Open access peer-reviewed chapter - ONLINE FIRST

A Numerical Study of the Minimum-Dissipation Model for Large-Eddy Simulation in OpenFOAM

Written By

Jing Sun and Roel Verstappen

Submitted: 13 February 2024 Reviewed: 08 May 2024 Published: 22 July 2024

DOI: 10.5772/intechopen.1005638

Computational Fluid Dynamics - Analysis, Simulations, and Applications IntechOpen
Computational Fluid Dynamics - Analysis, Simulations, and Applica... Edited by Mahboub Baccouch

From the Edited Volume

Computational Fluid Dynamics - Analysis, Simulations, and Applications [Working Title]

Prof. Mahboub Baccouch

Chapter metrics overview

15 Chapter Downloads

View Full Metrics

Abstract

The minimum-dissipation model (QR) has been utilized in studying turbulent channel flows at Reynolds numbers Reτ of up to 2000, as well as in investigating the flow past a circular cylinder at a Reynolds number (Re) of 3900, and flow over periodic hills at Re = 10,595. In our investigations, we have employed both symmetry-preserving discretizations and standard second-order accurate discretization methods within the OpenFOAM framework. The outcomes are compared with Direct Numerical Simulation (DNS) and experiment, indicating a favorable alignment between the QR results and the reference data. The findings suggest that the static QR model achieves comparable performance to dynamic models while cutting computational costs by a factor of three. The model coefficient C = 0.024 produces the most precise predictions, and as the mesh resolution increases, the influence of the subgrid model decreases, dropping to less than 20% of the molecular viscosity at the finest mesh. Furthermore, the QR model can predict the mean and root-mean-square velocity accurately up to Reτ=2000 without a wall damping function. The characteristics of turbulence strongly rely on spatial discretization methods. Various comparisons demonstrate the QR model conjugated with symmetry-preserving discretization performs better than the standard OpenFOAM discretization. Within the realm of OpenFOAM discretization, central difference schemes outperform other approaches.

Keywords

  • minimum-dissipation model
  • turbulence
  • large-eddy simulation
  • symmetry-preserving discretization
  • OpenFOAM

1. Introduction

Turbulent flows are prevalent in numerous engineering applications, yet simulating them using Direct Numerical Simulation (DNS) can be prohibitively expensive, especially for high-Reynolds-number flows. The Unsteady Reynolds-Averaged Navier-Stokes (RANS) method, while reducing computational costs, introduces significant simplifications, leading to decreased accuracy due to its treatment of large-scale unsteadiness. To tackle these challenges, large-eddy simulation (LES) has been developed. LES directly resolves the larger-scale, unsteady turbulent motions while modeling the effects of smaller-scale motions. This is achieved by representing the unresolved scale of motion using a subgrid model.

Among the subgrid models available for LES, the Smagorinsky model stands out as the most widely used one [1]. While the Smagorinsky model demonstrates satisfactory performance in decaying homogeneous turbulence simulations [2], it tends to dissipate eddies inappropriately for laminar and transitional flows. The performance of the Smagorinsky model can be enhanced by dynamically computing the model coefficient, but this approach comes with increased computational costs. Another alternative is the WALE model [3], which addresses behavior near walls by utilizing the square of the velocity gradient tensor. Additionally, there is the Vreman model, which remains insensitive to pure shear but may lead to eddy dissipation in cases of back-scatter and solid body rotation [4].

Minimum-dissipation models offer a straightforward alternative to Smagorinsky-type approaches for parameterizing subfilter turbulent fluxes in large-eddy simulation (LES). The QR model, proposed by Verstappen [5, 6], is the first minimum-dissipation model introduced in this context. The QR model possesses several desirable characteristics: it is more computationally efficient than dynamic models, effectively deactivates in laminar and transitional flows, and maintains consistency with the exact subfilter stress tensor on isotropic grids. Subsequently, the anisotropic minimum-dissipation model (AMD) was developed by Rozema et al. specifically for flows on anisotropic grids [7]. Abkar and Moin utilized the AMD model to examine high-Reynolds-number rough-wall boundary layer flow [8]. Zahiri et al. further extended the application of the AMD model by implementing it into OpenFOAM and conducting tests on both single-phase and multi-phase flows [9].

