In this work, a generalized wall-modeled large eddy simulation model (GWMLES) is presented. An extended formulation of the classical WMLES approach is proposed that also enables the modeling of the entire log-layer by using a Reynolds-averaged Navier–Stokes (RANS) model. GWMLES is validated against direct numerical simulations, large eddy simulations (LES), WMLES, hybrid RANS/LES, unsteady RANS (URANS), and experimental data of test cases featuring industrial flows. It is demonstrated that GWMLES does not share the main shortcoming of current WMLES models. When the entire log-layer is solved with a RANS model, GWMLES gives a level of accuracy similar to recent LES results, as well as computational cost savings that are proportional to the Reynolds number in wall-bounded flows. The model shows superior performance than URANS even when the resolved portion of the energy spectrum is reduced. Motivated by the different time scales of the flow and RANS variables, it requires approximately 30% lower computational costs than the detached eddy simulation family models in the turbulent flows considered. These features represent significant advancements in the simulation of wall-bounded flows at high Reynolds numbers, particularly in industrial applications.

Turbulent flows are present in multiple industrial applications such as aircraft and vehicle aerodynamics, turbomachinery, medical devices, and weather prediction. The understanding of turbulent flow is important, since turbulence produces effects that include the increase in friction losses and heat transfer, the delay of separation in flows exposed to adverse pressure gradients, and the noise generation. The complete flow description is obtained with the direct numerical simulations (DNS), in which the entire turbulence spectrum is resolved. Estimates derived from Kolmogorov's hypotheses show that, for a turbulent flow above a flat plate of length L, the number of grid points is N DNS R e L 37 / 14.1 

The computational cost of DNS is reduced with the large eddy simulation (LES) models, which solve part of the turbulence spectrum and model the remainder. The dynamics of the large-scale motions, which are affected by the flow geometry and are not universal, are computed explicitly. The small scales have a universal character and follow the Kolmogorov hypotheses, and they are represented by simple models. As a general rule, LES simulations require a minimum of 80% of the turbulent kinetic energy to be resolved,2 which is restrictive in wall-bounded flows at high Reynolds numbers; for example, in the turbulent flow above a flat plate, the number of grid points is N LES R e L 13 / 7.1 The scaling problem near the wall is reduced in the wall-modeled LES (WMLES) models, where a mixing-length model or a wall function is used up to the inner part of the logarithmic layer. For a WMLES simulation of a turbulent flow above a flat plate, the number of grid points is N WMLES R e L.1 The accuracy of current WMLES models is limited by the log-layer mismatch (LLM), which is the deviation of the velocity from the log-law, especially where the model switches from the wall function (or the mixing-length model) to LES.3 Multiple solutions have been proposed, including the addition of forcing terms in the momentum equations,4 although the constants of the forcing terms are not universal. This is partially solved in dynamic approaches that are based on control systems,5 although at the expense of the increase in the complexity and the numerical stability. The error of the LLM is proportional to the dimensionless wall distance y ̃ + of the first near-wall cell, and mesh refinement is not always possible due to the requirement of using cells of low aspect ratio for resolved turbulence. If a mixing-length model is used near the wall,6 the use of high aspect ratio cells is limited due to the reduced width in which the Reynolds-averaged Navier–Stokes (RANS) part is applicable.

In the Reynolds-averaged Navier–Stokes (RANS) models, the computational cost does not depend on the Reynolds number. The entire energy spectrum is modeled, and since the variables of the equations are mean quantities, temporal resolution is not required in the simulations. RANS models have shown great accuracy in wall-bounded attached flows, where the law of the wall establishes a reference for the derivation of the model constants. Two-equation models are currently the reference for industrial simulations, since they provide the optimum balance between computational cost, stability, and accuracy. However, mean quantities are not well predicted in free shear flows7 and separated flows,8 which constitute an important part of industrial applications. Furthermore, due to their time-averaged formulation, RANS models are usually not suitable for predicting time-dependent quantities, for example, in acoustics simulations,9 fluid–structure interactions,10 and multiphase flows.11 There are particular cases—namely, massively separated flows—where unsteady RANS simulations (URANS) provide unsteady solutions, and specifically part of the turbulence spectrum, although the accuracy is limited in these situations.12 

In wall-bounded flows at high Reynolds numbers, the computational cost of LES is unaffordable for practical engineering applications. On the contrary, the cost in RANS is several orders of magnitude lower, and the results given by these models for boundary layers without detachment are often acceptable. In free shear flows, much less computational power is required to perform LES simulations. A compromise solution in terms of accuracy and computational cost is given by the hybrid RANS/LES models, where RANS is used in the boundary layers, and LES in the rest of the domain. Both WMLES with the mixing-length approach and hybrid RANS/LES models share the idea of blending RANS and LES models, which leads to the use of both names interchangeably by some authors. In the present article, WMLES is assumed to be a model where the blending takes place at a y ̃ + value in the inner part of the log-layer (i.e., 10 < y ̃ + < 50). On the contrary, hybrid RANS/LES models are considered those where transition takes place at higher values of y ̃ + up to the edge of the boundary layer.

A reference hybrid RANS/LES model is the detached eddy simulation (DES) model. The two-equation DES model introduces the grid spacing Δ = max ( Δ x , Δ y , Δ z ) in the dissipation term of the turbulent kinetic energy equation,13 where Δi refers to the grid spacing in the i direction. The model switches to LES if the turbulence length scale l t > C DES Δ for a constant C DES 1; therefore, it reduces to a RANS model in coarse meshes, and to LES in fine meshes. However, for intermediate conditions, the transition from RANS and LES can occur without control. If the original mesh satisfies l t < C DES Δ, then the model stays in RANS [see Fig. 1(a)]. If the mesh is refined so that l t > C DES Δ in some regions of the boundary layer, then the model switches to LES. Since lt reaches its maximum value in the middle of the boundary layer,14 a real situation can be the one shown in Fig. 1(b), where δ is the thickness of the boundary layer. This is a stable flow because there is no phenomenon that triggers an instability, so no vortex structures are generated, leading to early separation, especially in flows exposed to an adverse pressure gradient. This numerical behavior is called grid-induced separation (GIS).15 If the mesh is further refined, the switching of the model from RANS to LES formulations occurs closer to the wall; that is, the width of the LES zone is extended, while there is no certainty that the mesh is fine enough for a LES simulation. Thus, in DES, mesh refinement does not guarantee a more accurate prediction.

FIG. 1.

Schematic of the grid-induced separation in DES. (a) Coarse mesh. (b) Fine mesh.

FIG. 1.

Schematic of the grid-induced separation in DES. (a) Coarse mesh. (b) Fine mesh.

Close modal

An improvement of DES called delayed-DES (DDES) was published later15 that introduces a shielding function in the formulation. While in DES the effect of GIS is noticeable from Δ δ, in DDES, it is reduced to Δ 0.25 δ.14 However, the impact is much more drastic than in DES, since for Δ < 0.25 δ, small changes in the mesh size produce significant variations in the solution. As a consequence, DES/DDES can provide less accurate solutions than RANS.16 Another important aspect of DES/DDES is that the transition from RANS to LES takes place in a distance whose size is not negligible; in particular, in DDES, it has been shown that it is around 0.3 δ in a flat plate.17 This is called gray area in the literature, since the turbulence model is not well defined there. Gray areas delay the generation of turbulent structures and generate spurious buffer layers18 and the LLM.19 

Variants of the DDES model have been proposed, based on the use of the blending functions of the k ω SST model,20 which reduce the GIS down to Δ 0.1 δ, but the gray areas are not eliminated. More elaborated formulations involve the use of additional complex functions that depend on flow variables that act as sensors.21 The model cited solves the problem of the gray areas, although the GIS is not addressed. Moreover, it includes multiple additional adjusting constants, and it has not been verified that the values proposed are universal. The improved-DDES (IDDES) model is a modification of DES developed with the capability to run in WMLES mode,6 but with limited shielding against GIS. For meshes that are not suitable for WMLES, the accuracy of DDES and IDDES is similar,22 and recent studies have shown that IDDES does not solve completely the LLM.23 

Another weakness shared by all DES family is the reduced accuracy of the LES part. By assuming an equilibrium of the source terms of the RANS turbulence variables, it can be demonstrated that the LES formulation corresponds to a Smagorinsky model; specifically, for the original constant C DES = 0.61 (Ref. 13) used in DDES with the k ω SST model as the RANS model,24 the corresponding Smagorinsky constant outside the boundary layers is C s 0.175. This value is too dissipative for free shear flows, for which C s 0.1 has been shown to be optimal.25 DDES also inherits limitations from the Smagorinsky model, namely, the need to adjust the model constant based on the specific flow being simulated. An additional deficiency is the persistent presence of non-zero turbulent viscosity in standard laminar regions, including laminar shear layers, where it should ideally vanish.

Similar to the DES family, the scale-adaptive simulation (SAS) model is a formulation that takes RANS equations as reference, and a second equation for the turbulence scale is reformulated.26 This results in the inclusion of the second velocity space derivative in the RANS equations, so that the scale equation is able to adjust to resolved scales in the flow. SAS has been shown to provide performances superior to that of RANS (URANS) models in flows with strong instabilities.27,28 However, for a given mesh suitable for LES far from walls, when the flow instability is not enough to enable the LES model, SAS solves turbulent structures of a size similar to those of URANS,29 and the model can be less accurate than URANS.30 Thus, in these cases, gray areas are produced, since the model does not solve the equations neither in URANS nor in LES mode. While in the DES family the location, width, and conditions of the gray areas are reasonably known, in SAS that knowledge does not exist. Other hybrid RANS/LES models available in the literature solve the partially averaged Navier–Stokes (PANS) equations.31 The difference between RANS and PANS is that the latter contains modified parameters in the turbulence dissipation and diffusion terms, so that, under some restrictions, the required level of modeled-to-resolved turbulent kinetic energy and dissipation rate can be set. PANS does not include additional shielding functions, so the presence of gray areas, the LLM, and the GIS are potential problems. The exact transition point from RANS to LES can be defined with the Zonal RANS/LES methods, where RANS and LES are solved in different parts of the domain, and a synthetic turbulence generator transfers modeled RANS turbulence to resolved LES turbulence.32 However, there are not many industrial applications in which the location of the RANS/LES interface can be accurately defined.

The stress-blended eddy simulation (SBES) approach has been developed by Ansys to overcome the deficiencies of the original DES model and its variants.14 The basis is the blending of the RANS and LES models at the stress level, which is equivalent to the blending of the turbulent viscosities if both models use the Boussinesq turbulent viscosity assumption. All RANS and LES models can be combined, which is especially relevant in the LES case, since more consistent models than the Smagorinsky model can be used. In addition, motivated by the higher time scale of the RANS turbulence quantities in comparison with the rest flow variables, in SBES the equations of RANS variables can be solved every 5–10 flow time steps. This benefit is obtained due to the decoupling of the RANS and LES turbulent viscosities and leads to a computational cost similar to the WMLES/LES models, which is lower than the one of DES/DDES/IDDES that solve the equations of the RANS variables at all time steps. Ansys claims that the blending function is defined, such that the RANS model covers the entire boundary layer, with no GIS problem, and with gray areas of negligible size. It has been verified that the model switches to WMLES mode depending on the mesh size and the turbulence content of the flow. However, the details of the formulation are not published, and it is available only in Ansys Fluent/CFX. SBES has shown similar or greater accuracy than IDDES/DDES in a mixing layer,14 automotive applications,33 and multiphase flows.34 

Hybrid RANS/LES models can also be applied to transitional flows with transition RANS models, namely, the γ R e θ t,35 which has been validated in turbomachinery applications.36 The model has shown to be capable of predicting some of the physics of the laminar transition in further studies on wing profiles37 and wind turbines.38,39 However, in some cases, discrepancies with experimental data have been encountered near the separation regions. Transitional flows have been studied with LES models, although significant errors have been observed in the early stages of separation bubbles.40 This is because the residual stress models do not detect the entire laminar regions in medium and coarse meshes, so that the resolution needs to be increased to levels that are close to DNS simulations.41 Alternatives have been proposed using sensors to detect the transition point,42 although this methodology implies modeling and the addition of uncertainties in general cases. Since the LES approach combines the high CPU cost and the modeling requirement, an interesting approach is to use a transition RANS model as a basis for a hybrid RANS/LES model in these cases. However, as noted in Ref. 43, the number of publications on this topic is reduced.

In summary, in the industrial context, DNS simulations are not an option in the near future, and LES is limited to free shear flows due to the high computing power required to perform these simulations. Given its limited accuracy, RANS is an option just for preliminary simulations and/or non-complex turbulent flows. Thus, WMLES simulations for moderate Reynolds numbers and hybrid RANS/LES models for high Reynolds numbers are stated as reference for industrial flows in the near future.44,45 In the hybrid RANS/LES case, provided there is no GIS, the requirement of solving the outer layer of the boundary layer in RANS can be dropped, since it is not universal and the mesh requirements are of the same order as in the freestream region. Thus, both WMLES and hybrid RANS/LES can be unified to a model that solves in RANS the viscous sublayer and/or the log-layer, depending on the simulation requirements.

