence modeling focusing on L-type equations faces significant uncertainty because of the unknown required structure of model parameters involved. These problems are solved by the approach presented which mathematically determines the appearance of a quadratic length scale ratio in the turbulence model considered. The model obtained is consistent with log-law requirements as long as the model parameter relationship κ 2 = C μ 1 / 2 σ ( C 2 C 1 ) is satisfied, with κ being the von Kármán constant and C 2 = C 1 α 0 .
  • The expression α = 1 + L 2 ( α 0 1 ) is simply a consequence of ensuring the validity of the turbulence model considered in between a reference state (characterized by L 0 and the constant α 0 ) and a partially or fully resolving state characterized by L and α . The model functioning depends essentially on the setting of the reference length scale L 0 , which spans the hybridization range in between L 0 and L. Variants of performing this are discussed in detail elsewhere [159] by including DES, SAS, Mellor–Yamada-type, and LES scaling variants. However, the most natural, most effective setting of L 0 is L 0 = L t o t , where L t o t refers to the total length scale that accounts for resolved and modeled contributions (the calculation of L t o t and L + = L / L t o t is addressed in the Appendix C). This setting enables a seamless transition between LES and RANS regimes.

3.3. Advantages Compared to Existing Approaches

Given L 0 = L t o t , what are the essential advantages compared to DES type concepts used currently to address the atmospheric gray-zone problem? By taking reference to the DES problems P1–P3 reported in Section 2.3, we observe the following (see also the summary in Table 3).
P1.
Core problem: The DES core problem does not exist because there is no switch of RANS and LES length scales. More specifically, there is no switch at all, and there is no discrepancy between resolution driving parameters (LES and RANS length scales) and the actual flow resolution. Instead, the actual flow resolution is measured (in analogy to the corresponding energy ratio k + = k / k t o t ) via L + = L / L t o t , and the model is directly informed about the actual flow resolution via α . The latter enables the model to properly respond to variations in the actual resolved motion via increasing or decreasing its contribution to the simulation. Thus, there is no mismatch between resolved and modeled motion.
P2.
Uncertainty of proper approach and predictions: The variety of different DES versions that exist is implied by the shortcomings of DES: these DES modifications were presented to overcome these problems. This relates, e.g., to the main DES shortcoming to block resolved motion or to contaminate modeled flow regions with fluctuations, and the DES problem to sensitively depend on the mesh organization. These problems simply do not appear in the approach described here. First, there is no discrepancy between resolved and modeled motion because of the model adjustment mechanism. Second, there is no explicit grid dependence of the model presented; the grid-related filter width Δ does not enter the model.
P3.
Balanced simulation performance and computational cost: All previous applications of CES methods presented [138,139,140] reveal a very essential feature: the simulation predictions are well balanced. This means that predictions of a variety of relevant flow characteristics like separation zone characteristics, pressure and skin-friction distributions, profiles of mean velocities and stresses agree very well (and better than the results of other methods) with experimental data: there is simply no wiggle room for flow-dependent model adjustments as given in DES methods. This fact has remarkable implications for the computational cost of CES methods, which are found significantly below the cost of other methods [139,140].
The statements made in the preceding paragraph will be illustrated by applications of CES and DES methods to periodic hill flow simulations at a high R e = 37 K (the highest R e case for which experimental data are available for model validation). An illustration of this flow (a channel flow involving periodic restrictions) is provided by Figure 2 [138]. This flow considered involves features such as separation, recirculation, and natural reattachment [171,172]. It is used a lot for the evaluation of turbulence models [73] because of its reasonable balance of relative simplicity and complexity. DES applications were presented by Saini et al. [173]. Two variants of IDDES methods were considered: k ω -SST IDDES (an implementation of IDDES in Menter’s shear-stress transport (SST) k ω model [174]), and Spalart–Allmaras (S-A) IDDES (an implementation of IDDES in the Spalart–Allmaras turbulence model [175]). Two grids were involved in these simulations: a coarse 0.3 M grid, and a fine 0.9 M grid. CES applications were presented by Heinz et al. [138]. Three hybrid variants of a k ω model were considered: CES-KOS (hybridization in the ω scale equation), CES-KOK (hybridization in the k equation, and CES-KOKU (a hybridization according to unified RANS-LES). At the bottom and top, the channel is constrained by solid walls: no-slip and impermeability boundary conditions were used at these walls, and wall functions were not employed. Periodic boundary conditions are employed in the streamwise and spanwise directions. A significant range of R e ( R e = [ 37 , 250 , 500 ] K) and a variety of coarse and finer grids ( G 500 , G 250 , G 120 involving 500 K, 250 K, and 120 K grid points, respectively) were considered. It is worth noting that these grids cover the full range of almost modeled to almost completely resolving simulations.
Table 4 shows that DES is not as accurate as CES in regard to the calculation of separation zone characteristics (in particular the calculation of the reattachment point). We note that the CES grids are coarser than the DES grids, respectively. A disturbing fact is that the accuracy of DES significantly reduces if the grid is refined. What are the reasons for that? Figure 3 shows the f ˜ d distribution, which controls the transition from URANS to the LES regime, for the coarse grid for both models considered at x / h = 4 . The URANS region is defined by f ˜ d = 1 , the scale-resolving region (SRR) is defined by f ˜ d 0 , and the shaded intermediate region represents the gray-zone. It may be seen that the extend of the gray-zone is strongly influenced by the model applied. The expectation is that the corresponding distribution of k + = k / k t o t reflects this transition of f ˜ d from unity to zero. But Figure 4 (which shows coarse grid results) shows that this is not the case. The x / h = 4 curves show little variations, there is basically no difference between gray-zone and SRR behavior, this means the k + = distribution is not controlled by the f ˜ d distribution (it is of interest that the same can be seen in other approaches like PITM and PANS [138] where the resolution control is decoupled from the actual flow resolution). The fact that the k + values are rather small may lead to the expectation that flow simulation results are excellent because the flow is apparently very well resolved. Figure 5 shows that this is not the case. The velocity profiles appear to be reasonable, but stress profiles show significant shortcomings: there is a significant model and grid dependence, and the profiles can be qualitatively incorrect.
The CES features are very different from DES features. By design, there is no mismatch between the resolution control and the actual resolution. Figure 6 shows that the k + distributions are almost constant over the core of the flow. There are minor variations with the model considered and grid applied. As expected, a grid refinement implies a systematic reduction of k + . It is of interest to see the coincidence between k + and L + 2 / 3 . The latter can be explained by the exact relationship k + = L + 2 / 3 ϵ + 2 / 3 in conjunction with the fact that ϵ + is equal to unity except very close to the wall. Figure 7 shows implications for flow simulations. The streamwise velocity and stresses are very well predicted (the experimental stresses are known to be slightly underpredicted). It is remarkable that a grid coarsening has very little effect.
It is of interest that the conclusions of the CES-DES comparison reported for the periodic hill flow case considered here are fully in line with corresponding conclusions obtained for other flow simulations, e.g., the NASA wall-mounted hump flow and the axisymmetric transonic bump flow [139,140,162]. Flow-adjusted specific DES variants can provide good predictions of some flow characteristics but fail in regard to other flow characteristics (e.g., the separation zone characteristics or friction coefficient distributions). On the other hand, CES methods perform very well in regard to all relevant flow characteristics.

4. Conclusions

This paper presented an analysis of ongoing efforts to deal with the gray-zone problem in atmospheric simulations. The main conclusions can be summarized as follows.
1.
Significance and trend: There is currently no indication that simple fixes can overcome the gray-zone problem. The existing problems are serious. Without substantial changes in solution approaches, this problem has to be expected to significantly hamper future atmospheric simulations (possibly for decades to come). Specifically, current (DES-type) concepts to address this problem in regard to atmospheric simulations follow, basically, developments made by the engineering community years ago. However, there are clear indications that these developments significantly fall behind expectations: see the problems P1–P3 reported in Section 2.3.
2.
New modeling approach: There exist new (CES-type) simulation concepts developed in the context of engineering applications. Both theoretical features and applications provide at least strong indications that such CES-type simulations have the potential to overcome problems as seen in the context of DES applications. Specific technical features of the new modeling approach are (i) independence of the main LES restriction (i.e., the inclusion of the filter width Δ ), (ii) inclusion of an inherent actual resolution indicator ( L + ), and (iii) identification of conceptual issues of alternative methods (where different L 0 settings are used).
3.
Resolution-aware and anisotropy-aware modeling: One of the most essential problems of atmospheric simulations is to provide appropriate (gray-zone applicable) turbulence length scales. The analysis presented identifies a resolution-aware length scale equation where significant uncertainty of the structure of previous k L type modeling efforts (e.g., regarding the model parameter structure) is excluded. In addition, as described in the Appendix A, a significant model extension covering the structure of the Mellor–Yamada hierarchy is well possible. An attractive feature of the model presented is the identification of length-scale relationships, which generalize the Mellor–Yamada master length scale concept.

Funding

This research was funded by the National Science Foundation USA (AGS, Grant No. 2137351, with N. Anderson as Technical Officer).