However, there have been limited investigations into the QR model, particularly within open-source software. In this study, we introduce the QR model into OpenFOAM and conduct simulations in scenarios involving high Reynolds numbers and complex geometries, marking the first instance of such exploration.

The discretization error discretizing the partial differential equations into algebraic equations. Verstappen and Veldman [10] put forward a proposition to precisely maintain the symmetry properties of the underlying differential operators. This involves representing the convective operator with a skew-symmetric matrix and the diffusive operator with a symmetric, positive-definite matrix. Trias et al. [11] extended this concept to unstructured collocated grids and introduced an approach based on fully conservative regularization of the convective term to address the issue of checkerboard spurious modes. Expanding upon these concepts, Komen et al. [12] devised a symmetry-preserving second-order time-accurate PISO-based method for solving incompressible flow equations on collocated grids. Hopman et al. made the code available to the public [13].

This chapter starts by introducing the QR model, the symmetry-preserving discretization, and OpenFOAM standard discretization methods. Then, simulations are conducted for channel flow, flow over periodic hills, and flow over a cylinder. The model coefficient is optimized and an investigation of the numerical discretization schemes is carried out. The obtained results are compared with the references [14, 15, 16, 17, 18, 19, 20, 21]. The chapter concludes with a summary.

Advertisement

2. Numerical schemes

2.1 Minimum-dissipation model

The dynamics of large eddies in incompressible fluid flow are governed by the following momentum and continuity equations

tυ+υυ+p2υSυ=τv,υ=0,E1

In this context, v represents the spatially filtered velocity, ν denotes the kinematic molecular viscosity, and p represents the pressure. Sυ=υ+υT/2 refers to the symmetric portion of the velocity gradient. The subgrid tensor τυ can be expressed as:

τij=13τkkδij+τij13τkkδij=23ksgsδij+τij13τkkδij.

where ksgs=12τkk is the subgrid kinetic energy. The subgrid stress tensor τij is split into an isotropic part 13τkkδij and anisotropic part τij13τkkδij. The isotropic part can be included into the pressure. The eddy-viscosity model describes the anisotropic part of the subgrid tensor as

τij13τkkδij=2νeSυE2

Note that the trace of Sυ is zero due to the divergence-free condition, υ=0. The coefficient νe is termed the eddy viscosity of the model. This subgrid model is time irreversible (for νe>0), meaning it contributes to dissipation forward in time. In the classical Smagorinsky eddy-viscosity model [1], the eddy viscosity is set equal to:

νe=CS2Δ24q.E3

Here qυ=12trS2υ represents the second invariant of the strain-rate tensor Sυ. It’s worth noting that Sυ=2trSυ2=4q. Various Smagorinsky coefficient values Cs have been suggested for different scenarios, spanning from Cs=0.10.17 for channel flow and temporal mixing layers, to Cs=0.200.22 for decaying homogeneous isotropic turbulence [2].

The first minimum-dissipation model, introduced by Verstappen [5], is based on the invariants of the strain-rate tensor, and is designed to deactivate in laminar flow and flows exhibiting negative eddy dissipation. This model operates under the assumption that the eddy-viscosity model should prevent the residual field, denoted as υ=υυ¯ where the bar represents the LES filter, from being dynamically significant. This criterion is formalized by constraining the subgrid kinetic energy using Poincaé’s inequality. According to Poincaré’s inequality, there exists a constant CΔ, contingent solely on ΩΔ, such that for every function υ in the Sobolev space W1,2ΩΔ