The main goal of this work is to develop a generalized WMLES formulation (GWMLES) that extends the classical WMLES approach. The model is divided into two separated formulations: GWMLESsub, where turbulence is modeled up to the inner part of the log-layer, and GWMLES log, where the viscous sublayer and the entire log-layer are modeled. The GWMLESsub aims to improve current WMLES models, while the purpose of the GWMLES log is to reduce the computational cost of the DES family, as well as to eliminate their shortcomings that lead to solution deterioration (see Fig. 2). The GWMLES model is based on the SBES approach of blending the turbulent viscosities, but with two key differences based on the information provided by Ansys: first, that in GWMLES log the RANS model does not cover the outer layer of the boundary layer, which is not universal, and second, that the selection between GWMLES log and GWMLESsub is done by the user and not depending on the mesh size. The remaining requirements for the blending function are shared with SBES to overcome the deficiencies of the DES family, namely, the dissipative LES part, GIS, and the gray areas. This paper is organized as follows. First, the methodology is presented in Sec. II, including the details of the model. Then, in Sec. III, the results given by the model are compared with DDES, IDDES, SBES, and URANS, as well as with experimental data and DNS/LES/WMLES results available in the literature; in particular, in four test cases that cover three key industrial applications: a free shear flow, two turbulent wall-bounded flows with both GWMLES submodels, and a laminar–turbulent transitional flow. Finally, the conclusions and future work are drawn in Sec. IV.

FIG. 2.

Overview of the main approaches for simulating turbulent flows. The differences between the models are not in real scale, since they are Reynolds number-dependent.

FIG. 2.

Overview of the main approaches for simulating turbulent flows. The differences between the models are not in real scale, since they are Reynolds number-dependent.

Close modal

The model equations are listed in Sec. II A, which are closed with the expression of the blending function of the model that is developed in Sec. II A 1. The numerical implementation is discussed in Sec. II B, including, in Sec. II B 3, the definition of quality estimators used in numerical simulations for model validation.

The compressible Navier–Stokes equations are2 
(1)
where ρ is the density, v = ( u , v , w ) the velocity vector, t the time, p the pressure, e the specific internal energy, and T the temperature. Equations (1)–(3) contain the following assumptions, terms, and parameters:
  • The viscous stress tensor τ uses the Boussinesq turbulent viscosity assumption2 
    (4)

    where μt is the turbulent viscosity, δij is the Kronecker delta, S i j = ( v i / x j + v j / x i ) / 2 are the components of the strain rate tensor, and k is the turbulent kinetic energy.

  • The turbulent viscosity μt blends the RANS and LES formulations
    (5)
  • Unless otherwise specified, μ t RANS is given by the k ω SST model,24 and μ t LES is given by the WALE model46 with a constant C w = 0.325. The blending function f GWMLES is presented in Sec. II A 1. Since the RANS and LES formulations are used in different domain zones, it is considered that the flow variables are Reynolds-averaged in the RANS regions ( f GWMLES = 1) and implicitly filtered in the LES regions ( f GWMLES = 0 ).

  • The production terms of the RANS equations—which are proportional to μ t 2—use the turbulent viscosity μ t RANS in the entire domain [and not μt, see Eq. (5)], even in the zones where f GWMLES = 0. This ensures that f GWMLES does not depend explicitly on the mesh size, since, as shown hereafter, f GWMLES is a function of the RANS turbulence variables.

  • The specific internal energy e is given by e = c v ( T T 0 ), where cv is the gas specific heat capacity at constant volume, and T 0 = 298.15 K is the reference temperature.

  • Fourier's law is used for conductive heat flux in Eq. (3), with κ eff = κ + μ t c p / P r t referring to the effective thermal conductivity, with the turbulent Prandtl number P r t = 0.85.47 

  • The viscous dissipation function ϕ v = · ( τ · v ) v · ( · τ ).

1. GWMLES blending function

The expression for f GWMLES is developed to provide two different formulations: GWMLES log, where f GWMLES 1 up to the outermost part of the log-layer and f GWMLES 0 elsewhere, and GWMLESsub, where f GWMLES 1 up to the inner part of the log-layer and f GWMLES 0 elsewhere. The starting point is the F2 function of the k ω SST model
(6)
that is, F 2 1 if arg1 or arg 2 1, while F 2 0 if arg1 and arg 2 1. In arg1, β * = 0.09 is a model constant, ω is the turbulent specific dissipation rate, and y ̃ is the wall distance, while in arg2, the kinematic viscosity ν = μ / ρ. The expressions of arg1 and arg2 are based on perturbation analyses of the boundary layer (see Refs. 24 for an overview and 48 for all details):
  • arg1: the equations of k and ω can be simplified in the log-layer by making the same assumptions done in the derivation of the law of the wall. As a result, the turbulence length scale l t = k / ( β * ω ) 2.5 y ̃ in the log-layer. This means that tanh [ ( k / ( β * ω y ̃ ) ) 2 ] tanh ( 2.5 2 ) 1 holds in the log-layer for the k ω SST model. In Eq. (6), the coefficient equal to 2 acts as safeguard to ensure that F 2 1 in the log-layer of all flows.

  • arg2: the equations of k and ω in the viscous sublayer are obtained as a limit y ̃ 0 of the log-layer equations. This expansion yields ω = ω sub 85 ν / y ̃ 2 for the k ω SST model. The coefficient equal to 500 in Eq. (6) avoids that F 2 = 0 in the buffer layer.

The F2 function does not depend on the mesh size if an asymptotic solution is reached in the RANS equations, which means that GIS is eliminated by definition. The shape of the function for different velocity profiles is shown in Fig. 3. It covers the viscous sublayer and the log-layer with separate terms and switches from 1 to 0 between 0.7 0.8 δ and δ, which means that the transition from RANS to LES is done in a region with a characteristic length of 0.2 δ 0.3 δ. This is not optimal for GWMLES because the gray area covers approximately the entire outer layer. The F2 function is taken as reference for the development of the blending function f GWMLES proposed in this work, which consists of a modification that reads
(7)
where the constants n, c log, and c sub are constants that are derived as follows:
  • n: it controls the length of the gray area. The F2 function of Fig. 3 with the largest gray area of length 0.3 δ is taken as reference, which corresponds to Re2. The reduction of the gray area implies that n > 2. The case n = 8 gives a gray area of 0.1 δ, while a further increase to n = 12 does not provide relevant differences. The value of n chosen for the GWMLES log is n log = 8. A mesh size Δ = 0.1 δ is considered representative of a GWMLES log simulation in the outer layer; therefore, the transition takes place in one cell of the computational mesh. In the GWMLESsub mode, the choice is less universal, since the ratio δ sub / δ—with δ sub referred to the thickness of the viscous sublayer—depends on the Reynolds number. In GWMLESsub, considering that the mesh size Δ = 0.01 δ, then a similar reasoning to that of the GWMLES log model leads to a constant n sub = 16.

  • c log: it controls the value of y ̃ / δ in which the transition from RANS to LES occurs. In wall-bounded flows, the viscous sublayer and the log-layer constitute the universal part of the boundary layer, and the mesh requirements in the outer layer are of the same order as in the freestream region. Thus, it is assumed that for a mesh intending to solve 80% of the turbulence spectrum, the outer layer is solved in LES mode: thus, taking Fig. 4(a) as reference, c log = 2 is chosen for the GWMLES log mode. The model also allows to cover the entire boundary layer in RANS for coarse meshes by setting c log 16 [see Fig. 4(b)].

  • c sub: In both GWMLESsub and GWMLES log models, the goal is to solve the viscous sublayer and the buffer layer in RANS mode. This is the same idea behind the development of the k ω SST model; therefore, the second argument of f GWMLES [Eq. (7)] is proposed to be the same as the one of the F2 function [Eq. (6)], which means that c sub = 500. The GWMLESsub mode is activated by setting c log = 0, and in this case f GWMLES = 1 only up to the innermost part of the log-layer. The ratio between the thickness of the viscous sublayer y ̃ sub and δ depends strongly on the Reynolds number; therefore, in the GWMLESsub mode, the transition point y ̃ sub / δ from RANS to LES is not universal.

FIG. 3.

Blending function F2 and velocity profiles at different Reynolds numbers, in flows exposed to an adverse pressure gradient. Adapted from Ref. 49.

FIG. 3.

Blending function F2 and velocity profiles at different Reynolds numbers, in flows exposed to an adverse pressure gradient. Adapted from Ref. 49.

Close modal
FIG. 4.

GWMLES blending function for different combinations of model constants. (a) c log = 2 , c sub = 500 and varying n. (b) c sub = 500, n = 8, and varying c log.

FIG. 4.

GWMLES blending function for different combinations of model constants. (a) c log = 2 , c sub = 500 and varying n. (b) c sub = 500, n = 8, and varying c log.

Close modal

The parameters of GWMLES derived from the previous analyses are given in Table I.

TABLE I.

Parameters of the two formulations of the GWMLES model.

Mode c sub c log n
GWMLESsub  500  16 
GWMLES log  500 
Mode c sub c log n
GWMLESsub  500  16 
GWMLES log  500 

Ansys Fluent 2023R1 is used,50 which incorporates Eq. (5) by default, and the f GWMLES blending function defined in Eq. (7) is implemented with a user-defined function (UDF).

1. Mesh

The poly-hexcore method of Ansys Fluent is used for meshing, which creates a mesh consisting of octrees in the bulk region, and keeps a layered hexahedral mesh in the boundary layers. These elements are connected with general polyhedral elements (see Fig. 5), which have been used successfully in LES simulations of jets51 and external aerodynamics.52 The computational mesh is generated with different cell sizes Δ ranging from Δ min to Δ max, and the finest cell size is placed near the most relevant walls of the computational domain (see Fig. 5).

FIG. 5.

Poly-hexcore mesh used in the simulations performed in this work. The schematic is applicable to both GWMLESsub (where the RANS part covers up to the inner part of the log-layer) and GWMLES log (where the RANS part covers up to the outermost part of the log-layer).

FIG. 5.

Poly-hexcore mesh used in the simulations performed in this work. The schematic is applicable to both GWMLESsub (where the RANS part covers up to the inner part of the log-layer) and GWMLES log (where the RANS part covers up to the outermost part of the log-layer).

Close modal

The use of cells with an aspect ratio greater than 1 has been shown to deteriorate the accuracy of scale-resolving simulations.53–55 Thus, the mesh is generated so that f GWMLES = 1 (RANS) in the hexahedral cells with an aspect ratio greater than 1, and LES is considered in the rest of the domain. For flows at high Reynolds numbers, the instantaneous profile of f GWMLES is expected to be time-dependent due to velocity fluctuations. To take this into account and avoid that LES is solved in cells of high aspect ratios, the transition ratio of the boundary layer cells is adapted such that the aspect ratio of the outermost hexahedral cells is close to 1 (see Fig. 5). It should be noted that the mean (time-averaged) contour of the function f GWMLES is obtained after a precursor RANS simulation; that is, no unsteady simulations are needed to adapt the mesh to achieve the right numerical setup. The selection between GWMLESsub and GWMLES log is done so that f GWMLES = 0 (LES) in a computational cell if and only if the cell aspect ratio is close to 1; in particular, based on the cited references, values lower than 2 are considered acceptable. The set of constants of Table I is not optimal for all cases; for a given mesh, GWMLES log may lead to solving in RANS large regions suitable for LES, while GWMLESsub may lead to solving in LES a region that should be solved in RANS (with the corresponding risk of GIS). In those cases, the modification of the constant c sub is proposed. The effect of this modification will be assessed in one of the test cases presented in Sec. III.

2. Resolution of the equations

Equations (1)–(3) are solved with the finite volume method.56 Flow variables are stored at cell centers, and cell face values are computed with interpolation schemes. A central difference interpolation scheme is used in all terms of the equations except from the convective terms of the k and ω equations of the k ω SST model. Ansys Fluent does not allow the definition of non-diffusive schemes for these equations, although this is considered to be not relevant, since the equations for RANS turbulence variables are inherently diffusive. In addition, in the continuity equation, a corrected momentum interpolation is proposed to avoid pressure checkerboarding.57 Gradients are obtained by a least squares procedure, the temporal discretization is a second-order implicit scheme, and a linearized form of each equation for the flow variable Φ is
(8)
In Eq. (8), nb is used to denote neighbor cells, superindexes n and n + 1 refer to the current and next time step, respectively, and A and B are constants. The pressure–velocity coupling is performed with the SIMPLEC algorithm,56 and in each iteration, the systems of linear algebraic equations are solved using an algebraic multigrid method that uses the Gauss–Seidel method as smoother.58 The residuals of Eq. (8) in the mth iteration, which are globally scaled, are defined as follows:
(9)

It is verified that, at the end of the iterative process that is done every time step, the globally scaled residuals are < 5 × 10 4 (continuity), < 10 5 (momentum and turbulent variables), and < 10 8 (energy). These values are lower than those by default in Ansys Fluent of < 10 3 (continuity, momentum, and turbulent variables) and < 10 6 (energy). Simulations are started at t = t0 from a precursor RANS simulation (steady-state solver with second-order upwind discretization in the convective terms), and the time step dt is selected such that the maximum cell convective Courant number Co < 0.5. The same time step is used in all simulations of the mesh independence studies. In SBES and GWMLES models, the equations of the RANS turbulence variables (k and ω) are solved every 5 time steps. Simulations are finished at t = t1 when the flow statistics reach a statistically steady state (see Algorithm 1).