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

I am thankful to B. Kosović, T. Juliano, and J. Singh for related work. This work was supported by Wyoming NASA Space Grant Consortium (NASA Grant No. 80NSSC20M0113) and the University of Wyoming School of Computing (Wyoming Innovation Partnership grant).

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Mellor–Yamada Hierarchy

A realizable stochastic model [145,152,168] can be used to derive the following transport equations for filtered velocities and scalars,
D U ˜ i D t = ( p ˜ / ρ ) x i + 2 ν S ˜ i k x k u i u k ˜ x k ϵ i k n f k U ˜ n β θ Δ θ g i ,
D Θ ˜ D t = x k ν P r Θ ˜ x k u k θ ˜ x k + S θ .
These equations generalize Equations (1) and (2) by the appearance of the stress u i u k ˜ and turbulent scalar flux u k θ ˜ , for which no eddy viscosity assumption is made. To complement Equations (A1) and (A2), we also use the realizable stochastic model to derive exact equations for u i u k ˜ and u k θ ˜ (and θ 2 ˜ which appears in the turbulent scalar equation),
D u i u j ˜ D t + τ i j k x k P i j + β θ g j u i θ ˜ + g i u j θ ˜ + f k ϵ j k n u n u i ˜ + ϵ i k n u n u j ˜ =
= c 1 τ u i u j ˜ 2 k 3 δ i j + 2 c 2 U ˜ i x k u k u j ˜ + P 3 δ i j 2 3 c 1 + c 2 P ϵ 3 2 C 0 ϵ δ i j ,
D u i θ ˜ D t + τ i k θ x k P i θ + β θ g i θ 2 ˜ + f k ϵ i k n u n θ ˜ = c 1 2 τ + c θ 1 τ θ u i θ ˜ + c 2 u k θ ˜ U ˜ i x k ,
D θ 2 ˜ D t + τ k θ θ x k 2 P θ = 2 c θ 1 τ θ θ 2 ˜ .
We have here P = u k u n ˜ U ˜ n / x k , and τ i j k , τ i j θ , τ i θ θ refer to corresponding triple correlations, and we introduced for production terms the abbreviations
P i j = u k u j ˜ U ˜ i x k u k u i ˜ U ˜ j x k , P i θ = u i u k ˜ Θ ˜ x k u k θ ˜ U ˜ i x k , P θ = u k θ ˜ Θ ˜ x k .
The scalar mixing frequency is given by c θ 1 / τ θ , i.e., this mixing frequency is not assumed to be proportional to τ 1 ( τ θ refers to the scalar dissipation time scale, it is related to the scalar dissipation rate by ϵ θ = k θ / τ θ , where k θ = θ 2 ˜ / 2 ). In addition, c 1 , c 2 , c θ 1 , and C 0 are stochastic model parameters. The consistency of these equations with the definitions of dissipation rates ϵ = k / τ and ϵ θ = k θ / τ θ in the transport equations for k = u n u n ˜ / 2 and k θ = θ 2 ˜ / 2 requires two coefficient relations,
c 1 + c 2 P / ϵ 3 C 0 / 2 = 1 , c θ 1 = 1 / 2 .
It is essential to note that Equations (A3)–(A5) agree with the level 4 closure model of Mellor and Yamada [126] with the exception of Mellor and Yamada’s neglect of higher-order anisotropy effects. Specifically, the box terms in Equations (A3) and (A4) are differently modeled by Mellor and Yamada, leading to an isotropic version of the c 2 term on the RHS of Equation (A3) (which reads then 8 k c 2 S ˜ i j / 3 ) and the neglect of the last term in Equation (A4). A significant advantage of Equations (A3)–(A5) is the ability to calculate the closure parameters in Mellor and Yamada’s model. A comparison of coefficients applied here with coefficients 1 , 2 , Λ 1 , Λ 2 applied by Mellor and Yamada reveals in conjunction with c θ 1 = 1 / 2 the relations
c 1 τ = ( 2 k ) 1 / 2 3 1 , 4 c 2 3 = 2 C 1 , ϵ = ( 2 k ) 3 / 2 Λ 1 , c 1 2 τ + 1 2 τ θ = ( 2 k ) 1 / 2 3 2 , c θ 1 τ θ = ( 2 k ) 1 / 2 Λ 2 .
We find in addition to the relationship C 1 = 2 c 2 / 3 the following conditions for 1 , 2 , Λ 1 , Λ 2 ,
( 1 , Λ 1 , 2 , Λ 2 ) = 1 2 c 1 , 3 , 1 c 1 + Γ , 3 2 c θ 1 Γ 2 3 / 2 3 L .
L = k 1 / 2 τ and the mechanical to scalar time scale ratio Γ = τ / τ θ are involved here (a constant Γ = 1.5 was applied for stratified flow [176,177]). It may be seen, therefore, that the model parameters 1 , Λ 1 involved in Mellor and Yamada’s model are not constant, but they vary with the production-to-dissipation ratio according to Equation (A7).

Appendix B. Implications for Length Scale Equation

It is of interest to compare Equation (3) with the k L model that is implied by Equation (3). On the basis of D ( k L ) / D t = k D L / D t + L D k / D t / [ 2 k ] we obtain
D k D t = P + P b ϵ + D k , D L D t = L k P + P b C 1 1 2 ϵ C 1 α 1 2 + F L + D L .
The nondimensional F L reads,
F L = C μ L 2 1 4 k 2 k x j k x j 1 L k k x j L x j ,
and the diffusion coefficient is given by D L = [ ( ν t / σ L ) L / x j ] / x j . To simplify matters, the only assumption applied to derive Equation (A10) is the following: we applied σ L = σ k ( σ k is the corresponding diffusion factor in the diffusion term of the k equation, which is always considered to be equal to unity). The consequence of not making this assumption is the appearance of spurious source terms [159]. In conjunction with Equation (4), we see that C 1 α 1 / 2 becomes negative, meaning the dissipation sink term turns into a production term. On the other hand, a dissipation term appears based on the first term of Equation (A11). The second (cross-diffusion) term in Equation (A11) can be seen to be negligible, see the discussion in Ref. [108]. Thus, the immediate consideration of an k L model faces significantly enhanced difficulty of applications and uncertainty.

Appendix C. Resolution Indicators

The resolution indicator L + = L / L t o t can be calculated in the following way. The modeled contribution is calculated by L = k 3 / 2 / ϵ , where the brackets refer to averaging in time. The total length scale is calculated correspondingly by L t o t = k t o t 3 / 2 / ϵ t o t . Here, k t o t = k + k r e s is the sum of modeled and resolved contributions, where the resolved contribution is calculated by
k r e s = 1 2 U ˜ i U ˜ i U ˜ i U ˜ i .
Correspondingly, ϵ t o t is the sum of modeled and resolved contributions, ϵ t o t = ϵ + ϵ r e s , where
ϵ r e s = ν U ˜ i x j U ˜ i x j U ˜ i x j U ˜ i x j .
Both k r e s and ϵ r e s are calculated during the simulation by using Equations (A12) and (A13). We note that L + is related to the corresponding kinetic energy ratio k + = k / k t o t by L + = k + 3 / 2 / ϵ + , where ϵ + = ϵ / ϵ t o t . Away from walls, ϵ + = 1 represents a good approximation which implies L + = k + 3 / 2 . By construction we have 0 L + 1 where L + = 0 refers to fully resolved flow and L + = 1 to fully modeled flow.