ΩΔυυ¯2dxCΔΩΔυ2dxE4

where the residual field υ=υυ¯ contains the eddies of size smaller than the length of the filter Δ, and = is the standard norm of the inner product on the space of real-valued L2ΩΔ functions, where ΩΔ is a domain with diameter Δj here it will be taken equal to a grid cell. The Poincaré constant CΔ, which remains unaffected by υ, is equal to the inverse of the smallest non-zero eigenvalue of the dissipative operator Δ==T on a grid cell ΩΔ [22]. =T is utilized for the L2ΩΔ inner product and periodic domain ΩΔ. For convex domains, the Poincaré constant is denoted as CΔ=Δ/π2 [23]. Poincaré’s inequality dictates that the kinetic energy of the residual field υ is bounded by a constant multiplied by the velocity gradient energy

ΩΔ12υ2dxCΔΩΔ12υ2dxE5

The equation governing the evolution of velocity gradient energy can be obtained by computing the inner product of L2 with 2υ. Integration by part yields

ddtΩΔ12υ2dx=νΩΔ2υ2dx+ΩΔυυΔυdxνeΩΔ2υ2dx.E6

The boundary terms arising from integration by parts vanish since ΩΔ is a periodic box. The second term on the right-hand side of Eq. (6) signifies the generation of velocity gradient energy by the convective term. It can be represented as rυ=trS3υ/3=detSυ, which corresponds to the third invariant of the strain-rate tensor Sυ (for further details, see Ref. [5, 6, 7]). We make the assumption that the molecular and eddy viscosity remains constant over a grid cell. The third term in Eq. (6) accounts for the dissipation caused by eddy viscosity and can be expressed as qω=trS2ω/2, where ω denotes the vorticity, ω=×υ.

Incorporating rυ and qω into the evolution equation of velocity gradient energy (Eq. (6))

ddtΩΔ12υ2dx=νΩΔ2υ2dx+4ΩΔrυdx4νeΩΔqωdx.E7

Let us consider that the eddy viscosity is chosen in such a way that the last two terms on the right-hand side of Eq. (7) counterbalance each other

ΩΔrυdx=νeΩΔqωdxE8

Then, we obtain

ddtΩΔ12υ2dx=νΩΔ2υ2dx.E9

Utilizing Poincaré’s inequality and Gronwall’s lemma on the right-hand side of the aforementioned equation yields:

ΩΔυ2xtdxCΔΩΔυ2xtdxCΔe2νtCΔΩΔυ2x0dx.

The energy of the subgrid-scale decays at least as fast as CΔe2νtCΔ, for any filter length Δ. Therefore, we are able to keep the subfilter component υ under control through the utilization of Eq. (8). The minimum amount of eddy dissipation must adhere to the dissipation condition (9).

The expression for qω in Eq. (8) can be equivalently written in terms of qv:

ΩΔqωdx=14ΩΔω2dx=14ΩΔωΔωdx=ΩΔωΔωdxΩΔωωdxΩΔqvdx.E10

Thus, Eq. (8) becomes

νe=ΩΔωωdxΩΔωΔωdxΩΔrυdxΩΔqυdxE11

The first fraction on the right-hand side above is at most CΔ, which corresponds to the reciprocal of the smallest eigenvalue of Δ on ΩΔ.

Therefore, we establish νeΩΔqυdx=CΔΩΔrυdx. This equation guarantees that the subgrid scales remain dynamically insignificant, implying that their energy is bounded by Eq. (9), where the subgrid scales energy ΩΔυ2xtdx decays at least as rapidly as CΔe2νtCΔ, for any filter length Δ. Consequently, the minimum eddy viscosity required to counteract the nonlinear production is determined by

νe=CΔrυ¯qυ¯.E12

To ensure non-negativity of the eddy viscosity, we employ the absolute value of rυ. The quantities qυ¯ and rυ¯ represent the grid cell averages of the second and third invariant of the strain-rate tensor, respectively. In practical applications, grid cell averaging is typically approximated using midpoint integration. This characterizes the QR model