Algorithm 1. Generalized wall-modeled LES algorithm.

t = t0       ⊳ Precursor RANS simulation is converged 
while t < t 1 do      ⊳ LES solution not statistically steady 
  if mod ( t / d t , 5 ) = 0 then  ⊳ RANS solved every 5 time steps 
    μ t RANS k ω SST model   ⊳ different RANS model can be used 
  end if 
   μ t LES WALE model   ⊳ different LES model can be used 
   f GWMLES Eq. (7)     ⊳ implemented as UDF in Fluent (see  Appendix A
   μ t Eq. (5)       ⊳ included in Ansys Fluent 
  Solve Navier–Stokes equations Eqs. (1)–(3) 
   t = t + d t 
end while 
t = t0       ⊳ Precursor RANS simulation is converged 
while t < t 1 do      ⊳ LES solution not statistically steady 
  if mod ( t / d t , 5 ) = 0 then  ⊳ RANS solved every 5 time steps 
    μ t RANS k ω SST model   ⊳ different RANS model can be used 
  end if 
   μ t LES WALE model   ⊳ different LES model can be used 
   f GWMLES Eq. (7)     ⊳ implemented as UDF in Fluent (see  Appendix A
   μ t Eq. (5)       ⊳ included in Ansys Fluent 
  Solve Navier–Stokes equations Eqs. (1)–(3) 
   t = t + d t 
end while 

Simulations are performed in an eight-node cluster, with each node consisting of a 16-core AMD EPYC 7302 processors, which gives a total number of 128 cores. For each test case, the computational cost is obtained taking the following flow parameters as references:

  • Dimensionless time step, dt ̂: the ratio between the simulation time step dt and the characteristic time t c = L c / v , d t ̂ = d t / t c , where Lc is the flow characteristic length.

  • Dimensionless domain length, L ̂ Ω: the ratio between the length of the simulation domain L Ω and the characteristic length scale, L ̂ Ω = L Ω / L c.

  • Number of iterations per time step, N iter: for a given dimensionless time step, it is defined as the number of iterations performed by the numerical scheme before reaching the convergence conditions shown in previous paragraphs. As previously noted, the number of equations solved is not the same in all time steps for the case of SBES and GWMLES models; therefore, the value given is assumed to be the average value.

  • Iteration time t iter: computational time required to perform one iteration.

  • Number of flow times during transition N flow tr: number of times that the mean flow passes through the computational domain before the extraction of flow statistics starts.

  • Number of flow times during data extraction, N flow dat: number of times that the mean flow passes through the computational domain, while the flow statistics are extracted.

  • Simulation time t sim: time required to perform the entire simulation, including the extraction of flow statistics. It is calculated as t sim = ( L ̂ Ω / d t ̂ ) ( N flow t r + N flow dat ) N iter t iter.

3. Quality estimators

Three parameters related to the accuracy of the simulations will be considered.

  • Resolution quality, Q r: this is an estimate of the resolved turbulence spectrum2 
    (10)
    where k ¯ = ( u 2 + v 2 + w 2 ) / 2 is the resolved turbulent kinetic energy, and kr is the residual kinetic energy obtained from the kinetic energy transport model59 
    (11)

    for a value of the constant C k = 0.05 derived from simulations and experimental data.59 It is worth noting that, with additional equilibrium assumptions, the value of Ck can be also obtained from the Smagorinsky model.60 For a typical range of the Smagorinsky constant C s [ 0.1 , 0.2 ], the corresponding range of Ck is [ 0.05 , 0.12 ], which means that the value chosen from Ref. 59 gives a conservative prediction of the resolution quality.

  • Number of computational cells per integral length scale, N: since there is no universal method for estimating the resolution quality, another method will be considered that consists of computing the integral length scale, which is obtained as follows:2 
    (12)
    where R i i ( r , x , t ) = v i ( x , t ) v i ( x + r , t ) are the components of the two-point correlation tensor, with e i referring to the unit vector in the xi coordinate direction, and v i to the time-averaged value of vi. In this work, the spanwise integral length scale L will be calculated; therefore, i = 3, and L = L 33. The number of computational cells per integral length scale N is used as indicator for the quality of a LES simulation;61 in particular, in a subdomain S 1 Ω of the computational domain Ω,
    (13)

    An approximation of N / Δ can be obtained for the particular case of a slight variation of the von Kármán spectrum.62 The assumption that four cells are needed for a particular wavelength—which is considered as the characteristic length of the eddies—leads to a value of N 4 for Q r = 0.5, and N 10 for Q r = 0.8.

  • Q criterion, Q: it is defined as the second invariant of v,63 
    (14)

    where W is the vorticity tensor, and | | W | | = [ tr ( W W t ) ] 1 / 2 and | | S | | = [ tr ( S S t ) ] 1 / 2. The iso-surfaces of Q allow the visualization of the turbulence structures.

  • Taylor microscale, λ: the length scale that represents the onset of the turbulence dissipation range, which is defined from the two-point correlation function2 
    (15)
    The calculation of the Taylor microscale for meshes that differ more than one order of magnitude from the Kolmogorov scale can lead to large uncertainties, since the number of points close to the r = 0 is reduced. An approximation for the Taylor microscale will be used to estimate the cutoff limit of the turbulence spectrum of numerical solutions2 
    (16)

    where ϵ ¯ = ν | | S | | 2 is the resolved turbulent dissipation rate.

  • Normalized error, ϵ: in each test case, the numerical result g num will be compared with the reference experimental or DNS data g ref in a subdomain S 2 Ω of the computational domain Ω. The normalized error ϵ is defined as follows:
    (17)

The accuracy of the GWMLES model will be compared hereafter with other models such as SBES, DES family, and URANS, as well as with results from LES/WMLES simulations available in the literature. DNS results or experimental data will be taken as reference in each case. First, the subsonic round jet is studied in Sec. III A to verify the requirement that the blending function does not activate in turbulent regions far away from walls, which is equivalent to show that it gives a LES solution in a free shear flow; second, the periodic channel in Sec. III B serves as reference for the validation of the classical WMLES mode, which is given by GWMLESsub; third, the flow around the Ahmed body will be studied in Sec. III C, which is a complex turbulent wall-bounded flow used to test the GWMLES log model; finally, the capabilities of using a transition RANS model in the GWMLES log approach will be evaluated by studying the flow around a circular cylinder around the critical regime in Sec. III D.

The subsonic round jet has been extensively studied in the literature to obtain the acoustic properties of jets,64 which are key to the design of future aircraft engines.65 Taking experimental data as Ref. 66, simulation results will be compared with LES results obtained by the present authors, as well as with WMLES results available on the literature.67 

1. Domain and boundary conditions

The axisymmetric domain with respect to the x axis is shown in Fig. 6. The Mach number M = 0.5; therefore, the pressure at the outlet of the convergent nozzle is equal to the ambient pressure. The model assumes that the turbulence at the nozzle outlet is negligible compared to that produced by the Kelvin–Helmholtz instability. Thus, the nozzle is not part of the computational domain, and no turbulence fluctuations are defined at the inlet. The dimensions of the computational domain have been deduced after performing a domain independence study.

FIG. 6.

Computational domain Ω jet and boundary conditions of the subsonic round jet.

FIG. 6.

Computational domain Ω jet and boundary conditions of the subsonic round jet.

Close modal

The ambient pressure and temperature are p = 101 325 Pa and T = 298.15 K, while the Reynolds number R e D = 2.8 × 10 5 referred to D = 25.4 mm and the ambient density, which is obtained with the ideal gas law. At the inlet, the turbulence boundary conditions are referred to the RANS turbulence variables, so that the turbulent kinetic energy and the turbulent specific dissipation rate can be calculated. The turbulence intensity I = 2 k / 3 / u is obtained from Ref. 66. Like in the rest of the validation cases, it has been verified that the value of the inlet turbulent viscosity does not have an impact on the results. The boundary condition for the turbulent specific dissipation rate is the asymptotic solution in the viscous sublayer.

2. Results

a. Mesh independence study

Three different meshes are considered, with Q r [Eq. (10)] ranging from 0.5 to 0.8 in the region inside the edge of the shear layer. Table II shows that the estimation N = 10 [Eq. (13)] for Q r = 0.8 presented in Sec. II B 3 is met, so that both methods show that the resolution of the fine mesh is acceptable for a LES simulation. The normalized error decreases as the mesh size decreases. The differences between the fine and the medium mesh are lower than those between the medium and the coarse mesh; therefore, a convergent behavior is achieved.

TABLE II.

Results of the mesh independence study of the subsonic round jet performed with the GWMLES log model. The number of cells per integral length is obtained in S jet 1 = { x Ω jet | x / D = 10 , y = 0 , z / D 1 }, while the comparison with experimental data is done in the axis of revolution: S jet 2 = { x Ω jet | x / D 20 , y = 0 , z = 0 }. Experimental data obtained from Ref. 66.

Δ min Δ max Number of cells Q r N S jet 1 | | ϵ ( u ) | | S jet 2 | | ϵ ( u ) | | S jet 2
D / 20  D  0.4 M  0.5  2.3  0.062  0.253 
D / 40  D / 2  2.6 M  0.6  5.1  0.048  0.177 
D / 80  D / 4  17.3 M  0.8  9.5  0.039  0.142 
Δ min Δ max Number of cells Q r N S jet 1 | | ϵ ( u ) | | S jet 2 | | ϵ ( u ) | | S jet 2
D / 20  D  0.4 M  0.5  2.3  0.062  0.253 
D / 40  D / 2  2.6 M  0.6  5.1  0.048  0.177 
D / 80  D / 4  17.3 M  0.8  9.5  0.039  0.142 

In the fine mesh, the normalized error in the fluctuations is lower than 15%, which is related to the Q r value that is reasonably uniform along the jet [see Fig. 7(b)]. Performing a Richardson extrapolation does not lead to ϵ 0 for Δ 0, which is explained by two reasons: first, neglecting the turbulence content coming from the nozzle flow is expected to lead to errors that are important when Q r 1; second, as shown in Sec. III A 2 b, the experimental curve is the mean of several measurements; thus, for Q r 1, the uncertainty of the measurements dominates against the normalized error related to the simulations.

FIG. 7.

Turbulence variables of the GWMLES log model in the plane z = 0 of the subsonic round jet. (a) f GWMLES log . (b) Q r.

FIG. 7.

Turbulence variables of the GWMLES log model in the plane z = 0 of the subsonic round jet. (a) f GWMLES log . (b) Q r.

Close modal
b. Comparison between simulations and experimental data

Comparisons with experimental data will be shown for the fine mesh in all cases. Figure 8 shows the results of the mean and fluctuating axial velocities given by different models. The mean differences between GWMLES log and SBES are lower than 2%, and closer to experimental data than DDES in Fig. 8(d). Even if f GWMLES log is activated far from the jet region [see Fig. 7(a)], LES results (WALE) are shown to verify that the regions with f GWMLES log > 0 do not affect the results. The data available in the literature are for a WMLES case that includes the nozzle region, for a fine mesh consisting of 10 2 M elements. As expected, its accuracy is higher than the one of the simulations performed in this work. URANS gives a reasonable prediction of the mean flow, although this model significantly underpredicts the fluctuations. The normalized errors (Table II) of the URANS solution obtained with the fine mesh are larger than those given by GWMLES log with the coarse mesh in the velocity fluctuations; therefore, for free shear flows, using the GWMLES log model is recommended over URANS even for coarse meshes.

FIG. 8.

Velocity statistics of the subsonic round jet. (a) Mean x-velocity in the x axis. (b) Fluctuating x-velocity in the x axis. (c) Mean x-velocity at ( x / D , z / D ) = ( 4 , 0 ). (d) Fluctuating x-velocity at ( x / D , z / D ) = ( 4 , 0 ).

FIG. 8.

Velocity statistics of the subsonic round jet. (a) Mean x-velocity in the x axis. (b) Fluctuating x-velocity in the x axis. (c) Mean x-velocity at ( x / D , z / D ) = ( 4 , 0 ). (d) Fluctuating x-velocity at ( x / D , z / D ) = ( 4 , 0 ).

Close modal

Using a different residual stress model in GWMLES log (WALE) and DDES (Smagorinsky) does not lead to important variations in the results. However, as previously mentioned, the model constant used in the DDES model is too dissipative in free shear flows. This is translated into the resolution of coarser turbulence structures, in particular near the nozzle (see Fig. 9).

FIG. 9.

Iso-surfaces of Q =  5 × 10 9 s−2 of the subsonic round jet. (a) GWMLES log. (b) DDES.

FIG. 9.

Iso-surfaces of Q =  5 × 10 9 s−2 of the subsonic round jet. (a) GWMLES log. (b) DDES.

Close modal

The increased dissipation produced by DDES can be verified by looking at the turbulence spectrum at x / D = ( 10 , 0 , 0 ) [see Fig. 10(a)], which is outside of the potential core of the jet. Both turbulence models give almost the same spectrum for high wavenumbers. However, the cutoff limit is higher in GWMLES log, which is an important feature for acoustics calculations. This can be further verified quantitatively by calculating the Taylor microscale [see Eq. (15)]. The higher values of λ close to y = 0 in the case of x / D = 2 are because this zone is inside the potential core of the jet, where the development of the turbulence structures is limited. Analyzing the Q criterion, the turbulence spectrum, and the Taylor microscales leads to the same conclusion: GWMLES log solves finer turbulence structures than DDES.

FIG. 10.

Turbulence statistics of the subsonic round jet. (a) Spectrum of the resolved turbulent kinetic energy at x / D = ( 10 , 0 , 0 ). (b) Dimensionless Taylor microscale at z / D = 0.

FIG. 10.

Turbulence statistics of the subsonic round jet. (a) Spectrum of the resolved turbulent kinetic energy at x / D = ( 10 , 0 , 0 ). (b) Dimensionless Taylor microscale at z / D = 0.

Close modal

It has been verified that, for the same mesh, GWMLES log requires 30% lower computational cost than DDES, and 10% more computing time than LES (see Table III). The reason is that, in GWMLES log, RANS equations are solved every five time steps, while in DDES they are solved at all time steps, and no additional equation is solved for LES. Considering that the number of cores (128) is moderate, the simulation time is reasonable for an industrial framework.

TABLE III.

Simulation parameters of the subsonic round jet for a characteristic length scale Lc = D.

Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLES log  3.4 × 10 3  40  0.65  3.9 days 
DDES  3.4 × 10 3  40  0.9  5.9 days 
LES  3.4 × 10 3  40  0.55  3.6 days 
Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLES log  3.4 × 10 3  40  0.65  3.9 days 
DDES  3.4 × 10 3  40  0.9  5.9 days 
LES  3.4 × 10 3  40  0.55  3.6 days 

The periodic channel flow is a benchmark case that provides extensive information on near-wall turbulence without the requirement of simulating a large and/or complex domain. DNS results will be taken as Refs. 68 and 69, and the accuracy of the GWMLESsub model will also be compared with WMLES results obtained by other authors.70 The influence of the constant c sub on the performance of the model will be studied by considering two simulation approaches: the baseline with c sub = 500 ( GWMLES sub 500) and a variation with c sub = 1000 ( GWMLES sub 1000).

1. Domain and boundary conditions

In what follows, the air is assumed to be incompressible. The domain is a rectangular periodic channel (see Fig. 11), with the dimensions validated in Ref. 71 against larger domains. In the longitudinal (x) direction, the mass flow rate m ̇ is defined from the freestream velocity v , while in the spanwise direction (z) the mass flow rate vanishes. In the no-slip walls, the wall-parallel velocity is given by the solution in the viscous sublayer.48 The Reynolds numbers R e = 2 ρ u h / μ = 8.7 × 10 4 and R e τ = ρ u τ h / μ = 2000, where u τ is the friction velocity.

FIG. 11.

Computational domain Ωch and boundary conditions of the periodic channel.

FIG. 11.

Computational domain Ωch and boundary conditions of the periodic channel.

Close modal

2. Results

a. Mesh independence study

The results of the mesh independence study are shown in Table IV. The values given by GWMLES sub 500 converge toward a value whose normalized error is lower than 10% in all variables. This error is lower than 20% in the coarse mesh, even if the resolution quality is only 60%. This is attributed to the fact that this test is one of the classical benchmarks that are used to validate numerical methods and turbulence models, with no particular complex flow features such as recirculation or separation zones. The value of the (time-averaged) y ̃ w + = 1 is uniform due to the periodicity of the flow, and 10 cells normal to the wall are defined where f GWMLES = 1. Following the trends shown before, solving 80% of the turbulence spectrum corresponds to defining approximately ten computational cells per integral length scale. It can be concluded that both mesh quality indicators are related, with a relationship N ( Q r = 0.8 ) 10. Thus, in the following mesh independence studies, only the resolution quality will be specified.

TABLE IV.

Results of the mesh independence study of the periodic channel performed with the GWMLES sub 500 model. The number of cells per integral length is obtained in S c h 1 = { x Ω c h | x / h = π , y / h = 8 , 0 z / h 4 }, while the comparison with DNS is done in the boundary layer S c h 2 = { x Ω c h | x / h = π , 0 y / h 1 , z = 0 }. DNS data obtained from Refs. 68 (u) and 69 (Cf).

Δ min + Δ max + Number of cells Q r N S ch 1 | | ϵ ( u ) | | S ch 2 | | ϵ ( u ) | | S ch 2 | | ϵ ( C f ) | | S ch 2
100  400  1.8 M  0.6  2.4  0.039  0.193  0.08 
50  200  7.2 M  0.7  7.9  0.015  0.123  0.02 
25  100  28.9 M  0.85  14.7  0.009  0.087  0.01 
Δ min + Δ max + Number of cells Q r N S ch 1 | | ϵ ( u ) | | S ch 2 | | ϵ ( u ) | | S ch 2 | | ϵ ( C f ) | | S ch 2
100  400  1.8 M  0.6  2.4  0.039  0.193  0.08 
50  200  7.2 M  0.7  7.9  0.015  0.123  0.02 
25  100  28.9 M  0.85  14.7  0.009  0.087  0.01 

As expected from the reasoning followed in Sec. III A 1, most of the domain is solved in LES mode [see Fig. 12(a)]. Details of the transition from RANS to LES will be given below. The resolution quality is uniform in all meshes considered [see Fig. 12(b)].

FIG. 12.

Turbulence variables of the GWMLES sub 500 model in the plane z = 0 of the periodic channel. (a) f GWMLES sub. (b) Q r.

FIG. 12.

Turbulence variables of the GWMLES sub 500 model in the plane z = 0 of the periodic channel. (a) f GWMLES sub. (b) Q r.

Close modal
b. Comparison between simulations and experimental data

The comparison between velocity statistics in the boundary layer given by different simulations is illustrated in Fig. 13. The agreement with DNS data is excellent in the GWMLESsub and SBES models, while IDDES predicts a 50% lower thickness of the log-layer [see Fig. 13(a)]. The WMLES results70 are obtained with a wall function approach and provide great agreement in the log-layer, although errors up to 20% in the mean dimensionless x-velocity are noticed in the viscous sublayer and the buffer layer.

FIG. 13.

Velocity statistics in the boundary layer of the periodic channel. (a) Mean dimensionless x-velocity. (b) Fluctuating dimensionless x-velocity. (c) Fluctuating dimensionless y-velocity. (d) Fluctuating dimensionless z-velocity.

FIG. 13.

Velocity statistics in the boundary layer of the periodic channel. (a) Mean dimensionless x-velocity. (b) Fluctuating dimensionless x-velocity. (c) Fluctuating dimensionless y-velocity. (d) Fluctuating dimensionless z-velocity.

Close modal

The deviations between different methods are more pronounced in the fluctuating variables. GWMLESsub, SBES, and IDDES results differ up to 20% in Fig. 13(b), although they are significantly more accurate than WMLES [see Figs. 13(b)–13(d)]. The mesh is coarser in the simulation of Ref. 71 ( Δ min + = 79), but this is partially balanced by the fifth-order spectral difference method that is used. Independently of the mesh and the numerical method, the issues of the wall function approach have been remarked in the introduction. This is relevant, since the fluctuations in the buffer layer are of the same order of magnitude as those in the log-layer. Even if RANS models are calibrated for steady cases, it is demonstrated that the model developed in this work is able to capture the velocity fluctuations that are produced in the viscous sublayer and the log-layer, while the wall function approach leads to errors that are as high as 50% in Fig. 13(b). Solving the viscous sublayer and the buffer layer in RANS leads to a higher computational cost than with the wall function-based approach. Nevertheless, in these zones, the cell aspect ratio is much higher than 1; therefore, both mesh sizes are of the same order of magnitude.

URANS simulations have also been performed, leading to a deviation lower than 1% in the mean velocity, and with negligible fluctuations. Starting from a precursor RANS simulation, the high dissipation of URANS leads to no fluctuations. The same conclusion is extracted after sufficient time if the URANS simulation is initialized from a GWMLESsub solution. This could be anticipated before performing the simulation, since all parts of the domain are inside the boundary layer, with no relevant flow instabilities. Given that in GWMLESsub, the normalized error of the mean values obtained with the coarse mesh is lower than 4% (see Table IV), and that URANS predicts negligible fluctuating variables, GWMLESsub provides more accurate results than URANS in the periodic channel flow for all meshes considered.

The increase in c sub defined in Eq. (7) leads to a RANS-LES transition at higher y ̃ + (see Fig. 14). The transition region of the mean value of the blending function contains more than two cells. However, as illustrated in Fig. 12(a), the fluctuations in the function are important because of the high resolution quality, which means that the time-averaged value is smoothed with respect to the instantaneous value; that is, in Fig. 14, the transition region is of the order of y ̃ +, while Fig. 12(a) shows that the instantaneous transition region is much thinner. The mean SBES blending function is essentially the GWMLES sub 500 function translated approximately by y ̃ + = 5. In GWMLES sub 1000, the y ̃ + in which RANS-LES transition takes place is twice the one of GWMLES sub 500, with a lower slope of the function. No LLM is noticed in Fig. 13(a), and the velocity fluctuations are underpredicted, which is attributed to the higher dissipation as a consequence of the extended RANS region. The great accuracy of the GWMLESsub, SBES, and IDDES models is also proved in the mean skin friction coefficient, which is predicted with an error lower than 5% in all models (see Table V). This study confirms the capabilities of the proposed model to improve the current wall function and mixing-length approaches for WMLES.

FIG. 14.

Mean blending functions in the boundary layer of the periodic channel.

FIG. 14.

Mean blending functions in the boundary layer of the periodic channel.

Close modal
TABLE V.

Mean skin friction coefficient of the periodic channel.

Parameter Simulation C f ( × 1 0 3)
DNS69   4.375 
GWMLES sub 500  4.385 
GWMLES sub 1000  4.392 
SBES  4.381 
IDDES  4.396 
URANS  4.391 
Parameter Simulation C f ( × 1 0 3)
DNS69   4.375 
GWMLES sub 500  4.385 
GWMLES sub 1000  4.392 
SBES  4.381 
IDDES  4.396 
URANS  4.391 

The IDDES model has been particularly calibrated for periodic channel flows.6 Moreover, there are no transition regions, which means that the transition from modeled-to-resolved turbulence is not as relevant as in other more complex wall-bounded flows. This can be verified in the size of the resolved turbulence structures, in which no qualitative differences can be observed between GWMLESsub and IDDES (see Fig. 15). Quantitative results derived from the distribution of the Taylor length scale lead to the same conclusion (see Fig. 16).

FIG. 15.

Iso-surfaces of Q = 106 s−2 in the boundary layer of the periodic channel. (a) GWMLES sub 500. (b) IDDES.

FIG. 15.

Iso-surfaces of Q = 106 s−2 in the boundary layer of the periodic channel. (a) GWMLES sub 500. (b) IDDES.

Close modal
FIG. 16.

Dimensionless Taylor microscale in the boundary layer of the periodic channel.

FIG. 16.

Dimensionless Taylor microscale in the boundary layer of the periodic channel.

Close modal

The computational cost for GWMLESsub (same as for SBES) is around 35% lower than the one for DDES (Table VI). The reduction is greater than in the subsonic round jet case, since in compressible flow RANS equations represent 2 / 7 = 28 % of the total number of equations, while in incompressible flow without heat transfer the ratio is 2 / 6 = 33 %. The computational cost is of the same order as the one of the simulations of the subsonic round jet (see Table III). The higher cell count and lower time step, as well as the large number of time steps required for the turbulence structures to be fully developed, are compensated by the reduced numbers of iterations per time step and dimensionless length of the computational domain.

TABLE VI.

Simulation parameters of the periodic channel for a characteristic length scale Lc = h.

Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLESsub  1.4 × 10 3  2 π  0.9  15  2.8 days 
IDDES  1.4 × 10 3  2 π  1.35  15  4.2 days 
Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLESsub  1.4 × 10 3  2 π  0.9  15  2.8 days 
IDDES  1.4 × 10 3  2 π  1.35  15  4.2 days 

The Ahmed body is a model of a generic road car72 (see Fig. 17). The flow shares some features with the real cars such as a large three-dimensional recirculation zone at the back of the geometry, and the interaction of two counter-rotating vortices that are generated in the slant edges. For slant angles lower than the critical angle of 30°, the flow separates in the beginning of the slant, and it reattaches downstream, leading to complex flow features. Taking experimental data as Refs. 72–75, simulation results will be compared with the results obtained from LES and WMLES simulations available in the literature.76,77

FIG. 17.

Geometry of the Ahmed body.72 Dimensions in mm. (a) Back view. (b) Lateral view. (c) Bottom view.

FIG. 17.

Geometry of the Ahmed body.72 Dimensions in mm. (a) Back view. (b) Lateral view. (c) Bottom view.

Close modal

1. Domain and boundary conditions

The computational domain is a 3/4 open wind tunnel with the dimensions of the cross section obtained from Ref. 73 (see Fig. 18). The lengths upstream and downstream of the Ahmed body have been obtained after performing a domain independence study. The Reynolds number R e L = 2.7 × 10 6, and the turbulence intensity u / u at the inlet, which is used to calculate the turbulent kinetic energy and the turbulent specific dissipation rate, is obtained from Ref. 73.

FIG. 18.

Computational domain Ω ahmed and boundary conditions of the Ahmed body.

FIG. 18.

Computational domain Ω ahmed and boundary conditions of the Ahmed body.

Close modal

2. Results

a. Mesh independence study

Three different meshes are considered, with Q r in the slant region ranging from 0.5 to 0.8. Coarser meshes lead to numerical instabilities when the low-dissipation discretization methods mentioned in Sec. II B 2 are used. Table VII shows that the error in the velocities of the boundary layer decreases as the mesh size decreases, and a convergent behavior is achieved.

TABLE VII.

Results of the mesh independence study of the Ahmed body performed with the GWMLES log model in the boundary layer of the slant: S ahmed = { x Ω ahmed | x = 0.043 , y = 0 , 0.25 z 0.4 }. Experimental data obtained from Ref. 73.

Δ min + Δ max + Numberof cells Q r | | ϵ ( u ) | | S ahmed | | ϵ ( u ) | | S ahmed C d C d , f
375  6400  2.4 M  0.5  0.194  0.488  0.346  0.057 
275  4800  4.5 M  0.65  0.167  0.351  0.340  0.056 
175  3200  11.5 M  0.8  0.149  0.264  0.335  0.056 
Δ min + Δ max + Numberof cells Q r | | ϵ ( u ) | | S ahmed | | ϵ ( u ) | | S ahmed C d C d , f
375  6400  2.4 M  0.5  0.194  0.488  0.346  0.057 
275  4800  4.5 M  0.65  0.167  0.351  0.340  0.056 
175  3200  11.5 M  0.8  0.149  0.264  0.335  0.056 

The drag parameters—with C d , f referring to the friction component of the drag coefficient—are not compared with experimental data; the reason is that there is no consensus based on different experiments. In the normal direction of the boundary layers, 20 cells are defined, with y ̃ w + < 2 in the turbulent boundary layers [see Fig. 19(a)]. These conditions imply small differences in the quantities that strongly depend on the RANS model, such as the friction component of the drag coefficient. In contrast, there are separation and reattachment processes, and the drag coefficient and the fluctuating values are mainly determined by the accuracy of the LES region. This is the reason why these variables are more sensitive to mesh refinement. In the fine mesh, Δ min < 0.1 δ, even though, as noted in Ref. 20, the flow is not exposed to an adverse pressure gradient on the roof; therefore, GIS is not a problem for the present geometry.