References

  1. Muñoz-Esparza, D.; Kosović, B.; Mirocha, J.; van Beeck, J. Bridging the transition from mesoscale to microscale turbulence in numerical weather prediction models. Bound.-Layer Meteorol. 2014, 153, 409–440. [Google Scholar] [CrossRef]
  2. Sanz Rodrigo, J.; Chavez Arroyo, R.A.; Moriarty, P.; Churchfield, M.; Kosović, B.; Réthoré, P.E.; Hansen, K.S.; Hahmann, A.; Mirocha, J.D.; Rife, D. Mesoscale to microscale wind farm flow modeling and evaluation. Wires Energy Environ. 2017, 6, e214. [Google Scholar] [CrossRef]
  3. Mazzaro, L.J.; Muñoz-Esparza, D.; Lundquist, J.K.; Linn, R.R. Nested mesoscale-to-LES modeling of the atmospheric boundary layer in the presence of under-resolved convective structures. J. Adv. Model Earth Syst. 2017, 9, 1795–1810. [Google Scholar] [CrossRef]
  4. Veers, P.; Dykes, K.; Lantz, E.; Barth, S.; Bottasso, C.L.; Carlson, O.; Clifton, A.; Green, J.; Green, P.; Holttinen, H.; et al. Grand challenges in the science of wind energy. Science 2019, 366, eaau2027. [Google Scholar] [CrossRef] [PubMed]
  5. Meneveau, C. Big wind power: Seven questions for turbulence research. J. Turbul. 2019, 20, 2–20. [Google Scholar] [CrossRef]
  6. Haupt, S.E.; Kosović, B.; Shaw, W.; Berg, L.K.; Churchfield, M.; Cline, J.; Draxl, C.; Ennis, B.; Koo, E.; Kotamarthi, R.; et al. On bridging a modeling scale gap: Mesoscale to microscale coupling for wind energy. Bull. Amer. Meteor. Soc. 2019, 100, 2533–2550. [Google Scholar] [CrossRef]
  7. Marjanovic, N. Mesoscale and Large-Eddy Simulations for Wind Energy; Technical Report for Lawrence Livermore National Lab. (LLNL): Livermore, CA, USA, 2011. [Google Scholar]
  8. Zhong, S.; Chow, F.K. Meso-and fine-scale modeling over complex terrain: Parameterizations and applications. In Mountain Weather Research and Forecasting; Springer: Berlin/Heidelberg, Germany, 2013; pp. 591–653. [Google Scholar]
  9. Mehta, D.; Van Zuijlen, A.H.; Koren, B.; Holierhoek, J.G.; Bijl, H. Large Eddy Simulation of wind farm aerodynamics: A review. J. Wind Eng. Ind. Aerod. 2014, 133, 1–17. [Google Scholar] [CrossRef]
  10. Ito, J.; Niino, H.; Nakanishi, M.; Moeng, C.H. An extension of the Mellor–Yamada model to the terra incognita zone for dry convective mixed layers in the free convection regime. Bound.-Layer Meteorol. 2015, 157, 23–43. [Google Scholar] [CrossRef]
  11. Temel, O.; van Beeck, J. Adaptation of mesoscale turbulence parameterisation schemes as RANS closures for ABL simulations. J. Turbul. 2016, 17, 966–997. [Google Scholar] [CrossRef]
  12. Breton, S.P.; Sumner, J.; Sørensen, J.N.; Hansen, K.S.; Sarmast, S.; Ivanell, S. A survey of modelling methods for high-fidelity wind farm simulations using large eddy simulation. Phil. Trans. Math. Phys. Eng. Sci. 2017, 375, 20160097. [Google Scholar] [CrossRef]
  13. Temel, O.; Porchetta, S.; Bricteux, L.; van Beeck, J. RANS closures for non-neutral microscale CFD simulations sustained with inflow conditions acquired from mesoscale simulations. Appl. Math. Model. 2018, 53, 635–652. [Google Scholar] [CrossRef]
  14. Mazzaro, L.J. Mesoscale-to-LES Coupling of Atmospheric Boundary Layer Flows. Ph.D. Thesis, University of Colorado, Boulder, CO, USA, 2019. [Google Scholar]
  15. Chow, F.K.; Schär, C.; Ban, N.; Lundquist, K.A.; Schlemmer, L.; Shi, X. Crossing multiple gray zones in the transition from mesoscale to microscale simulation over complex terrain. Atmosphere 2019, 10, 274. [Google Scholar] [CrossRef]
  16. Rodrigo, J.S.; Churchfield, M.; Kosović, B. Atmospheric boundary layer modeling based on mesoscale tendencies and data assimilation at microscale. Wind Energy Sci. 2016, 2, 35–54. [Google Scholar]
  17. Muñoz-Esparza, D.; Lundquist, J.K.; Sauer, J.A.; Kosović, B.; Linn, R.R. Coupled mesoscale-LES modeling of a diurnal cycle during the CWEX-13 field campaign: From weather to boundary-layer eddies. J. Adv. Model. Earth Syst. 2017, 9, 1572–1594. [Google Scholar] [CrossRef]
  18. Rai, R.K.; Berg, L.K.; Kosović, B.; Mirocha, J.D.; Pekour, M.S.; Shaw, W.J. Comparison of measured and numerically simulated turbulence statistics in a convective boundary layer over complex terrain. Bound.-Layer Meteorol. 2017, 163, 69–89. [Google Scholar] [CrossRef]
  19. Mirocha, J.D.; Churchfield, M.J.; Muñoz-Esparza, D.; Rai, R.K.; Feng, Y.; Kosovic, B.; Haupt, S.E.; Brown, B.; Ennis, B.L.; Draxl, C.; et al. Large-eddy simulation sensitivities to variations of configuration and forcing parameters in canonical boundary-layer flows for wind energy applications. Wind Energy Sci. Discuss. 2018, 3, 589–613. [Google Scholar] [CrossRef]
  20. Rai, R.K.; Berg, L.K.; Kosović, B.; Haupt, S.E.; Mirocha, J.D.; Ennis, B.L.; Draxl, C. Evaluation of the impact of horizontal grid spacing in terra incognita on coupled mesoscale–microscale simulations using the WRF framework. Mon. Weather Rev. 2019, 147, 1007–1027. [Google Scholar] [CrossRef]
  21. Quon, E.W.; Ghate, A.S.; Lele, S.K. Enrichment methods for inflow turbulence generation in the atmospheric boundary layer. J. Phys. Conf. Ser. 2018, 1037, 072054. [Google Scholar] [CrossRef]
  22. Doubrawa, P.; Muñoz-Esparza, D. Simulating real atmospheric boundary layers at gray-zone resolutions: How do currently available turbulence parameterizations perform? Atmosphere 2020, 11, 345. [Google Scholar] [CrossRef]
  23. Haupt, S.E.; Berg, L.; Churchfield, M.; Kosović, B.; Mirocha, J.; Shaw, W. Mesoscale to Microscale Coupling for Wind Energy Applications: Addressing the Challenges. J. Phys. Conf. Ser. 2020, 1452, 012076. [Google Scholar] [CrossRef]
  24. Santoni, C.; García-Cartagena, E.J.; Ciri, U.; Zhan, L.; Valerio Iungo, G.; Leonardi, S. One-way mesoscale-microscale coupling for simulating a wind farm in North Texas: Assessment against SCADA and LiDAR data. Wind Energy 2020, 23, 691–710. [Google Scholar] [CrossRef]
  25. Bauer, H.S.; Muppa, S.K.; Wulfmeyer, V.; Behrendt, A.; Warrach-Sagi, K.; Späth, F. Multi-nested WRF simulations for studying planetary boundary layer processes on the turbulence-permitting scale in a realistic mesoscale environment. Tellus A 2020, 72, 1–28. [Google Scholar] [CrossRef]
  26. Wiersema, D.J.; Lundquist, K.A.; Chow, F.K. Mesoscale to microscale simulations over complex terrain with the immersed boundary method in the Weather Research and Forecasting Model. Mon. Weather Rev. 2020, 148, 577–595. [Google Scholar] [CrossRef]
  27. Wiersema, D. Mesoscale to Microscale Atmospheric Modeling Over Complex Terrain. Ph.D. Thesis, University of California, Berkeley, CA, USA, 2020. [Google Scholar]
  28. Haupt, S.E.; Anderson, A.; Berg, L.; Brown, B.; Churchfield, M.J.; Draxl, C.; Ennis, B.L.; Feng, Y.; Kosović, B.; Kotamarthi, R.; et al. First Year Report of the A2e Mesoscale to Microscale Coupling Project; Technical report no. PNNL-25108; Pacific Northwest National Laboratory (PNNL): Richland, WA, USA, 2015. [Google Scholar]
  29. Haupt, S.E.; Kotamarthi, R.; Feng, Y.; Mirocha, J.D.; Koo, E.; Linn, R.; Kosović, B.; Brown, B.; Anderson, A.; Churchfield, M.J.; et al. Second Year Report of the Atmosphere to Electrons Mesoscale to Microscale Coupling Project: Nonstationary Modeling Techniques and Assessment; Technical Report; Pacific Northwest National Laboratory (PNNL): Richland, WA, USA, 2017. [Google Scholar]
  30. Wyngaard, J.C. Toward numerical modeling in the “Terra Incognita”. J. Atmos. Sci. 2004, 61, 1816–1826. [Google Scholar]
  31. Deardorff, J.W. A numerical study of three-dimensional turbulent channel flow at large Reynolds numbers. J. Fluid Mech. 1970, 41, 453–480. [Google Scholar] [CrossRef]
  32. Schumann, U. Subgrid scale model for finite difference simulations of turbulent flows in plane channels and annuli. J. Comput. Phys. 1975, 18, 376–404. [Google Scholar] [CrossRef]
  33. Grötzbach, G. Direct numerical and large eddy simulation of turbulent channel flow. Encycl. Fluid Mech. 1987, 6, 1337–1391. [Google Scholar]