τ13trτI=2νeSυ=2CΔrυqυSυE13

2.2 Numerical schemes in OpenFOAM

2.2.1 Spatial and temporal discretization

The discretization schemes typically maintain a second-order accuracy level. OpenFOAM employs a collocated arrangement where all variable values are computed and stored at the center xP of the control volume VP. These values are depicted by a piecewise constant profile, commonly known as the mean value

ϕxP=ϕP1VPVPϕxdV,E14

where ϕ denotes a discretized quantity. By employing the Divergence theorem (or Gauss theorem), the volume integrals present in the governing equations are transformed into surface integrals. Consequently, the problem shifts toward interpolating cell-centered values to face-centered values. The face values involved in the convective and diffusive fluxes must be determined through some form of interpolation from the centroid of the control volumes to their respective faces. The Gauss linear interpolation scheme in OpenFOAM is given by ϕf=αϕP+1αϕN, where α=fNPN, which yields a central difference scheme on a uniform mesh. The solver we use is pimpleFOAM if no other specification is provided.

If no specific instruction is provided, the continuous temporal derivative /t is discretized using an implicit backward scheme, as indicated by Eq. (15)

3ϕn+14ϕn+ϕn12Δttϕ,E15

where the n + 1 represents the value at the subsequent time level t+Δt, n denotes the value at the current time level, and n – 1 signifies the value at the preceding time level tΔt.

2.3 Symmetry-preserving discretization scheme

2.3.1 Navier-stokes equation

The governing equation for incompressible Navier-Stokes (NS) is expressed as

tu+uu1Reu+p=0,u=0.E16

Here, u stands for the velocity (no filter applied), p is the pressure, and Re represents the Reynolds number.

When there are no external sources, for instance, body or boundary forces, the change rate of the total energy remains unaffected by both pressure disparities and convective transport. Instead, it is solely dictated by dissipation. This fundamental physical characteristic can be easily inferred from the symmetric properties of the differential operators in the NS Eq. (16).

The total energy of the flow (u, u) is defined using the conventional scalar product. Its temporal evolution can be derived by differentiating (u, u) with respect to time and then reformulating tu with the assistance of Eq. (16). This process yields:

ddtuu=uuuuuu+1Reuu+uupuup.

By performing integration by parts on the linear and trilinear forms on the right-hand side, disregarding any boundary contributions, we

u=uand=E17

As a result of these (skew-) symmetries, the convective- and pressure-dependent terms cancel, leading to a simplification in the change rate of the total energy to

ddtuu=2Reuu0E18

In the discrete context, the evolution of energy also adheres to Eq. (18), where u is substituted with the discrete velocity, and is replaced by its discrete counterpart. This holds true provided that the discretization of the differential operator upholds the (skew-) symmetries outlined in Eq. (17). Under this condition, the energy of any discrete solution remains conserved in the absence of viscosity, and diminishes over time in the presence of dissipation. Essentially, a symmetry-preserving spatial discretization of the NS equation exhibits unconditional stability and conservativeness.

2.3.2 Symmetry-preserving discretization

Consider the discretization of a first-order derivative in one spatial dimension. The Lagrangian interpolation (minimizing local truncation error) violates the skew-symmetry of the convective operator on nonuniform grids. Consequently, quantities conserved in the continuous formulation, such as kinetic energy, are not preserved at the discrete level. The energy undergoes systematic damping; for instance, in upwind methods, where the convective term introduces artificial dissipation that dampens the kinetic energy. Alternatively, explicit damping may be required to maintain stability. However, artificial dissipation disrupts the delicate balance between convective transport and physical dissipation, particularly at the smallest scales of motion, thereby distorting the essence of turbulence. Hence, symmetry-preserving discretization is employed.