FIG. 19.

Turbulence variables of the GWMLES log model in the plane y = 0 of the Ahmed body. (a) Mean dimensionless wall distance of the first near-wall cell. (b) f GWMLES log . (c) Q r.

FIG. 19.

Turbulence variables of the GWMLES log model in the plane y = 0 of the Ahmed body. (a) Mean dimensionless wall distance of the first near-wall cell. (b) f GWMLES log . (c) Q r.

Close modal

It should be noted that the normalized error of the fluctuating velocity of the fine mesh is still 26%. Two universal reasons are related to the use of a RANS model in the boundary layers, as well as the limited resolution of the turbulence spectrum. Even if Q r 0.8 in the slant, there are some regions in the wake where the value is 0.6 0.7 [see Fig. 19(c)]. In addition, in the experimental data available, there are only ten data points in the boundary layer, which are only spaced by 4 mm. This means that the accuracy of the measuring technique could have an important role in the comparison, as well as the manufacturing tolerances of the body. In Sec. III C 2 c, the results will be compared with data available on the literature, so that a better picture can be obtained in regard to the accuracy of the model proposed.

b. GWMLESlog blending function

The mean profiles of the x-velocity and the GWMLES log blending function in different positions of the roof of the Ahmed body are presented in Fig. 20. The equilibrium solutions in the viscous sublayer u sub + = y ̃ + and the log-layer u log + = C + + log ( y ̃ + ) / κ for values of the von Kármán constant κ = 0.41 and C + = 5 (Ref. 48) are plotted as reference. Figure 20(a) shows that the boundary layer is not fully developed in the first three reference points. The transition point from RANS to LES depends on the x-coordinate [see Fig. 20(b)], since the log-layer thickness is proportional to the distance to the leading edge when the flow is not fully developed. At x = −250 mm, the profile obtained from the numerical simulations matches the law of the wall. The mesh size in the roof region is Δ 0.05 δ (half of the size used in the slant region), and the transition from RANS to LES takes place in around two cells. This is the expected result, given that the reference considered in Sec. II A 1 was 1 cell for Δ 0.1 δ. In the fully developed case (x = −250 mm), the transition point is located near the end of the log-layer, as predicted.

FIG. 20.

Mean boundary layer profiles the plane y = 0 of the Ahmed body. (a) Dimensionless x-velocity. (b) Blending function.

FIG. 20.

Mean boundary layer profiles the plane y = 0 of the Ahmed body. (a) Dimensionless x-velocity. (b) Blending function.

Close modal
c. Comparison between simulations and experimental data

Figure 21 compares the streamwise velocity statistics in the slant region of the symmetry plane (see Fig. 17) with experimental data, LES, and WMLES simulations. The results of LES and WMLES simulations are those obtained with the fine mesh used in the corresponding work. The differences between GWMLES log, SBES, and DDES are lower than 5% at almost all points, especially in the mean velocity. These models are more accurate than WMLES in the mean values, although the deviation from the experimental data is more pronounced than in the LES simulations. Even if the accuracy of all models compared is similar in the fluctuating velocity, at x = −203 mm, there are some fluctuations predicted by the GWMLES log, SBES, and LES models that are not present in the DDES model and in the experimental data. URANS fails to determine the velocity profiles, with negligible temporal fluctuations; indeed, the normalized errors (Table VII) are greater than those obtained with GWMLES log and the coarse mesh.

FIG. 21.

Velocity statistics in the plane y = 0 of the Ahmed body. (a) Mean x-velocity. (b) Fluctuating x-velocity.

FIG. 21.

Velocity statistics in the plane y = 0 of the Ahmed body. (a) Mean x-velocity. (b) Fluctuating x-velocity.

Close modal

Figure 21(a) allows the detection of the separation and reattachment point from different models. In the attached flow, the streamwise velocity component is positive, while the opposite occurs when the flow is separated. Except from URANS, in which the flow is separated along the entire slant, the separation at the beginning of the slant and the subsequent reattachment before the end are predicted by all models. However, GWMLES log, SBES, and DDES models overpredict the width of the separated zone, while LES and WMLES provide the opposite behavior. It should be noted that a positive streamwise velocity component is not a sufficient condition for the flow to be attached, and it has been verified that in these cases, the vertical (z) velocity component is negative, as required to deduce that the flow is attached.

The effect of the prediction of the transition point will be analyzed in the drag parameters. Among the models that give a reasonable prediction of the velocity statistics, since the separated region is the largest in the DDES results, it is expected that DDES provides the highest drag coefficient value. In contrast, the lowest should be given by the LES model given the overprediction of the attached region by this model. This is indeed shown in Table VIII. As previously anticipated, the value of the drag coefficient is not clear taking previous articles as references. The quality of the test sample has a strong influence on the drag coefficient;75 it has been shown that small roundings in the upper edge of the slant lead to a decrease in the drag coefficient, and, in particular, the drag coefficient does not depend on the radius of curvature in the cases studied. That is, there is a reference value in the body with a sharp edge, 0.356, which is similar to the one obtained in Ref. 74. The drag coefficient reduces to 0.297 with a rounded edge, which is closer to the one of Ref. 72. Another key parameter is the friction component of the drag. The values given by GWMLES log, SBES, and DDES models differ by less than 10% and agree with the value given in Ref. 72. LES results provide a low value of the friction component, which may be a reason for the discrepancy in C d with respect to Refs. 74 and 75. Since friction effects are dominant in the attached flow, the friction component given by URANS is in line with the one given by GWMLES log, SBES, and DDES. Given the large deviations of the URANS result with respect to experimental data in the velocity statistics, it is expected that the friction coefficient predicted by URANS deviates significantly, although this is not the case. Furthermore, flow variables need to be considered to explain this behavior.

TABLE VIII.

Drag parameters of the Ahmed body.

Parameter EXP72  EXP74  EXP75  GWMLES log SBES DDES LES76  URANS
C d   0.285  0.343  0.356  0.335  0.331  0.347  0.295  0.298 
C d , f   ⋯  0.051  ⋯  0.055  0.057  0.052  0.037  0.052 
Parameter EXP72  EXP74  EXP75  GWMLES log SBES DDES LES76  URANS
C d   0.285  0.343  0.356  0.335  0.331  0.347  0.295  0.298 
C d , f   ⋯  0.051  ⋯  0.055  0.057  0.052  0.037  0.052 

Even if the drag coefficient is not obtained in the experiments taken as reference in Fig. 21 (Ref. 73), the pressure coefficient distributions in the slant and the rear surface of the Ahmed body are available. Figure 22 compares numerical and experimental results in the slant and shows that simulation values provide an accurate overall picture of the distribution. However, errors up to 20 % are obtained in the numerical simulations (see Table IX). As expected from the comparison of velocity statistics, the prediction of URANS is of reduced accuracy, since this model does not predict clearly two of the main flow physics: the separation and subsequent reattachment, as well as the two counter-rotating vortices, which are extremely weak [see Fig. 22(a)]. It has been verified that the normalized errors obtained with the coarse mesh in GWMLES log are lower than those of URANS in the fine mesh, which leads to the same conclusion as in the previous cases: GWMLES provides more accurate results than URANS even in coarse meshes.

FIG. 22.

Pressure coefficient in the slant of the Ahmed body. Due to the limited number of sampling points of the LDA measurements, the contours cover the spanwise domain 0.1875 y 0.1875, which is smaller than the width of the Ahmed body ( 0.1945 y 0.1945). (a) Experiment.73 (b) GWMLES log. (c) SBES. (d) DDES. (e) URANS.

FIG. 22.

Pressure coefficient in the slant of the Ahmed body. Due to the limited number of sampling points of the LDA measurements, the contours cover the spanwise domain 0.1875 y 0.1875, which is smaller than the width of the Ahmed body ( 0.1945 y 0.1945). (a) Experiment.73 (b) GWMLES log. (c) SBES. (d) DDES. (e) URANS.

Close modal
TABLE IX.

Pressure coefficients of the Ahmed body.

Parameter EXP73  GWMLES log SBES DDES URANS
C p slant   −0.585  −0.499  −0.487  −0.534  −0.365 
C p rear   −0.161  −0.161  −0.158  −0.171  −0.171 
Parameter EXP73  GWMLES log SBES DDES URANS
C p slant   −0.585  −0.499  −0.487  −0.534  −0.365 
C p rear   −0.161  −0.161  −0.158  −0.171  −0.171 

The contours of the pressure coefficient are less accurate at first sight in the rear of the body, which corresponds to the plane x = 0 (see Fig. 23). However, the area-averaged value is highly accurate in all cases, even in URANS (see Table IX); the latter is considered to be a coincidence, since the pointwise contour of URANS [Fig. 23(b)] deviates significantly from the experimental data [Fig. 23(a)]. Assuming that there is no error in the remaining surfaces, from the results of Table IX, it is deduced that GWMLES log underpredicts the drag obtained in the experiments of Ref. 73. This enforces the assumption that the correct drag coefficient is 0.34, so that the hybrid RANS/LES models would provide a better accuracy than the LES model for a much lower computational cost. In that situation, GWMLES log model is more accurate than the SBES, which can be due to solving the outer layer in LES mode instead of in RANS.

FIG. 23.

Mean pressure coefficient in the rear of the Ahmed body. (a) Experiment.73 (b) GWMLES log. (c) SBES. (d) DDES. (e) URANS.

FIG. 23.

Mean pressure coefficient in the rear of the Ahmed body. (a) Experiment.73 (b) GWMLES log. (c) SBES. (d) DDES. (e) URANS.

Close modal

In addition to the baseline GWMLES log with the k ω SST RANS model and the WALE residual stress model ( GWMLES log WALE), another simulation has been performed with the k ω SST and Smagorinsky models ( GWMLES log SMAG). This allows us to study separately the effect of the subgrid-scale model and f GWMLES log on the resolved turbulence structures. The Smagorinsky constant C s = 0.175, which is equal to the equivalent Smagorinsky constant of the DDES model, as mentioned in the introduction. Even if the accuracy of GWMLES log and DDES in terms of mean and fluctuating variables has been shown to be similar, Fig. 24 shows that both the use of a more consistent subgrid stress model (WALE) and the reduction of gray areas contribute to the increased resolution. In fact, the differences in the size of the turbulence structures between GWMLES log WALE and DDES are more noticeable than in the previous test cases, with a reduced computational cost required by the GWMLES log with respect to DDES (see Table X).

FIG. 24.

Iso-surfaces of Q =  2.5 × 10 6 s−2 in the slant of the Ahmed body. (a) GWMLES log WALE. (b) GWMLES log SMAG. (c) DDES.

FIG. 24.

Iso-surfaces of Q =  2.5 × 10 6 s−2 in the slant of the Ahmed body. (a) GWMLES log WALE. (b) GWMLES log SMAG. (c) DDES.

Close modal
TABLE X.

Simulation parameters of the Ahmed body for a characteristic length scale Lc = H.

Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLESsub  1.3 × 10 3  35  0.5  7.8 days 
DDES  1.3 × 10 3  35  0.75  11.7 days 
Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLESsub  1.3 × 10 3  35  0.5  7.8 days 
DDES  1.3 × 10 3  35  0.75  11.7 days 

The flow around a circular cylinder at R e 10 5 is representative of the laminar–turbulent transition process,78 with the type of transition depending on Re and the particular conditions of the experiments. In this test case, experimental data79 and LES results80,81 will be used for comparison. The work80 considers many cases with different spanwise lengths, mesh sizes, and turbulence models. It has been considered as reference the case with the same spanwise length as in this work (2D), the fine mesh, and the dynamic Smagorinsky residual stress model. Two simulation approaches will be tested within the GWMLES log model: the standard k ω SST model for turbulent boundary layers ( GWMLES log SST), and the γ R e θ t model for boundary layers with laminar–turbulent transition ( GWMLES log γ R e θ t). The latter adds two additional transport equations for the intermittency γ and the transition momentum thickness Reynolds number R e θ t.35 In both cases, the residual stress model is the baseline WALE model.

1. Domain and boundary conditions

The computational domain is a circular cylinder with periodic boundary conditions in the spanwise direction (see Fig. 25). The distance between the periodic boundaries has been deduced from a domain independence study. The definition of the inlet boundary conditions for the RANS turbulence variables has been done with the aim of representing laminar freestream conditions, which correspond to the LES results used for comparison and the aim of the experiment taken as Ref. 80. The Reynolds number of the flow R e D = 1.4 × 10 5.

FIG. 25.

Computational domain Ωcyl and boundary conditions of the circular cylinder.

FIG. 25.

Computational domain Ωcyl and boundary conditions of the circular cylinder.

Close modal

2. Results

a. Mesh independence study

Three different meshes are considered, with Q r in the wake of the cylinder ranging from 0.6 to 0.8. Table XI shows that the error in the velocity statistics and the mean drag coefficient converges to a constant value. In the following paragraphs, the reason why ϵ ̸ 0 when Q r 0 will be explained. The increase in the mesh resolution does not have a strong impact on the results, which is attributed to the low y w + value that is kept constant in all meshes (see Fig. 26), as well as the definition of 20 boundary layer cells in the normal direction. This study also allows us to conclude that RANS zones play an important role in the flow.