  34. Piomelli, U.; Ferziger, J.; Moin, P. New approximate boundary conditions for large eddy simulations of wall-bounded flows. Phys. Fluids A 1989, 6, 1061–1068. [Google Scholar] [CrossRef]
  35. Cabot, W.; Moin, P. Approximate wall boundary conditions in the large-eddy simulation of high Reynolds number flow. Flow Turbul. Combust. 1999, 63, 269–291. [Google Scholar] [CrossRef]
  36. Sagaut, P. Large Eddy Simulation for Incompressible Flows: An Introduction; Springer: Berlin/Heidelberg, Germany, 2002. [Google Scholar]
  37. Piomelli, U.; Balaras, E. Wall-layer models for large-eddy simulations. Annu. Rev. Fluid Mech. 2002, 34, 349–374. [Google Scholar] [CrossRef]
  38. Piomelli, U. Wall-layer models for large-eddy simulations. Prog. Aerosp. Sci. 2008, 44, 437–446. [Google Scholar] [CrossRef]
  39. Kawai, S.; Larsson, J. Wall-modeling in large eddy simulation: Length scales, grid resolution, and accuracy. Phys. Fluids 2012, 24, 015105. [Google Scholar] [CrossRef]
  40. Bose, S.T.; Moin, P. A dynamic slip boundary condition for wall-modeled large-eddy simulation. Phys. Fluids 2014, 26, 015104. [Google Scholar] [CrossRef]
  41. Park, G.I.; Moin, P. An improved dynamic non-equilibrium wall-model for large eddy simulation. Phys. Fluids 2014, 26, 015108. [Google Scholar] [CrossRef]
  42. Piomelli, U. Large eddy simulations in 2030 and beyond. Phil. Trans. R. Soc. A 2014, 372, 20130320. [Google Scholar] [CrossRef]
  43. Moin, P.; Bodart, J.; Bose, S.; Park, G.I. Wall-modeling in complex turbulent flows. In Advances in Fluid-Structure Interaction: Updated Contributions Reflecting New Findings, Proceedings of the ERCOFTAC Symposium on Unsteady Separation in Fluid-Structure Interaction, Mykonos, Greece, 17–21 June 2013; Braza, M., Bottaro, A., Thompson, M., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 207–219. [Google Scholar]
  44. Larsson, J.; Kawai, S.; Bodart, J.; Bermejo-Moreno, I. Large eddy simulation with modeled wall-stress: Recent progress and future directions. Mech. Eng. Rev. 2016, 3, 15–00418. [Google Scholar] [CrossRef]
  45. Bose, S.T.; Park, G.I. Wall-modeled large-eddy simulation for complex turbulent flows. Annu. Rev. Fluid Mech. 2018, 50, 535–561. [Google Scholar] [CrossRef]
  46. Yang, X.; Sadique, J.; Mittal, R.; Meneveau, C. Integral wall model for large eddy simulations of wall-bounded turbulent flows. Phys. Fluids 2015, 27, 025112. [Google Scholar] [CrossRef]
  47. Chen, S.; Xia, Z.; Pei, S.; Wang, J.; Yang, Y.; Xiao, Z.; Shi, Y. Reynolds-stress-constrained large-eddy simulation of wall-bounded turbulent flows. J. Fluid Mech. 2012, 703, 1–28. [Google Scholar] [CrossRef]
  48. Chen, S.; Chen, Y.; Xia, Z.; Qu, K.; Shi, Y.; Xiao, Z.; Liu, Q.; Cai, Q.; Liu, F.; Lee, C.; et al. Constrained large-eddy simulation and detached eddy simulation of flow past a commercial aircraft at 14 degrees angle of attack. Sci. China Phys. Mech. Astron. 2013, 56, 270–276. [Google Scholar] [CrossRef]
  49. Xia, Z.; Shi, Y.; Hong, R.; Xiao, Z.; Chen, S. Constrained large-eddy simulation of separated flow in a channel with streamwise-periodic constrictions. J. Turbul. 2013, 14, 1–21. [Google Scholar] [CrossRef]
  50. Jiang, Z.; Xiao, Z.; Shi, Y.; Chen, S. Constrained large-eddy simulation of wall-bounded compressible turbulent flows. Phys. Fluids 2013, 25, 106102. [Google Scholar] [CrossRef]
  51. Hong, R.; Xia, Z.; Shi, Y.; Xiao, Z.; Chen, S. Constrained large-eddy simulation of compressible flow past a circular cylinder. Commun. Comput. Phys. 2014, 15, 388–421. [Google Scholar] [CrossRef]
  52. Zhao, Y.; Xia, Z.; Shi, Y.; Xiao, Z.; Chen, S. Constrained large-eddy simulation of laminar-turbulent transition in channel flow. Phys. Fluids 2014, 26, 095103. [Google Scholar] [CrossRef]
  53. Xua, Q.; Yang, Y. Reynolds stress constrained large eddy simulation of separation flows in a U-duct. J. Propul. Power Res. 2014, 3, 49–58. [Google Scholar] [CrossRef]
  54. Xia, Z.; Xiao, Z.; Shi, Y.; Chen, S. Constrained large-eddy simulation for aerodynamics. In Progress in Hybrid RANS-LES Modelling: Papers in Proceedings of the 5th Symposium on Hybrid RANS-LES Methods, A&M University, College Station, TX, USA, 19–21 March 2014; Girimaji, S., Haase, W., Peng, S.H., Schwamborn, D., Eds.; Springer International Publishing: Cham, Switzerland, 2015; pp. 239–254. [Google Scholar]
  55. Jiang, Z.; Xiao, Z.; Shi, Y.; Chen, S. Constrained large-eddy simulation of turbulent flow and heat transfer in a stationary ribbed duct. Int. J. Numer. Methods Heat Fluid Flow 2016, 26, 1069–1091. [Google Scholar] [CrossRef]
  56. Verma, A.; Park, N.; Mahesh, K. A hybrid subgrid-scale model constrained by Reynolds stress. Phys. Fluids 2013, 25, 110805. [Google Scholar] [CrossRef]
  57. Xiao, Z.; Shi, Y.; Xia, Z.; Chen, S. Comment on “A hybrid subgrid-scale model constrained by Reynolds stress”. Phys. Fluids 2014, 26, 059101. [Google Scholar] [CrossRef]
  58. Verma, A.; Park, N.; Mahesh, K. Response to “Comment on ’A hybrid subgrid-scale model constrained by Reynolds stress”. Phys. Fluids 2014, 26, 059102. [Google Scholar] [CrossRef]
  59. Spalart, P.R.; Jou, W.H.; Strelets, M.; Allmaras, S.R. Comments on the feasibility of LES for wings, and on a hybrid RANS/LES approach. In Proceedings of the Advances in DNS/LES, Ruston, LA, USA, 4–8 August 1997; Liu, C., Liu, Z., Eds.; Greyden Press: Columbus, OH, USA, 1997; pp. 137–147. [Google Scholar]
  60. Travin, A.; Shur, M.L.; Strelets, M.; Spalart, P. Detached-eddy simulations past a circular cylinder. Flow Turbul. Combust. 1999, 63, 113–138. [Google Scholar] [CrossRef]
  61. Spalart, P.R. Strategies for turbulence modelling and simulations. Int. J. Heat Fluid Flow 2000, 21, 252–263. [Google Scholar] [CrossRef]
  62. Strelets, M. Detached eddy simulation of massively separated flows. In Proceedings of the 39th AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 8–11 January 2001; AIAA Paper 01-0879. pp. 1–18. [Google Scholar]
  63. Travin, A.; Shur, M.L. Physical and numerical upgrades in the detached-eddy simulation of complex turbulent flows. In Advances in LES of Complex Flows, Proceedings of the Euromech Colloquium 412, Munich, Germany, 4–6 October 2000; Friedrich, R., Rodi, W., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2002; pp. 239–254. [Google Scholar]
  64. Menter, F.R.; Kuntz, M.; Langtry, R. Ten years of industrial experience with SST turbulence model. Turb. Heat Mass Transf. 2003, 4, 625–632. [Google Scholar]
  65. Spalart, P.R.; Deck, S.; Shur, M.L.; Squires, K.D.; Strelets, M.K.; Travin, A. A new version of detached-eddy simulation, resistant to ambiguous grid densities. Theor. Comput. Fluid Dyn. 2006, 20, 181–195. [Google Scholar] [CrossRef]
  66. Shur, M.L.; Spalart, P.R.; Strelets, M.K.; Travin, A. A hybrid RANS-LES approach with delayed-DES and wall-modelled LES capabilities. Int. J. Heat Fluid Flow 2008, 29, 1638–1649. [Google Scholar] [CrossRef]
  67. Spalart, P.R. Detached-eddy simulation. Annu. Rev. Fluid Mech. 2009, 41, 181–202. [Google Scholar] [CrossRef]
  68. Mockett, C.; Fuchs, M.; Thiele, F. Progress in DES for wall-modelled LES of complex internal flows. Comput. Fluids 2012, 65, 44–55. [Google Scholar] [CrossRef]
  69. Friess, C.; Manceau, R.; Gatski, T.B. Toward an equivalence criterion for hybrid RANS/LES methods. Comput. Fluids 2015, 122, 233–246. [Google Scholar] [CrossRef]
  70. Chaouat, B. The state of the art of hybrid RANS/LES modeling for the simulation of turbulent flows. Flow Turbul. Combust. 2017, 99, 279–327. [Google Scholar] [CrossRef]
  71. Speziale, C.G. Turbulence modeling for time-dependent RANS and VLES: A review. AIAA J. 1998, 36, 173–184. [Google Scholar] [CrossRef]
  72. Fröhlich, J.; Terzi, D.V. Hybrid LES/RANS methods for the simulation of turbulent flows. Prog. Aerosp. Sci. 2008, 44, 349–377. [Google Scholar] [CrossRef]
  73. Heinz, S. A review of hybrid RANS-LES methods for turbulent flows: Concepts and applications. Prog. Aerosp. Sci. 2020, 114, 100597. [Google Scholar] [CrossRef]
  74. Menter, F.; Hüppe, A.; Matyushenko, A.; Kolmogorov, D. An overview of hybrid RANS–LES models developed for industrial CFD. Appl. Sci. 2021, 11, 2459. [Google Scholar] [CrossRef]
  75. Heinz, S. Unified turbulence models for LES and RANS, FDF and PDF simulations. Theoret. Comput. Fluid Dynam. 2007, 21, 99–118. [Google Scholar] [CrossRef]
  76. Heinz, S. Realizability of dynamic subgrid-scale stress models via stochastic analysis. Monte Carlo Methods Appl. 2008, 14, 311–329. [Google Scholar] [CrossRef]
  77. Heinz, S.; Gopalan, H. Realizable versus non-realizable dynamic subgrid-scale stress models. Phys. Fluids 2012, 24, 115105. [Google Scholar] [CrossRef]
  78. Gopalan, H.; Heinz, S.; Stöllinger, M. A unified RANS-LES model: Computational development, accuracy and cost. J. Comput. Phys. 2013, 249, 249–279. [Google Scholar] [CrossRef]
  79. Mokhtarpoor, R.; Heinz, S.; Stoellinger, M. Dynamic unified RANS-LES simulations of high Reynolds number separated flows. Phys. Fluids 2016, 28, 095101. [Google Scholar] [CrossRef]
  80. Mokhtarpoor, R.; Heinz, S. Dynamic large eddy simulation: Stability via realizability. Phys. Fluids 2017, 29, 105104. [Google Scholar] [CrossRef]
  81. Ahmadi, G.; Kassem, H.; Stoevesandt, B.; Peinke, J.; Heinz, S. Wall bounded turbulent flows up to high Reynolds numbers: LES resolution assessment. In Proceedings of the AIAA Scitech 2019 Forum, Orlando, FL, USA, 7–11 January 2019; AIAA Paper 19-1887. pp. 1–9. [Google Scholar]
  82. Girimaji, S.; Srinivasan, R.; Jeong, E. PANS turbulence for seamless transition between RANS and LES: Fixed-point analysis and preliminary results. In Proceedings of the ASME FEDSM03, Honolulu, HI, USA, 6–10 July 2003; ASME Paper FEDSM2003-45336. pp. 1–9. [Google Scholar]
  83. Girimaji, S.; Abdol-Hamid, K. Partially averaged Navier Stokes model for turbulence: Implemantation and validation. In Proceedings of the 43rd AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 10–13 January 2005; AIAA Paper 05-0502. pp. 1–14. [Google Scholar]
  84. Girimaji, S. Partially-averaged Navier-Stokes method for turbulence: A Reynolds-averaged Navier-Stokes to direct numerical simulation bridging method. ASME J. Appl. Mech. 2006, 73, 413–421. [Google Scholar] [CrossRef]
  85. Girimaji, S.; Jeong, E.; Srinivasan, R. Partially averaged Navier-Stokes method for turbulence: Fixed point analysis and comparisons with unsteady partially averaged Navier-Stokes. ASME J. Appl. Mech. 2006, 73, 422–429. [Google Scholar] [CrossRef]
  86. Lakshmipathy, S.; Girimaji, S.S. Extension of Boussinesq turbulence constitutive relation for bridging methods. J. Turbul. 2007, 8, 1–20. [Google Scholar] [CrossRef]
  87. Frendi, A.; Tosh, A.; Girimaji, S. Flow past a backward-facing step: Comparison of PANS, DES and URANS results with experiments. Int. J. Comput. Methods Eng. Sci. Mech. 2007, 8, 23–38. [Google Scholar] [CrossRef]
  88. Lakshmipathy, S.; Girimaji, S.S. Partially averaged Navier-Stokes (PANS) method for turbulence simulations: Flow past a circular cylinder. ASME J. Fluids Eng. 2010, 132, 121202. [Google Scholar] [CrossRef]
  89. Jeong, E.; Girimaji, S.S. Partially averaged Navier–Stokes (PANS) method for turbulence simulations: Flow past a square cylinder. ASME J. Fluids Eng. 2010, 132, 121203. [Google Scholar] [CrossRef]
  90. Basara, B.; Krajnovic, S.; Girimaji, S.S.; Pavlovic, Z. Near-wall formulation of the partially averaged Navier-Stokes turbulence model. AIAA J. 2011, 42, 2627–2636. [Google Scholar] [CrossRef]
  91. Krajnovic, S.; Lárusson, R.; Basara, B. Superiority of PANS compared to LES in predicting a rudimentary landing gear flow with affordable meshes. Int. J. Heat Fluid Flow 2012, 37, 109–122. [Google Scholar] [CrossRef]
  92. Foroutan, H.; Yavuzkurt, S. A partially averaged Navier Stokes model for the simulation of turbulent swirling flow with vortex breakdown. Int. J. Heat Fluid Flow 2014, 50, 402–416. [Google Scholar] [CrossRef]
  93. Drikakis, D.; Sofos, F. Can artificial intelligence accelerate fluid mechanics research? Fluids 2023, 8, 212. [Google Scholar] [CrossRef]
  94. Schiestel, R.; Dejoan, A. Towards a new partially integrated transport model for coarse grid and unsteady turbulent flow simulations. Theor. Comput. Fluid Dyn. 2005, 18, 443–468. [Google Scholar] [CrossRef]
  95. Chaouat, B.; Schiestel, R. A new partially integrated transport model for subgrid-scale stresses and dissipation rate for turbulent developing flows. Phys. Fluids 2005, 17, 065106. [Google Scholar] [CrossRef]
  96. Chaouat, B.; Schiestel, R. From single-scale turbulence models to multiple-scale and subgrid-scale models by Fourier transform. Theor. Comput. Fluid Dyn. 2007, 21, 201–229. [Google Scholar] [CrossRef]
  97. Befeno, I.; Schiestel, R. Non-equilibrium mixing of turbulence scales using a continuous hybrid RANS/LES approach: Application to the shearless mixing layer. Flow Turbul. Combust. 2007, 78, 129–151. [Google Scholar] [CrossRef]
  98. Chaouat, B.; Schiestel, R. Progress in subgrid-scale transport modelling for continuous hybrid nonzonal RANS/LES simulations. Int. J. Heat Fluid Flow 2009, 30, 602–616. [Google Scholar] [CrossRef]
  99. Chaouat, B. Subfilter-scale transport model for hybrid RANS/LES simulations applied to a complex bounded flow. J. Turbul. 2010, 11, 1–30. [Google Scholar] [CrossRef]
  100. Chaouat, B. Simulation of turbulent rotating flows using a subfilter scale stress model derived from the partially integrated transport modeling method. Phys. Fluids 2012, 24, 045108. [Google Scholar] [CrossRef]
  101. Chaouat, B.; Schiestel, R. Analytical insights into the partially integrated transport modeling method for hybrid Reynolds averaged Navier-Stokes equations-large eddy simulations of turbulent flows. Phys. Fluids 2012, 24, 085106. [Google Scholar] [CrossRef]
  102. Chaouat, B.; Schiestel, R. Partially integrated transport modeling method for turbulence simulation with variable filters. Phys. Fluids 2013, 25, 125102. [Google Scholar] [CrossRef]
  103. Chaouat, B.; Schiestel, R. Hybrid RANS-LES simulations of the turbulent flow over periodic hills at high Reynolds number using the PITM method. Comput. Fluids 2013, 84, 279–300. [Google Scholar] [CrossRef]
  104. Chaouat, B. Application of the PITM method using inlet synthetic turbulence generation for the simulation of the turbulent flow in a small axisymmetric contraction. Flow Turbul. Combust. 2017, 98, 987–1024. [Google Scholar] [CrossRef]
  105. Menter, F.R.; Kuntz, M.; Bender, R. A scale-adaptive simulation model for turbulent flow predictions. In Proceedings of the 41st AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 6–9 January 2003; AIAA Paper 03-0767. pp. 1–12. [Google Scholar]
  106. Menter, F.R.; Egorov, Y. A scale-adaptive simulation model using two-equation models. In Proceedings of the 43rd AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 10–13 January 2005; AIAA Paper 05-1095. pp. 1–13. [Google Scholar]
  107. Menter, F.R.; Egorov, Y. Revisiting the turbulent scale equation. In IUTAM Symposium on One Hundred Years of Boundary Layer Research, Proceedings of the IUTAM Symposium, Göttingen, Germany, 12–14 August 2004; Meier, G.E.A., Sreenivasan, K.R., Eds.; Springer: Dordrecht, The Netherlands, 2006; pp. 279–290. [Google Scholar]
  108. Menter, F.R.; Egorov, Y. The scale-adaptive simulation method for unsteady turbulent flow prediction: Part 1: Theory and model description. Flow Turbul. Combust. 2010, 78, 113–138. [Google Scholar] [CrossRef]
  109. Menter, F.R.; Egorov, Y. The scale-adaptive simulation method for unsteady turbulent flow prediction: Part 2: Application to complex flows. Flow Turbul. Combust. 2010, 78, 139–165. [Google Scholar]
  110. Jakirlić, S.; Maduta, R. Extending the bounds of “steady” RANS closures: Toward an instability-sensitive Reynolds stress model. Int. J. Heat Fluid Flow 2015, 51, 175–194. [Google Scholar] [CrossRef]