TABLE XI.

Results of the mesh independence study of the circular cylinder performed with the GWMLES log γ R e θ t model in the wake axis S cyl = { x Ω cyl | 1 x / D 8 , y / D = 0 , z / D = 0 }. Experimental data obtained from Ref. 79.

Δ min Δ max Number of cells Q r | | ϵ ( u ) | | S cyl | | ϵ ( u ) | | S cyl | | ϵ ( C d ) | | S cyl
D / 125  D / 4  2.4 M  0.6  0.114  0.402  0.502 
D / 175  D / 4  4.8 M  0.7  0.134  0.373  0.61 
D / 250  D / 4  12.1 M  0.8  0.129  0.365  0.674 
Δ min Δ max Number of cells Q r | | ϵ ( u ) | | S cyl | | ϵ ( u ) | | S cyl | | ϵ ( C d ) | | S cyl
D / 125  D / 4  2.4 M  0.6  0.114  0.402  0.502 
D / 175  D / 4  4.8 M  0.7  0.134  0.373  0.61 
D / 250  D / 4  12.1 M  0.8  0.129  0.365  0.674 
FIG. 26.

Mean dimensionless wall distance of the first near-wall cell of the circular cylinder.

FIG. 26.

Mean dimensionless wall distance of the first near-wall cell of the circular cylinder.

Close modal

Figure 27(a) shows that GWMLES log γ R e θ t runs in LES mode only in the wake of the cylinder, which is not the expected behavior. However, the RANS zones outside the boundary layers are fully resolved, since there μ t / μ 1 [see Fig. 27(b)]. The turbulence variables k, ω, and μt almost vanish at the inlet. These values are not significantly increased outside the boundary layer or the wake, since the production terms in the k and ω equations are proportional to the strain rate that vanishes in these zones. This is because there is no wall or other turbulence generator than can produce velocity gradients. Thus, the second argument of the blending function defined in Eq. (7) activates without compromising the accuracy of the modeling approach.

FIG. 27.

Turbulence variables of the GWMLES log model in the plane z = 0 of the circular cylinder. (a) f GWMLES log . (b) μ t / μ.

FIG. 27.

Turbulence variables of the GWMLES log model in the plane z = 0 of the circular cylinder. (a) f GWMLES log . (b) μ t / μ.

Close modal
b. Comparison between simulations and experimental data

Figure 28 shows the velocity statistics on the longitudinal axis and on one of the vertical axes close to the cylinder surface. The GWMLES log SST, SBES, and DDES models are more accurate than the LES cases considered in Figs. 28(a) and 28(c), with mean differences with respect to experimental data lower than 20%. The opposite occurs in Figs. 28(b) and 28(d); therefore, no clear conclusion about the accuracy of the models can be extracted. GWMLES log γ R e θ t results deviate significantly with respect to other data in the fluctuating variables and provide the lowest turbulence levels, while the highest levels are obtained with the LES simulations.

FIG. 28.

Velocity statistics of the circular cylinder. (a) Mean x-velocity in the x axis. (b) Fluctuating x-velocity in the x axis. (c) Fluctuating y-velocity in the x axis. (d) Mean y-velocity at ( x , z ) = ( D .0 ).

FIG. 28.

Velocity statistics of the circular cylinder. (a) Mean x-velocity in the x axis. (b) Fluctuating x-velocity in the x axis. (c) Fluctuating y-velocity in the x axis. (d) Mean y-velocity at ( x , z ) = ( D .0 ).

Close modal

Furthermore, comparison of flow parameters is presented in Table XII. Defining the separation angle φ sep as the value at the inflection point of the pressure coefficient, in LES, the flow separates at the lowest value of φ, leading to increased drag, while the opposite occurs in GWMLES log γ R e θ t. GIS is not a problem for DDES in the meshes considered, since Δ min > 0.3 δ due to the thin boundary layer. The differences in the mean base pressure coefficient C p , b = C p ( φ = 180 ° ) follow the same trends as in the drag coefficient, and the Strouhal number increases notably in the laminar–turbulent transitional flow, in line with the data available in the literature.78 

TABLE XII.

Parameters of the flow around the circular cylinder.

Parameter EXP79  GWMLES log SST SBES DDES GWMLES log γ R e θ t LES80  LES81 
C d   1.24  0.79  0.79  0.84  0.4  1.45  1.43 
C p , b   −1.21  −0.85  −0.84  −0.83  −0.39  −1.76  −1.59 
St  0.18  0.24  0.25  0.24  0.33  0.2  0.19 
φ sep (deg)  90  105  105  110  120  ⋯  95 
Parameter EXP79  GWMLES log SST SBES DDES GWMLES log γ R e θ t LES80  LES81 
C d   1.24  0.79  0.79  0.84  0.4  1.45  1.43 
C p , b   −1.21  −0.85  −0.84  −0.83  −0.39  −1.76  −1.59 
St  0.18  0.24  0.25  0.24  0.33  0.2  0.19 
φ sep (deg)  90  105  105  110  120  ⋯  95 

The discrepancies of the model developed in this work in the velocity statistics are translated into the pressure coefficient statistics. The mean Cp is well predicted until φ = 60 °, and after this point, all models overpredict the absolute value of the pressure minimum [see Fig. 29(a)]. One reason for the lack of accuracy of GWMLES log SST, SBES, and DDES may be that the flow is laminar in the experiment, and then, the separation-induced transition to turbulence occurs around φ = 90 °. In that case, the turbulence models treat all the boundary layer as fully turbulent, which leads to the failure of the prediction of the transition point. This reasoning is not applicable to GWMLES log γ Re θ t, which solves the Navier–Stokes equation without modeled turbulence at least in a portion of the boundary layer. Additional variables need to be studied to understand the physical reasoning behind the delayed separation point provided by this model.

FIG. 29.

Pressure coefficient statistics of the circular cylinder. (a) Mean value. (b) Fluctuating value.

FIG. 29.

Pressure coefficient statistics of the circular cylinder. (a) Mean value. (b) Fluctuating value.

Close modal

Even if no data of the fluctuating component of the Cp are presented, the article cited79 provides the peak-to-peak values of the pressure coefficient, C p p p. The experimental C p value given by Fig. 29(b) has been obtained assuming that the C p = C p ( t ) function is sinusoidal, and in this case, C p = C p p p / 2, which serves as an approximation. Figure 29(b) shows that the fluctuations are much higher in the experiments than in the simulations in the entire domain, in particular, in the region of favorable pressure gradient, which is supposed to be laminar, and the high experimental C p calls this into question. If the transition mechanism in the experiment is due to the separation-induced transition, then a laminar separation bubble would be expected, which is not the case, although possibly due to the lack of resolution of the experimental devices.

Based on the laminar boundary conditions considered, the flow is laminar at the leading edge of the cylinder. GWMLES log SST provides μ t > 0 even at φ < 60 ° [see Fig. 30(a)], while GWMLES log γ R e θ t keeps μ t 0 up to φ 110 ° inside the thin boundary layer. In GWMLES log SST, from φ = 0 ° intensity of the pressure fluctuations increases as φ increases, while in GWMLES log γ R e θ t, this happens notably for φ = 60 ° [see Fig. 29(b)]. Thus, the transition model is expected to capture from φ 60 ° some unstable Tollmien–Schlichting waves that are the mechanism of the natural laminar–turbulent transition process. This can be verified in Fig. 31(b).

FIG. 30.

Mean turbulent viscosity ratio in the boundary layer of the circular cylinder. (a) GWMLES log SST. (b) GWMLES log γ R e θ t.

FIG. 30.

Mean turbulent viscosity ratio in the boundary layer of the circular cylinder. (a) GWMLES log SST. (b) GWMLES log γ R e θ t.

Close modal
FIG. 31.

Iso-surfaces of Q =  2.5 × 10 7 s−2 in the boundary layer of the circular cylinder. (a) GWMLES log SST. (b) GWMLES log γ R e θ t.

FIG. 31.

Iso-surfaces of Q =  2.5 × 10 7 s−2 in the boundary layer of the circular cylinder. (a) GWMLES log SST. (b) GWMLES log γ R e θ t.

Close modal

Since the resolution quality Q r μ t 2 [see Eq. (10)], GWMLES log γ R e θ t keeps Q r 1 up to φ = 110 ° [see Fig. 32(b)]. In contrast, in GWMLES log SST, the turbulence production terms activate and the resolution quality deteriorates from φ = 60 °, even for μ t / μ < 5 [see Fig. 30(a)]. This translates into the resolution of coarser turbulent structures, which are notably damped around the separation point. Thus, flow separation is a consequence of the low resolution in GWMLES log SST, SBES (SST), and DDES. This fact explains the delayed separation of GWMLES log γ R e θ t, since in this model, the resolution drops only after the flow is fully separated. These findings emphasize the importance of using a low-dissipation numerical scheme. DDES can also be coupled with the γ R e θ t ( DDES γ R e θ t), but in this case, the number of RANS equations is four, while the total number of equations is eight. This means that GWMLES log γ R e θ t provides up to 50% of reduction of the computational cost with respect to DDES γ R e θ t (see Table XIII).

FIG. 32.

Resolution quality in the plane z = 0 of the circular cylinder. (a) GWMLES log SST. (b) GWMLES log γ R e θ t.

FIG. 32.

Resolution quality in the plane z = 0 of the circular cylinder. (a) GWMLES log SST. (b) GWMLES log γ R e θ t.

Close modal
TABLE XIII.

Simulation parameters of the circular cylinder for a characteristic length scale Lc = D.

Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLES log γ R e θ t  10 3  35  0.6  14.6 days 
DDES γ R e θ t  10 3  35  1.2  29.2 days 
Model dt ̂ L ̂ Ω N iter t iter (s) N flow tr N flow dat t sim
GWMLES log γ R e θ t  10 3  35  0.6  14.6 days 
DDES γ R e θ t  10 3  35  1.2  29.2 days 
#include “udf.h” 
#define c_sub 500.0 
#define c_log 2.0 
#define n1 8.0 
#define n2 16.0 
/
C_K: turbulent kinetic energy (m ⁁ 2/s ⁁ 2) 
C_O: turbulent specific dissipation rate (1/s) 
C_WALL_DIST: wall distance (m) 
C_MU_L: molecular viscosity (Pa*s) 
C_R: density (kg/m ⁁ 3) 
*/ 
DEFINE_SBES_BF(GWMLES_log, c, t) 
{ 
  double arg1, arg2, f; 
  arg1=c_log*sqrt(C_K(c,t))/(0.09*C_O(c,t)*C_WALL_DIST(c,t)); 
  arg2=c_sub*C_MU_L(c,t)/(pow(C_WALL_DIST(c,t),2)*C_O(c,t)*C_R(c,t)); 
  if (arg1>arg2) { 
   f = tanh(pow(arg1,n1)); 
  } 
  else { 
   f = tanh(pow(arg2,n1)); 
  } 
   return f; 
} 
DEFINE_SBES_BF(GWMLES_sub, c, t) 
  { 
  double arg2, f; 
  arg2=c_sub*C_MU_L(c,t)/(pow(C_WALL_DIST(c,t),2)*C_O(c,t)*C_R(c,t)); 
  bf = tanh(pow(arg2,n2)); 
  return f; 
  } 
#include “udf.h” 
#define c_sub 500.0 
#define c_log 2.0 
#define n1 8.0 
#define n2 16.0 
/
C_K: turbulent kinetic energy (m ⁁ 2/s ⁁ 2) 
C_O: turbulent specific dissipation rate (1/s) 
C_WALL_DIST: wall distance (m) 
C_MU_L: molecular viscosity (Pa*s) 
C_R: density (kg/m ⁁ 3) 
*/ 
DEFINE_SBES_BF(GWMLES_log, c, t) 
{ 
  double arg1, arg2, f; 
  arg1=c_log*sqrt(C_K(c,t))/(0.09*C_O(c,t)*C_WALL_DIST(c,t)); 
  arg2=c_sub*C_MU_L(c,t)/(pow(C_WALL_DIST(c,t),2)*C_O(c,t)*C_R(c,t)); 
  if (arg1>arg2) { 
   f = tanh(pow(arg1,n1)); 
  } 
  else { 
   f = tanh(pow(arg2,n1)); 
  } 
   return f; 
} 
DEFINE_SBES_BF(GWMLES_sub, c, t) 
  { 
  double arg2, f; 
  arg2=c_sub*C_MU_L(c,t)/(pow(C_WALL_DIST(c,t),2)*C_O(c,t)*C_R(c,t)); 
  bf = tanh(pow(arg2,n2)); 
  return f; 
  } 

Even if the results given by GWMLES log γ R e θ t have physical sense, the generation of turbulent structures for Φ < 90 ° is not consistent with the prediction of the transition point given by the correlation implemented in the γ R e θ model. In fact, Fig. 30(b) shows that μ t > 0 (which means γ > 0) for Φ < 90 °, so that the correlation predicts laminar flow for Φ < 90 °, which is in line with the laminar separation claimed by the authors of the experiment taken as Ref. 79. With the numerical method used in this work, a flow instability is produced that leads to the natural laminar–turbulent transition. This can be a particular feature of the flow, but it cannot be discarded that it is a consequence of the instability of the numerical scheme. Indeed, hybrid discretizations using a blended formulation between the central differencing and first-order upwind schemes (see Ref. 56) with low blending coefficients (up to 0.25 for the first-order upwind) have been tried, leading to the conclusion that the transition point is a function of the numerical diffusion of the scheme. This is the same conclusion extracted from a dedicated study on this phenomenon, in which it is shown that, in a similar solver based on the finite volume method, the central differencing scheme generates nonphysical instabilities that lead to early natural laminar–turbulent transition.82 In addition, taking as reference DNS data, the article cited shows that, for under-resolved simulations, the transition point can be correctly predicted using a hybrid discretization, in particular, tuning the blending constant to a value that depends on the mesh, which is far from optimal. The instability of the central differencing scheme is well known for a long time.83,84 These instabilities are also expected to be produced in turbulent flows, although in these cases, the flow instability is much intense than the instability generated by the numerical scheme, so that using a central difference scheme is not detrimental for the quality of the results in these cases. In contrast, in flows with laminar–turbulent transition, the numerical instability can dominate against the physical instability, leading to the failure of the prediction of the transition point.

The discrepancy between simulations and the experiment can also be explained from the uncertainties of the experiment. Apart from turbulence sources such as surface imperfections or measurement devices, a key aspect in laminar–turbulent transition is the freestream turbulent intensity. The work cited79 says that I 0.6 %, while in the numerical simulations, I = 10 8 %. From classical correlations that relate the transition momentum thickness Reynolds number with the dimensionless pressure gradient λ θ and the freestream turbulence intensity, R e θ t = f ( λ θ , I ), it is known the high sensitivity of the correlation with respect to the freestream turbulence intensity [see Fig. 33(a)]. Thus, a range of 0 I 0.6 % is restrictive for comparison between numerical and experimental data. Indeed, as shown in Fig. 33(b), there is a wide spread in the values obtained from different experiments,86–93 which means that the flow is highly sensitive to the experimental setup.

FIG. 33.

Experimental data on laminar–turbulent transition. (a) Correlation of R e θ t = f ( λ θ , I ).85 (b) Drag coefficient of the circular cylinder.

FIG. 33.

Experimental data on laminar–turbulent transition. (a) Correlation of R e θ t = f ( λ θ , I ).85 (b) Drag coefficient of the circular cylinder.

Close modal

In this work, a generalized wall-modeled large simulation model (GWMLES) has been developed. The model improves and expands upon the classical WMLES approach, which models turbulence up to the inner part of the log-layer, by introducing a formulation that also allows for the modeling of the entire log-layer. This is particularly relevant for the industry given the current needs and available computing power. The model is based on the main idea of the unpublished SBES model, which blends RANS and LES turbulent viscosities. The blending function has been designed to be independent of the mesh, thereby avoiding the grid-induced separation problem that is one of the main drawbacks of state-of-the-art models like the DES family. Furthermore, GWMLES has been shown to require approximately 30% lower computational cost than other hybrid RANS/LES models such as DDES in the turbulent flows considered.

The numerical methodology has been validated using experimental data and DNS, WMLES, and LES simulations available in the literature, focusing on four test cases that represent industrial flows. Simulating a free shear flow has led to the conclusion that the model is consistent, since it provides a LES solution in the absence of boundary layers. The simulation of a periodic channel flow demonstrates the excellent performance of the model in a standard WMLES test case. Additionally, the advantages of resolving the inner part of the logarithmic layer in RANS instead of using wall functions or a mixing-length model have been presented.

More complex test cases, such as the flow around the Ahmed body, which involves turbulent flow with separation and subsequent reattachment, have been studied. GWMLES has shown its ability to capture finer turbulence structures compared to other hybrid RANS/LES models. This is attributed to the use of a more consistent LES model and a faster transition from modeled-to-resolved turbulence provided by the formulation. The results obtained by the model are in excellent agreement with experimental data, showing accuracy similar to LES simulations, but at a significantly lower computational cost. In addition, it has been shown to provide superior performance than URANS, even for coarse meshes. GWMLES is also promising as an alternative to LES in laminar–turbulent transitional flows, fully resolving laminar regions, and being able to solve the natural transition from laminar to fully turbulent flow.

The test cases considered have showcased the model's capabilities in various flow scenarios. However, further validation of the model is necessary, with additional discussion about time scales, in particular in different flow types that will be the basis of future work; first, additional standard test cases of turbulent flows such as the backward facing step61 allow additional comparison with experimental data and LES; second, complex flows with real geometries used in the aeronautical industry94 serve as additional challenging cases using spectral analysis as basis for the result comparison; third, the instabilities generated by numerical schemes can be studied in detail in laminar–turbulent transitional flows with laminar separation bubbles.41,95

The author acknowledge Universidad Politécnica de Madrid for providing computing resources on Magerit Supercomputer. Gonzalo Rubio also acknowledges the funding received by the Grant DeepCFD (Project No. PID2022-137899OB-I00) funded by No. MCIN/AEI/10.13039/501100011033 and by ERDF A way of making Europe.

The authors have no conflicts to disclose.

Aitor Amatriain: Conceptualization (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal). Corrado Gargiulo: Funding acquisition (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Writing – review & editing (equal). Gonzalo Rubio: Conceptualization (supporting); Investigation (supporting); Methodology (supporting); Supervision (lead); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

The f GWMLES blending function defined in Eq. (7) with the parameters given in Table I is implemented in Ansys Fluent with a code written in C programming language. This code is compiled with the in-built compiler of Ansys Fluent (Fluent—user-defined functions (UDF)—compiled), so that a user-defined blending function is applied to the SBES model of Ansys Fluent (models—turbulence—k–omega SST—SBES with user-defined function). The expressions for the blending functions contain cell macros, whose definition is given after the parameter definition of the code given below. In the same way as the cell macros, the blending function is defined for all cells (input argument “c”) of every domain thread (input argument “t”).