  111. Heinz, S.; Heinz, J.; Brant, J.A. Mass transport in membrane systems: Flow regime identification by Fourier analysis. Fluids 2022, 7, 369. [Google Scholar] [CrossRef]
  112. Honnert, R.; Masson, V.; Couvreux, F. A diagnostic for evaluating the representation of turbulence in atmospheric models at the kilometric scale. J. Atmos. Sci. 2011, 68, 3112–3131. [Google Scholar] [CrossRef]
  113. Honnert, R. Representation of the grey zone of turbulence in the atmospheric boundary layer. Adv. Sci. Res. 2016, 13, 63–67. [Google Scholar] [CrossRef]
  114. Juliano, T.W.; Kosović, B.; Jiménez, P.A.; Eghdami, M.; Haupt, S.E.; Martilli, A. “Gray Zone” simulations using a three-dimensional planetary boundary layer parameterization in the Weather Research and Forecasting Model. Mon. Weather Rev. 2022, 150, 1585–1619. [Google Scholar] [CrossRef]
  115. Honnert, R.; Efstathiou, G.A.; Beare, R.J.; Ito, J.; Lock, A.; Neggers, R.; Plant, R.S.; Shin, H.H.; Tomassini, L.; Zhou, B. The atmospheric boundary layer and the “gray zone” of turbulence: A critical review. J. Geophys. Res. Atmos. 2020, 125, e2019JD030317. [Google Scholar] [CrossRef]
  116. Efstathiou, G.A.; Beare, R.J. Quantifying and improving sub-grid diffusion in the boundary layer grey zone. Q. J. R. Meteorol. Soc. 2015, 141, 3006–3017. [Google Scholar] [CrossRef]
  117. Kealy, J.C.; Efstathiou, G.A.; Beare, R.J. The onset of resolved boundary layer turbulence at grey zone resolutions. Bound.-Layer Meteorol. 2019, 171, 31–52. [Google Scholar] [CrossRef]
  118. Beare, R.J.; Efstathiou, G.A. The role of diffusion in boundary layer turbulence simulation in the grey zone. Bound.-Layer Meteorol. 2025, 191, 30. [Google Scholar] [CrossRef]
  119. De Roode, S.R.; Siebesma, A.P.; Jansson, F.; Janssens, M. Dependency of mesoscale organization on grid anisotropy in large-eddy simulations of convective boundary layers at gray zone resolutions. J. Adv. Model Earth Syst. 2022, 14, e2022MS003095. [Google Scholar] [CrossRef]
  120. Sun, Y.; Yi, L.; Zhong, Z.; Hu, Y.; Ha, Y. Dependence of model convergence on horizontal resolution and convective parameterization in simulations of a tropical cyclone at gray-zone resolutions. J. Geophys. Res. Atmos. 2013, 118, 7715–7732. [Google Scholar] [CrossRef]
  121. Park, H.; Kim, G.; Cha, D.H.; Chang, E.C.; Kim, J.; Park, S.H.; Lee, D.K. Effect of a scale-aware convective parameterization scheme on the simulation of convective cells related heavy rainfall in South Korea. J. Adv. Model Earth Syst. 2022, 14, e2021MS002696. [Google Scholar] [CrossRef]
  122. Tomassini, L.; Willett, M.; Sellar, A.; Lock, A.; Walters, D.; Whitall, M.; Sanchez, C.; Heming, J.; Earnshaw, P.; Rodriguez, J.M.; et al. Confronting the convective gray zone in the global configuration of the Met Office Unified Model. J. Adv. Model Earth Syst. 2023, 15, e2022MS003418. [Google Scholar] [CrossRef]
  123. Honnert, R. Grey zone turbulence in the neutral atmospheric boundary layer. Bound.-Layer Meteorol. 2018, 170, 191–204. [Google Scholar] [CrossRef]
  124. Honnert, R.; Masson, V.; Lac, C.; Nagel, T. A theoretical analysis of mixing length for atmospheric models from micro to large scales. Front. Earth Sci. 2021, 8, 582056. [Google Scholar] [CrossRef]
  125. Sullivan, P.P.; Patton, E.G. The effect of mesh resolution on convective boundary layer statistics and structures generated by large-eddy simulation. J. Atmos. Sci. 2011, 68, 2395–2415. [Google Scholar] [CrossRef]
  126. Mellor, G.L.; Yamada, T. Development of a turbulence closure model for geophysical fluid problems. Rev. Geophys. 1982, 20, 851–875. [Google Scholar] [CrossRef]
  127. Detering, H.W.; Etling, D. Application of the E-ε turbulence model to the atmospheric boundary layer. Bound.-Layer Meteorol. 1985, 33, 113–133. [Google Scholar] [CrossRef]
  128. Duynkerke, P.; Driedonks, A. A model for the turbulent structure of the stratocumulus–topped atmospheric boundary layer. J. Atmos. Sci. 1987, 44, 43–64. [Google Scholar][Green Version]
  129. Duynkerke, P.G. Application of the E–ε turbulence closure model to the neutral and stable atmospheric boundary layer. J. Atmos. Sci. 1988, 45, 865–880. [Google Scholar]
  130. Starchenko, A.V.; Karyakin, A.S. Simulation of turbulent transport during 24-hour evolution of the atmospheric boundary layer. Proceeding of the Seventh International Symposium on Atmospheric and Ocean Optics, Tomsk, Russia, December 2000; International Society for Optics and Photonics: Bellingham, WA, USA, 2000; Volume 4341, pp. 626–633. [Google Scholar]
  131. Freedman, F.R.; Jacobson, M.Z. Modification of the standard ε-equation for the stable ABL through enforced consistency with Monin–Obukhov similarity theory. Bound.-Layer Meteorol. 2003, 106, 383–410. [Google Scholar] [CrossRef]
  132. Castelli, S.T.; Ferrero, E.; Anfossi, D.; Ohba, R. Turbulence closure models and their application in RAMS. Environ. Fluid Mech. 2005, 5, 169–192. [Google Scholar] [CrossRef]
  133. Zhang, C.; Wang, Y.; Xue, M. Evaluation of an E–ε and three other boundary layer parameterization schemes in the WRF model over the Southeast Pacific and the Southern Great Plains. Mon. Weather Rev. 2020, 148, 1121–1145. [Google Scholar] [CrossRef]
  134. Tran, V.; Ng, E.Y.K.; Skote, M. CFD simulation of dense gas dispersion in neutral atmospheric boundary layer with OpenFOAM. Meteorol. Atmos. Phys. 2020, 132, 273–285. [Google Scholar] [CrossRef]
  135. Mason, P.J.; Thomson, D.J. Stochastic backscatter in large-eddy simulations of boundary layers. J. Fluid Mech. 1992, 242, 51–78. [Google Scholar] [CrossRef]
  136. Senocak, I.; Ackerman, A.S.; Kirkpatrick, M.P.; Stevens, D.E.; Mansour, N.N. Study of near-surface models for large-eddy simulations of a neutrally stratified atmospheric boundary layer. Bound.-Layer Meteorol. 2007, 124, 405–424. [Google Scholar] [CrossRef]
  137. McWilliams, J.C.; Meneveau, C.; Patton, E.G.; Sullivan, P.P. Stable boundary layers and subfilter-scale motions. Atmosphere 2023, 14, 1107. [Google Scholar] [CrossRef]
  138. Heinz, S.; Mokhtarpoor, R.; Stoellinger, M.K. Theory-based Reynolds-averaged Navier-Stokes equations with large eddy simulation capability for separated turbulent flow simulations. Phys. Fluids 2020, 32, 065102. [Google Scholar] [CrossRef]
  139. Fagbade, A.; Heinz, S. Continuous eddy simulation (CES) of transonic shock-induced flow separation. Appl. Sci. 2024, 14, 2705. [Google Scholar] [CrossRef]
  140. Fagbade, A.; Heinz, S. Continuous eddy simulation vs. resolution-imposing simulation methods for turbulent flows. Fluids 2024, 9, 22. [Google Scholar] [CrossRef]
  141. Dong, T.; Minelli, G.; Wang, J.; Liang, X.; Krajnović, S. Numerical investigation of a high-speed train underbody flows: Studying flow structures through large-eddy simulation and assessment of steady and unsteady Reynolds-averaged Navier–Stokes and improved delayed detached eddy simulation performance. Phys. Fluids 2022, 34, 015126. [Google Scholar] [CrossRef]
  142. Temel, O.; van Beeck, J. Two-equation eddy viscosity models based on the Monin–Obukhov similarity theory. Appl. Math. Model. 2017, 42, 1–16. [Google Scholar] [CrossRef]
  143. Argyropoulos, C.D.; Markatos, N.C. Recent advances on the numerical modelling of turbulent flows. Appl. Math. Model. 2015, 39, 693–732. [Google Scholar] [CrossRef]
  144. Heinz, S. Theory-based mesoscale to microscale coupling for wind energy applications. Appl. Math. Model. 2021, 98, 563–575. [Google Scholar] [CrossRef]
  145. Heinz, S. From two-equation turbulence models to minimal error resolving simulation methods for complex turbulent flows. Fluids 2022, 7, 368. [Google Scholar] [CrossRef]
  146. Rotta, J.C. Turbulente Strömumgen; BG Teubner: Stuttgart, Germany, 1972. [Google Scholar]
  147. Speziale, C.G. On nonlinear kL and k-ϵ models of turbulence. J. Fluid Mech. 1987, 178, 459–475. [Google Scholar] [CrossRef]