1.
H.
Choi
and
P.
Moin
, “
Grid-point requirements for large eddy simulation: Chapman's estimates revisited
,”
Phys. Fluids
24
,
011702
(
2012
).
2.
S. B.
Pope
,
Turbulent Flows
(
Cambridge University Press
,
2000
).
3.
S. T.
Bose
and
G. I.
Park
, “
Wall-modeled large-eddy simulation for complex turbulent flows
,”
Annu. Rev. Fluid Mech.
50
(
1
),
535
561
(
2018
).
4.
U.
Piomelli
,
E.
Balaras
,
H.
Pasinato
,
K. D.
Squires
, and
P. R.
Spalart
, “
The inner-outer layer interface in large-eddy simulations with wall-layer models
,”
Int. J. Heat Fluid Flow
24
(
4
),
538
550
(
2003
).
5.
A.
Keating
and
U.
Piomelli
, “
A dynamic stochastic forcing method as a wall-layer model for large-eddy simulation
,”
J. Turbul.
7
,
N12
(
2006
).
6.
M. L.
Shur
,
P. R.
Spalart
,
M. K.
Strelets
, and
A. K.
Travin
, “
A hybrid RANS-LES approach with delayed-DES and wall-modelled LES capabilities
,”
Int. J. Heat Fluid Flow
29
(
6
),
1638
1649
(
2008
).
7.
D. A.
Yoder
,
J. R.
DeBonis
, and
N. J.
Georgiadis
, “
Modeling of turbulent free shear flows
,”
Comput. Fluids
117
,
212
232
(
2015
).
8.
D.
Corson
,
R.
Jaiman
, and
F.
Shakib
, “
Industrial application of RANS modelling: Capabilities and needs
,”
Int. J. Comput. Fluid Dyn.
23
(
4
),
337
347
(
2009
).
9.
A.
Uzun
,
A. S.
Lyrintzis
, and
G. A.
Blaisdell
, “
Coupling of integral acoustics methods with LES for jet noise prediction
,”
Int. J. Aeroacoust.
3
(
4
),
297
346
(
2004
).
10.
M.
Breuer
,
G.
De Nayer
,
M.
Münsch
,
T.
Gallinger
, and
R.
Wüchner
, “
Fluid-structure interaction using a partitioned semi-implicit predictor-corrector coupling scheme for the application of large-eddy simulation
,”
J. Fluids Struct.
29
,
107
130
(
2012
).
11.
R. O.
Fox
, “
Large-eddy-simulation tools for multiphase flows
,”
Annu. Rev. Fluid Mech.
44
(
1
),
47
76
(
2012
).
12.
E.
Palkin
,
R.
Mullyadzhanov
,
M.
Hadžiabdić
, and
K.
Hanjalić
, “
Scrutinizing URANS in shedding flows: The case of cylinder in cross-flow in the subcritical regime
,”
Flow Turbul. Combust.
97
(
4
),
1017
1046
(
2016
).
13.
M.
Strelets
, “
Detached eddy simulation of massively separated flows
,” AIAA Paper No. 2001-879,
2001
.
14.
F. R.
Menter
, “
Stress-blended eddy simulation (SBES)—A new paradigm in hybrid RANS-LES modeling
,” in
Progress in Hybrid RANS-LES Modelling
(Springer,
2018
), pp.
27
37
.
15.
P. R.
Spalart
,
S.
Deck
,
M. L.
Shur
,
K. D.
Squires
,
M. K.
Strelets
, and
A.
Travin
, “
A new version of detached-eddy simulation, resistant to ambiguous grid densities
,”
Theor. Comput. Fluid Dyn.
20
(
181
),
181
(
2006
).
16.
O. M. F.
Browne
,
J. A.
Housman
,
G. K.
Kenway
,
A. S.
Ghate
, and
C. C.
Kiris
, “
Numerical investigation of c L , max prediction on the NASA high-lift common research model
,”
AIAA J.
61
(
4
),
1639
1658
(
2023
).
17.
M. S.
Gritskevich
,
A. V.
Garbaruk
,
J.
Schütze
, and
F. R.
Menter
, “
Development of DDES and IDDES formulations for the k-ω shear stress transport model
,”
Flow Turbul. Combust.
88
(
3
),
431
(
2012
).
18.
J.
Fröhlich
and
D.
von Terzi
, “
Hybrid LES/RANS methods for the simulation of turbulent flows
,”
Prog. Aerosp. Sci.
44
(
5
),
349
377
(
2008
).
19.
P. R.
Spalart
, “
Detached-eddy simulation
,”
Annu. Rev. Fluid Mech.
41
(
1
),
181
202
(
2009
).
20.
F. R.
Menter
and
M.
Kuntz
, “
Adaptation of eddy-viscosity turbulence models to unsteady separated flow behind vehicles
,” in
The Aerodynamics of Heavy Vehicles: Trucks, Buses, and Trains
(
Springer
,
2004
), pp.
339
352
.
21.
S.
Deck
and
N.
Renard
, “
Towards an enhanced protection of attached boundary layers in hybrid RANS/LES methods
,”
J. Comput. Phys.
400
,
108970
(
2020
).
22.
D.
Schwamborn
and
M.
Strelets
, “
ATAAC—An EU-project dedicated to hybrid RANS/LES methods
,” in
Progress in Hybrid RANS-LES Modelling
(Springer,
2012
), pp.
59
75
.
23.
Y.
Han
,
Y.
He
, and
J.
Le
, “
Modification to improved delayed detached-eddy simulation regarding the log-layer mismatch
,”
AIAA J.
58
(
2
),
712
721
(
2020
).
24.
F. R.
Menter
, “
Two-equation eddy-viscosity turbulence models for engineering applications
,”
AIAA J.
32
(
8
),
1598
1605
(
1994
).
25.
A.
Yoshizawa
, “
Statistical theory for compressible turbulent shear flows, with the application to subgrid modeling
,”
Phys. Fluids
29
(
7
),
2152
2164
(
1986
).
26.
F. R.
Menter
and
Y.
Egorov
, “
The scale-adaptive simulation method for unsteady turbulent flow predictions. I. Theory and model description
,”
Flow Turbul. Combust.
85
(
1
),
113
138
(
2010
).
27.
Y.
Egorov
,
F. R.
Menter
,
R.
Lechner
, and
D.
Cokljat
, “
The scale-adaptive simulation method for unsteady turbulent flow predictions. II. Application to complex flows
,”
Flow Turbul. Combust.
85
(
1
),
139
165
(
2010
).
28.
C. Y.
Xu
,
T.
Zhang
,
Y. Y.
Yu
, and
J. H.
Sun
, “
Effect of von karman length scale in scale adaptive simulation approach on the prediction of supersonic turbulent flow
,”
Aerosp. Sci. Technol.
86
,
630
639
(
2019
).
29.
A.
Rezaeiha
,
H.
Montazeri
, and
B.
Blocken
, “
CFD analysis of dynamic stall on vertical axis wind turbines using scale-adaptive simulation (SAS): Comparison against URANS and hybrid RANS/LES
,”
Energy Convers. Manage.
196
,
1282
1298
(
2019
).
30.
L.
Davidson
, “
Evaluation of the SST-SAS model: Channel flow, asymmetric diffuser and axi-symmetric hill
,” in Proceedings of the
European Conference on Computational Fluid Dynamics
(Springer,
2006
), p.
20
.
31.
S. S.
Girimaji
, “
Partially-averaged Navier-Stokes model for turbulence: A Reynolds-averaged Navier-Stokes to direct numerical simulation bridging method
,”
J. Appl. Mech.
73
(
3
),
413
421
(
2006
).
32.
M.
Terracol
, “
A zonal RANS/LES approach for noise sources prediction
,”
Flow Turbul. Combust.
77
(
1
),
161
184
(
2006
).
33.
P.
Ekman
,
D.
Wieser
,
T.
Virdung
, and
M.
Karlsson
, “
Assessment of hybrid RANS-LES methods for accurate automotive aerodynamic simulations
,”
J. Wind Eng. Ind. Aerodyn.
206
,
104301
(
2020
).
34.
X.
Wang
,
J.
Zhang
,
Y.
Chen
, and
Z.
Kuai
, “
Numerical study of rising Taylor bubbles driven by buoyancy and additional pressure
,”
Int. J. Multiphase Flow
159
,
104309
(
2023
).
35.
R. B.
Langtry
and
F. R.
Menter
, “
Correlation-based transition modeling for unstructured parallelized computational fluid dynamics codes
,”
AIAA J.
47
(
12
),
2894
2906
(
2009
).
36.
R. B.
Langtry
,
F. R.
Menter
,
S. R.
Likki
,
Y. B.
Suzen
,
P. G.
Huang
, and
S.
Völker
, “
A correlation-based transition model using local variables. II. Test cases and industrial applications
,”
J. Turbomach.
128
,
423
434
(
2006
).
37.
H.
Dong
,
T.
Xia
,
L.
Chen
,
S.
Liu
,
Y. D.
Cui
,
B. C.
Khoo
, and
A.
Zhao
, “
Study on flow separation and transition of the airfoil in low Reynolds number
,”
Phys. Fluids
31
(
10
),
103601
(
2019
).
38.
A.
Rezaeiha
,
H.
Montazeri
, and
B.
Blocken
, “
On the accuracy of turbulence models for CFD simulations of vertical axis wind turbines
,”
Energy
180
,
838
857
(
2019
).
39.
M. S.
Campobasso
,
A.
Castorrini
,
L.
Cappugi
, and
A.
Bonfiglioli
, “
Experimentally validated three-dimensional computational aerodynamics of wind turbine blade sections featuring leading edge erosion cavities
,”
Wind Energy
25
(
1
),
168
189
(
2022
).
40.
S.
Lardeau
,
M.
Leschziner
, and
T.
Zaki
, “
Large eddy simulation of transitional separated flow over a flat plate and a compressor blade
,”
Flow Turbul. Combust.
88
,
19
44
(
2012
).
41.
O.
Lehmkuhl
,
I.
Rodríguez
,
A.
Baez
,
A.
Oliva
, and
C. D.
Pérez-Segarra
, “
On the large-eddy simulations for the flow around aerodynamic profiles using unstructured grids
,”
Comput. Fluids
84
,
176
189
(
2013
).
42.
J.
Bodart
and
J.
Larsson
, “
Sensor-based computation of transitional flows using wall-modeled large eddy simulation
,” in
Center for Turbulence Research Annual Briefs
(Stanford University,
2012
), pp.
229
240
.
43.
M.
Bouchard
,
J.
Marty
,
S.
Deck
, and
M.
Costes
, “
Numerical investigation of self-sustained oscillations of stall cells around a leading edge-separating airfoil
,”
Phys. Fluids
34
(
11
),
115153
(
2022
).
44.
B.
Chaouat
, “
The state of the art of hybrid RANS/LES modeling for the simulation of turbulent flows
,”
Flow Turbul. Combust.
99
(
2
),
279
327
(
2017
).
45.
G.
Pont
,
P.
Brenner
,
P.
Cinnella
, and
J. C.
Robinet
, “
High-order hybrid RANS/LES strategy for industrial applications
,” in
Direct and Large-Eddy Simulation X
(Springer,
2018
), pp.
313
319
.
46.
F.
Nicoud
and
F.
Ducros
, “
Subgrid-scale stress modelling based on the square of the velocity gradient tensor flow
,”
Flow Turbul. Combust.
62
(
3
),
183
200
(
1999
).
47.
W. M.
Kays
, “
Turbulent Prandtl number—Where are we?
,”
J. Heat Transfer
116
(
2
),
284
295
(
1994
).
48.
D. C.
Wilcox
,
Turbulence Modeling for CFD
, 3rd ed. (
DCW Industries
,
2006
).
49.
F. R.
Menter
, “
Improved two-equation k ω turbulence models for aerodynamic flows
,”
Technical Report No. NASA-TM-103975
(
NASA
,
1992
).
50.
Ansys, Inc
.,
Ansys Fluent Theory Guide
(
Ansys, Inc
.,
2023
), p.
2023R1
.
51.
G. A.
Bres
,
S. T.
Bose
,
M.
Emory
,
F. E.
Ham
,
O. T.
Schmidt
,
G.
Rigas
, and
T.
Colonius
, “
Large-eddy simulations of co-annular turbulent jet using a Voronoi-based mesh generation framework
,” AIAA Paper No. 2018-3302,
2018
.
52.
A.
Lozano-Durán
,
S. T.
Bose
, and
P.
Moin
, “
Performance of wall-modeled LES with boundary-layer-conforming grids for external aerodynamics
,”
AIAA J.
60
(
2
),
747
766
(
2022
).
53.
S.
Nishizawa
,
H.
Yashiro
,
Y.
Sato
,
Y.
Miyamoto
, and
H.
Tomita
, “
Influence of grid aspect ratio on planetary boundary layer turbulence in large-eddy simulations
,”
Geosci. Model Dev.
8
(
10
),
3393
3419
(
2015
).
54.
D. K.
Kolmogorov
,
F.
Menter
, and
A. V.
Garbaruk
, “
On mesh requirements for large eddy simulation with wall functions
,”
J. Phys.: Conf. Ser.
2103
,
012212
(
2021
).
55.
K.
Abe
, “
Detailed investigation of subgrid scale models in large-eddy simulation using high aspect-ratio grid spacing
,”
Phys. Fluids
33
,
115120
(
2021
).
56.
J. H.
Ferzinger
,
M.
Peric
, and
R. L.
Street
,
Computational Methods for Fluid Dynamics
, 4th ed. (
Springer
,
2020
).
57.
S.
Majumdar
, “
Role of underrelaxation in momentum interpolation for calculation of flow with nonstaggered grids
,”
Numer. Heat Transfer
13
(
1
),
125
132
(
1988
).
58.
S. F.
McCormick
,
Multigrid Methods
(
SIAM
,
1987
).
59.
S.
Menon
,
P. K.
Yeung
, and
W. W.
Kim
, “
Effect of subgrid models on the computed interscale energy transfer in isotropic turbulence
,”
Comput. Fluids
25
(
2
),
165
180
(
1996
).
60.
P. P.
Sullivan
,
J. C.
McWilliams
, and
C. H.
Moeng
, “
A subgrid-scale model for large-eddy simulation of planetary boundary-layer flows
,”
Boundary Layer Meteorol.
71
,
247
276
(
1994
).
61.
L.
Davidson
, “
How to estimate the resolution of an LES of recirculating flow
,” in
Quality and Reliability of Large-Eddy Simulations II
(
Springer
,
2011
), pp.
269
286
.
62.
T.
Von Kármán
, “
Progress in the statistical theory of turbulence
,”
Proc. Natl. Acad. Sci. U.S.A.
34
(
11
),
530
539
(
1948
).
63.
J.
Jeong
and
F.
Hussain
, “
On the identification of a vortex
,”
J. Fluid Mech.
285
,
69
94
(
1995
).
64.
K.
Viswanathan
, “
Aeroacoustics of hot jets
,”
J. Fluid Mech.
516
,
39
82
(
2004
).
65.
J. E.
Bridges
and
M. P.
Wernet
, “
Noise of internally mixed exhaust systems with external plug for supersonic transport applications
,” AIAA Paper No. 2021-2218,
2021
.
66.
J.
Bridges
and
M. P.
Wernet
, “
The NASA subsonic jet particle image velocimetry (PIV) dataset
,”
Technical Report No. TM-2011-216807
(
NASA
,
2011
).
67.
G. D.
Stich
,
A. S.
Ghate
,
J. A.
Housman
, and
C. C.
Kiris
, “
Wall modeled large eddy simulations for NASA's jet noise consensus database of single-stream, round, convergent jets
,” AIAA Paper No. 2022-0684,
2022
.
68.
S.
Hoyas
and
J.
Jiménez
, “
Scaling of the velocity fluctuations in turbulent channels up to R e τ = 2003
,”
Phys. Fluids
18
(
1
),
011702
(
2006
).
69.
M.
Bernardini
,
S.
Pirozzoli
, and
P.
Orlandi
, “
Velocity statistics in turbulent channel flow up to
R e τ = 4000,”
J. Fluid Mech.
742
,
171
191
(
2014
).
70.
G.
Lodato
,
P.
Castonguay
, and
A.
Jameson
, “
Structural wall-modeled LES using a high-order spectral difference scheme for unstructured meshes
,”
Flow Turbul. Combust.
92
,
579
606
(
2014
).
71.
A.
Lozano-Durán
and
J.
Jiménez
, “
Effect of the computational domain on direct simulations of turbulent channels up to R e τ = 4200
,”
Phys. Fluids
26
,
011702
(
2014
).
72.
S. R.
Ahmed
,
G.
Ramm
, and
G.
Faltin
, “
Some salient features of the time-averaged ground vehicle wake
,”
SAE Trans.
93
,
473
503
(
1984
).
73.
H.
Lienhart
,
C.
Stoots
, and
S.
Becker
, “
Flow and turbulence structures in the wake of a simplified car model (Ahmed model)
,” in
New Results in Numerical and Experimental Fluid Mechanics III
(
Springer
,
2002
), pp.
323
330
.
74.
B.
Conan
,
J.
Anthoine
, and
P.
Planquart
, “
Experimental aerodynamic study of a car-type bluff body
,”
Exp. Fluids
50
,
1273
1284
(
2011
).
75.
G.
Rossitto
,
C.
Sicot
,
V.
Ferrand
,
J.
Borée
, and
F.
Harambat
, “
Influence of afterbody rounding on the pressure distribution over a fastback vehicle
,”
Exp. Fluids
57
(
43
),
43
(
2016
).
76.
O.
Lehmkuhl
,
G.
Houzeaux
,
H.
Owen
,
G.
Chrysokentis
, and
I.
Rodriguez
, “
A low-dissipation finite element scheme for scale resolving simulations of turbulent flows
,”
J. Comput. Phys.
390
,
51
65
(
2019
).
77.
D. K.
Kolmogorov
,
A.
Hüppe
,
F.
Menter
, and
A. V.
Garbaruk
, “
Large eddy simulation with wall functions of Ahmed body
,”
J. Phys.: Conf. Ser.
2103
,
012213
(
2021
).
78.
H.
Schlichting
,
Boundary Layer Theory
, 7th ed. (
McGraw-Hill
,
1979
).
79.
B.
Cantwell
and
D.
Coles
, “
An experimental study of entrainment and transport in the turbulent near wake of a circular cylinder
,”
J. Fluid Mech.
136
,
321
374
(
1983
).
80.
M.
Breuer
, “
A challenging test case for large eddy simulation: High Reynolds number circular cylinder flow
,”
Int. J. Heat Fluid Flow
21
(
5
),
648
654
(
2000
).
81.
M.
de la Llave Plata
,
E.
Lamballais
, and
F.
Naddei
, “
On the performance of a high-order multiscale DG approach to LES at increasing Reynolds number
,”
Comput. Fluids
194
,
104306
(
2019
).
82.
T. A.
Smith
and
Y.
Ventikos
, “
Boundary layer transition over a foil using direct numerical simulation and large eddy simulation
,”
Phys. Fluids
31
(
12
),
124102
(
2019
).
83.
R. C.
Swanson
and
E.
Turkel
, “
On central-difference and upwind schemes
,”
J. Comput. Phys.
101
(
2
),
292
306
(
1992
).
84.
W.
Shyy
,
M. H.
Chen
,
R.
Mittal
, and
H. S.
Udaykumar
, “
On the suppression of numerical oscillations using a non-linear filter
,”
J. Comput. Phys.
102
(
1
),
49
62
(
1992
).
85.
B. J.
Abu-Ghannam
and
R.
Shaw
, “
Natural transition of boundary layers-the effects of turbulence, pressure gradient, and flow history
,”
J. Mech. Eng. Sci.
22
(
5
),
213
228
(
1980
).
86.
N. K.
Delany
and
N. E.
Sorensen
, “
Low-speed drag of cylinders of various shapes
,”
Technical Report No. TN-3038
(
NACA
,
1953
).
87.
R. E.
Spitzer
, “
Measurements of unsteady pressures and wake fluctuations for flow over a cylinder at supercritical Reynolds number,” Ph.D. thesis
(
California Institute of Technology
,
1965
).
88.
E.
Achenbach
, “
Distribution of local pressure and skin friction around a circular cylinder in cross-flow up to
R e = 5 × 10 6,”
J. Fluid Mech.
34
(
4
),
625
639
(
1968
).
89.
W. J.
Bursnall
and
L. K.
Loftin
, “
Experimental investigation of the pressure distribution about a yawed circular cylinder in the critical Reynolds number range
,”
Technical Report No. TN-2463
(
NACA
,
1951
).
90.
G.
Vaz
,
C.
Mabilat
,
R.
van der Wal
, and
P.
Gallagher
, “
Viscous flow computations on a smooth cylinders: A detailed numerical study with validation
,” in
Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering
(ASME,
2007
), pp.
849
860
.
91.
G.
Schewe
, “
On the force fluctuations acting on a circular cylinder in crossflow from subcritical up to transcritical Reynolds numbers
,”
J. Fluid Mech.
133
,
265
285
(
1983
).
92.
C.
Wieselsberger
, “
New data on the laws of fluid resistance
,”
Technical Report No. TN-84
(
NACA
,
1922
).
93.
A.
Fage
, “
Drag of circular cylinders and spheres
,”
Technical Report No. R&M 1370
(
Aeronautical Research Council
,
1930
).
94.
M.
Terracol
and
E.
Manoha
, “
Wall-resolved large-eddy simulation of a three-element high-lift airfoil
,”
AIAA J.
58
(
2
),
517
536
(
2020
).
95.
J.
Slaughter
,
D.
Moxey
, and
S.
Sherwin
, “
Large eddy simulation of an inverted multi-element wing in ground effect
,”
Flow Turbul. Combust.
110
(
4
),
917
944
(
2023
).