  148. Smith, B. A near wall model for the k-L two equation turbulence model. In Proceedings of the Fluid Dynamics Conference, Colorado Springs, CO, USA, 20–23 June 1994; AIAA Paper 94-2386. pp. 1–9. [Google Scholar]
  149. Goldberg, U.; Chakravarthy, S. A k-L turbulence closure for wall-bounded flows. In Proceedings of the 35th AIAA Fluid Dynamics Conference and Exhibit, Toronto, ON, Canada, 6–9 June 2005; AIAA Paper 05-4638. pp. 1–11. [Google Scholar]
  150. Rotta, J.C. Über eine Methode zur Berechnung Turbulenter Scherströmungen. Aerodyn. Vers. 1968, 69, A14 . [Google Scholar]
  151. Rodi, W. Turbulence modelling for boundary-layer calculations. In Proceedings of the IUTAM Symposium, Göttingen, Germany, 12–14 August 2004; Meier, G., Sreenivasan, K., Eds.; Springer: Berlin/Heidelberg, Germany, 2006; pp. 247–256. [Google Scholar]
  152. Heinz, S. Statistical Mechanics of Turbulent Flows; Springer: Berlin/Heidelberg, Germany, 2003. [Google Scholar]
  153. Heinz, S. Nonlinear Lagrangian equations for turbulent motion and buoyancy in inhomogeneous flows. Phys. Fluids 1997, 9, 703–716. [Google Scholar] [CrossRef][Green Version]
  154. Heinz, S. Connections between Lagrangian stochastic models and the closure theory of turbulence for stratified flows. Int. J. Heat Fluid Flow 1998, 19, 193–200. [Google Scholar] [CrossRef]
  155. Heinz, S.; van Dop, H. Buoyant plume rise described by a Lagrangian turbulence model. Atmos. Environ. 1999, 33, 2031–2043. [Google Scholar] [CrossRef]
  156. Baumert, H.; Peters, H. Second-moment closures and length scales for weakly stratified turbulent shear flows. J. Geophys. Res. Oceans 2000, 105, 6453–6468. [Google Scholar] [CrossRef]
  157. Sogachev, A. A note on two-equation closure modelling of canopy flow. Bound.-Layer Meteorol. 2009, 130, 423–435. [Google Scholar] [CrossRef]
  158. Stull, R.B. An Introduction to Boundary Layer Meteorology; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1988. [Google Scholar]
  159. Heinz, S. Strategy analysis of seamlessly resolving turbulent flow simulations. Aerospace 2025, 12, 597. [Google Scholar] [CrossRef]
  160. Burchard, H.; Petersen, O.; Rippeth, T.P. Comparing the performance of the Mellor-Yamada and the κ-ε two-equation turbulence models. J. Geophys. Res. Oceans 1998, 103, 10543–10554. [Google Scholar] [CrossRef]
  161. Burchard, H.; Petersen, O. Models of turbulence in the marine environment—A comparative study of two-equation turbulence models. J. Mar. Syst. 1999, 21, 29–53. [Google Scholar] [CrossRef]
  162. Heinz, S.; Fagbade, A. Evaluation metrics for partially and fully resolving simulations methods for turbulent flows. Int. J. Heat Fluid Flow 2025, 115, 109867. [Google Scholar] [CrossRef]
  163. Heinz, S. The potential of machine learning methods for separated turbulent flow simulations: Classical versus dynamic methods. Fluids 2024, 9, 278. [Google Scholar] [CrossRef]
  164. Heinz, S. Physically consistent resolving simulations of turbulent flows. Entropy 2024, 26, 1044. [Google Scholar] [CrossRef]
  165. Heinz, S. A mathematical solution to the Computational Fluid Dynamics (CFD) dilemma. Mathematics 2023, 11, 3199. [Google Scholar] [CrossRef]
  166. Heinz, S. Minimal error partially resolving simulation methods for turbulent flows: A dynamic machine learning approach. Phys. Fluids 2022, 34, 051705. [Google Scholar] [CrossRef]
  167. Heinz, S. Remarks on energy partitioning control in the PITM hybrid RANS/LES method for the simulation of turbulent flows. Flow, Turb. Combust. 2022, 108, 927–933. [Google Scholar] [CrossRef]
  168. Heinz, S. The continuous eddy simulation capability of velocity and scalar probability density function equations for turbulent flows. Phys. Fluids 2021, 33, 025107. [Google Scholar] [CrossRef]
  169. Heinz, S.; Peinke, J.; Stoevesandt, B. Cutting-edge turbulence simulation methods for wind energy and aerospace problems. Fluids 2021, 6, 288. [Google Scholar] [CrossRef]
  170. Heinz, S. The large eddy simulation capability of Reynolds-averaged Navier-Stokes equations: Analytical results. Phys. Fluids 2019, 31, 021702. [Google Scholar] [CrossRef]
  171. Rapp, C.; Manhart, M. Flow over periodic hills—An experimental study. Exp. Fluids 2011, 51, 247–269. [Google Scholar] [CrossRef]
  172. Kähler, C.J.; Scharnowski, S.; Cierpka, C. Highly resolved experimental results of the separated flow in a channel with streamwise periodic constrictions. J. Fluid Mech. 2016, 796, 257–284. [Google Scholar] [CrossRef]
  173. Saini, R.; Karimi, N.; Duan, L.; Sadiki, A.; Mehdizadeh, A. Effects of near wall modeling in the improved-delayed-detached-eddy-simulation (IDDES) methodology. Entropy 2018, 20, 771. [Google Scholar] [CrossRef]
  174. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 1598–1605. [Google Scholar] [CrossRef]
  175. Spalart, P.; Allmaras, S. A one-equation turbulence model for aerodynamic flows. La Rech. Aérosp. 1994, 1, 5–21. [Google Scholar]
  176. Warhaft, Z. Passive scalars in turbulent flows. Annual Rev. Fluid Mech. 2000, 32, 203–240. [Google Scholar] [CrossRef]
  177. Das, S.K.; Durbin, P.A. A Lagrangian stochastic model for dispersion in stratified turbulence. Phys. Fluids 2005, 17, 025109. [Google Scholar] [CrossRef]
Figure 1. An illustration of the Terra Incognita challenge [30]. Simulations of the ABL on (a) very coarse, (b) medium, and (c) very fine grids are considered. ABL turbulent eddies and clouds above them are illustrated.
Figure 1. An illustration of the Terra Incognita challenge [30]. Simulations of the ABL on (a) very coarse, (b) medium, and (c) very fine grids are considered. ABL turbulent eddies and clouds above them are illustrated.
Fluids 10 00301 g001
Figure 2. Velocity streamlines seen in periodic hill flows: results obtained by continuous eddy simulation at R e = 37 , 000 . Reprinted with permission from Ref. [138]. Copyright 2020 AIP Publishing.
Figure 2. Velocity streamlines seen in periodic hill flows: results obtained by continuous eddy simulation at R e = 37 , 000 . Reprinted with permission from Ref. [138]. Copyright 2020 AIP Publishing.
Fluids 10 00301 g002
Figure 3. Periodic hill flow at R e = 37 K [173]: response of the f ˜ d function to coarse grid resolution at x / H = 4 ; left column: k ω -SST IDDES, right column: Spalart–Allmaras (S-A) IDDES.
Figure 3. Periodic hill flow at R e = 37 K [173]: response of the f ˜ d function to coarse grid resolution at x / H = 4 ; left column: k ω -SST IDDES, right column: Spalart–Allmaras (S-A) IDDES.
Fluids 10 00301 g003
Figure 4. Periodic hill flow at R e = 37 K [173]: variation in the ratio of the modeled to total turbulent kinetic energy at four different streamwise locations under coarse grid resolution; SRR = scale-resolving region; left column (a): k ω -SST IDDES, right column (b): Spalart–Allmaras (S-A) IDDES.
Figure 4. Periodic hill flow at R e = 37 K [173]: variation in the ratio of the modeled to total turbulent kinetic energy at four different streamwise locations under coarse grid resolution; SRR = scale-resolving region; left column (a): k ω -SST IDDES, right column (b): Spalart–Allmaras (S-A) IDDES.
Fluids 10 00301 g004
Figure 5. Periodic hill flow at R e = 37 K [173]: profiles of the mean streamwise velocity U x , normal streamwise Reynolds stress u u , and shear stress u v normalized by the bulk velocity U b at x / H = 4 ; red = k ω -SST IDDES, blue = Spalart–Allmaras (S-A) IDDES, solid (dashed) lines refer to coarse (fine) grids.
Figure 5. Periodic hill flow at R e = 37 K [173]: profiles of the mean streamwise velocity U x , normal streamwise Reynolds stress u u , and shear stress u v normalized by the bulk velocity U b at x / H = 4 ; red = k ω -SST IDDES, blue = Spalart–Allmaras (S-A) IDDES, solid (dashed) lines refer to coarse (fine) grids.
Fluids 10 00301 g005
Figure 6. Left: CES model comparisons at R e = 37 K ( G 500 ): k + is shown at x / h = 4 . Right: CES-KOKU results at R e = 37 K on G 500 , G 250 , G 120 . The dotted (dashed) line shows k + ( L + 2 / 3 ). The insets show profiles close to the lower wall in y + scaling. Reprinted with permission from Ref. [138]. Copyright 2020 AIP Publishing.
Figure 6. Left: CES model comparisons at R e = 37 K ( G 500 ): k + is shown at x / h = 4 . Right: CES-KOKU results at R e = 37 K on G 500 , G 250 , G 120 . The dotted (dashed) line shows k + ( L + 2 / 3 ). The insets show profiles close to the lower wall in y + scaling. Reprinted with permission from Ref. [138]. Copyright 2020 AIP Publishing.
Fluids 10 00301 g006
Figure 7. CES-KOKU results at R e = 37 K, x / h = 4 , for the streamwise velocity U / U b , normal streamwise Reynolds stress u u / U b 2 , and shear stress u v / U b 2 on G 500 , G 250 , G 120 . Reprinted with permission from Ref. [138]. Copyright 2020 AIP Publishing.
Figure 7. CES-KOKU results at R e = 37 K, x / h = 4 , for the streamwise velocity U / U b , normal streamwise Reynolds stress u u / U b 2 , and shear stress u v / U b 2 on G 500 , G 250 , G 120 . Reprinted with permission from Ref. [138]. Copyright 2020 AIP Publishing.
Fluids 10 00301 g007
Table 1. Overview of terminology applied in this paper.
Table 1. Overview of terminology applied in this paper.
Resolving vs. nonresolving simulations: Resolving simulations performed on sufficiently fine grids enable the calculation of most of the spectrum of turbulent motions. They include turbulent fluctuations produced by the simulation. Nonresolving simulations do not include fluctuations; the turbulence is reflected by a turbulence model.
Resolved vs. modeled motion (or flow): Resolved motion stands for turbulent fluctuations produced by a resolving simulation. Modeled motion stands for turbulence described by an explicit model for turbulence.
RANS vs. LES [averaged vs. filtered equations]: RANS and LES equations can be derived using, e.g., ensemble averaging or filtering in space. With respect to simulations, there is no difference if the equation structure and parametrization of turbulence length scales are the same. The typical RANS-LES difference is the inclusion of a relatively large (small) turbulence length scale or viscosity, respectively. RANS (LES) are usually performed on relatively coarse (fine) grids focusing on flow modeling (flow resolution). Typically applied RANS equations can be used for resolving LES if the turbulence length scale and grid are sufficiently small. Used on appropriate grids, typically applied RANS equations can produce fluctuations; such RANS are called unsteady RANS (URANS).
Hybrid RANS-LES vs. RANS or LES: The term hybrid RANS-LES will be used for methods aiming at the ability to be seamlessly applicable to nonresolving and resolving simulations (in contrast to RANS or LES aiming at one of these cases). This view of hybrid RANS-LES includes, e.g., DES, which switches RANS and LES length scales to transition from one regime to the other (to enable a transition between modeled and resolved flow regions). Depending on the set-up, this view of hybrid RANS-LES can include WMLES, which applies RANS elements in some flow regions. This view of hybrid RANS-LES does not include typical atmospheric LES applied in conjunction with wall-modeling. Such simulations will be considered here simply as LES because the focus is still on flow resolution, i.e., not an a transition between resolution regimes.
Mesoscale vs. microscale simulations: Atmospheric mesoscale (microscale) simulations are performed with a resolution of a few km to hundreds of km (a resolution of meters to a few km). For simplicity, mesoscale and microscale simulations will be considered here as using incompletely resolving RANS and resolving LES. Specifically, such mesoscale simulations typically represent URANS, which includes the calculation of fluctuations.
Gray-zone vs. Terra Incognita: The term gray-zone refers to simulations on intermediate grids where the basic assumptions of RANS or LES are not satisfied. This may refer to coarse LES, where the resolved fluctuations do not need to represent realistic turbulence. This may refer to URANS where the appearance of fluctuations can lead to essentially misaligned production–dissipation mechanisms of equations. The term Terra Incognita is usually applied to refer to the gray-zone problem in atmospheric simulations. Dynamic LES is no option to address gray-zone problems. Dynamic LES is still resolving LES aiming at a better performance based on dynamic coefficient adjustments. The problem of under-resolution on coarse grids cannot properly be addressed by dynamic LES.
Scale-aware vs. resolution-aware parametrizations: Scale-aware parametrizations usually include reference to the filter width ( Δ ) in equations (like usual LES parametrizations). Under many conditions, such a reference is an insufficient reflection of the real flow resolution. Resolution-aware parametrizations (like CES) include explicit reference to the actual flow resolution in equations (by inclusion of k + = k / k t o t or L + = L / L t o t ; see below).
DES vs. CES simulations: Both address the gray-zone issue, but very differently. DES switches RANS and LES length scales, hoping for a seamless transition between modeled and resolved flow regions. Rather often, the latter transition happens with significant reluctance, which can lead to a significant mismatch of modeled and resolved flow. These problems are avoided in CES because of the direct interaction of modeled and resolved flow.
Table 3. Overview of significant DES and CES differences.
Table 3. Overview of significant DES and CES differences.
Detached Eddy Simulation (DES)Continuous Eddy Simulation (CES)
Core ideaswitch of RANS, LES length scalesresolution-aware modeling (model knows and appropriately responds to actual flow resolution)
Advantagesimple idea; simple to implement (simpler than CES implementation)exact math-based; direct actual resolution measurement; no grid influence; no DES problems
DisadvantageAlways: no implied resolution switch, resolved and modeled motion mismatch, many different versions of fixes, significant grid dependence, imbalanced predictions, significant cost to fix problemsOnly once: modeling approach requires calculation of actual flow resolution indicators on the fly, implementation of scale transport equation in general required
Functionalityspatial switch between RANS-type and LES-type flow regions (no CES transitionality)full hybrid model transitionality from modeled to resolved simulations
Table 4. Errors of reattachment point calculations by CES [138] and IDDES [173] methods. The experimentally determined reattachment point is x / h = 3.76 [171].
Table 4. Errors of reattachment point calculations by CES [138] and IDDES [173] methods. The experimentally determined reattachment point is x / h = 3.76 [171].
MethodReattachment ( x / h ) Error (%)
CES ( G 500 )3.780.5
CES ( G 120 )3.7 1.6
S-A IDDES (coarse 0.3M grid)3.95785.3
S-A IDDES (fine 0.9M grid)4.292214.2
k ω -SST IDDES (coarse 0.3M grid)3.95785.3
k ω -SST (fine 0.9M grid)4.707825.2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Heinz, S. The Atmospheric Gray-Zone (a.k.a. Terra Incognita) Problem: A Strategy Analysis from an Engineering Viewpoint. Fluids 2025, 10, 301. https://doi.org/10.3390/fluids10110301

AMA Style

Heinz S. The Atmospheric Gray-Zone (a.k.a. Terra Incognita) Problem: A Strategy Analysis from an Engineering Viewpoint. Fluids. 2025; 10(11):301. https://doi.org/10.3390/fluids10110301

Chicago/Turabian Style

Heinz, Stefan. 2025. "The Atmospheric Gray-Zone (a.k.a. Terra Incognita) Problem: A Strategy Analysis from an Engineering Viewpoint" Fluids 10, no. 11: 301. https://doi.org/10.3390/fluids10110301

APA Style

Heinz, S. (2025). The Atmospheric Gray-Zone (a.k.a. Terra Incognita) Problem: A Strategy Analysis from an Engineering Viewpoint. Fluids, 10(11), 301. https://doi.org/10.3390/fluids10110301

Article Metrics

Back to TopTop