Consider the continuing equation and linear momentum in one spatial dimension

tu+u¯xuxxu/Re+xp=0,xu=0.E19

For simplicity, the convective transport velocity u¯ is assumed to be constant. In matrix-vector notation, the spatial discretization of Eq. (19) can be expressed as:

Ω0duhdt+C0u¯uh+D0uh+Ω0G0ph=0,M0uh=0E20

where the diagonal matrix Ω0 is built of the spacing of mesh: Ω0i,i=12xi+1xi1, the discrete velocities uituxit constitute the vector uh; the tridiagonal matrices C0u¯ and D0 represent the convective and diffusive operate, respectively.

The continuity equation xu=0 is discretized as ui+1ui1xi+1xi1=0. This yields M0uh=0.

The discretization of the pressure gradient is facilitated by utilizing the symmetry relation outlined in Eq. (17). According to Eq. (17), the continuous gradient operator equals the negative transpose of the divergence, implying that any velocity and pressure fields adhere to the relation pu=pu. This relation also applies to the discrete pressure ph and the discrete pressure gradient G0ph, expressed as

G0phΩ0uh=phG0Ω0uh=phM0uh,E21

if the gradient operator is approximated by

G0=Ω01M0.E22

Note that the gradient matrix, describing the integration of the pressure over the control volumes, is represented by M0. Since the discrete pressure gradient also inherits boundary conditions from the discrete divergence, there is no need to explicitly specify boundary conditions for the pressure.

In the absence of diffusion, that is for D0=0, the energy uh2=uhΩ0uh of any solution uh of the dynamic system of Eq. (20) is conserved if and only if the right-hand side of

ddtuh2=uhC0u¯+C0u¯uh+uhM0ph+M0phuh

is zero. This property remains valid (for any uh) if and only if the coefficient matrix C0u¯ exhibits skew-symmetry,

C0u¯+C0u¯=0,E23

In other words, the discrete operator must retain the skew-symmetry observed in the continuous convective derivative u.

To meet the skew-symmetry condition stated in Eq. (23), the interpolation weights of the neighboring discrete variables should be set to 1/2. Consequently, symmetry-preserving discretization yields

u¯xuxiu¯ui+1ui1xi+1xi1=Ω01C0u¯uhiE24

The elements of the tridiagonal matrix C0u¯ are defined as follows: C0u¯i,i1=12u¯, C0u¯i,i=0, and C0u¯i,i+1=12u¯. It has been rigorously proven that Eq. (24) yields second-order accurate solutions on both uniform and nonuniform grids [24].

Diffusion is discretized in the same vein. The resulting coefficient matrix D0 is positive-definite, like the underlying differential operator xx

D0=1ReΔ0Λ01Δ0,

where the difference matrix Δ0 is defined as Δ0uhi=uiui1, and the non-zero entries of the diagonal matrix Λ0 are given by Λ0i,i=δxi. Consequently, the symmetric component of C0u¯+D0 is solely influenced by diffusion, and hence is positive-definite. The energy of any solution u¯h of the semi-discrete system Eq. (20) evolves in a manner analogous to the continuous case; compare Eq. (18) to

ddtuhΩ0uh=20+23uhD0+D0uh0.

The right-hand side equals zero if and only if uh lies within the null space of D0+D0. Thus, in summary, as the energy uh2=uhΩ0uh remains constant over time, a stable solution can be achieved on any grid. Similarly, higher-order symmetry-preserving discretization can be derived in a comparable manner (refer to [10] for details). Taking all ingredients together yields the symmetry-preserving discretization of the NS equation.

2.3.3 Symmetry-preserving discretization of Navier-Stokes equation

the incompressible Navier-Stokes equations in their semi-discrete form are expressed as

Ωduhdt+Cuhuh+DuhMph=0,Muh=0E25

Global conservation laws involve integrals across the flow domain, which are transformed into scalar products upon discretization of the flow. For example, the change of the total mass of the flow is discretized as the scalar product of the constant vector 1 (with dimensions equal to the number of grid cells) and the discrete mass flux Muh. Due to this scalar product equals zero (Muh=0) the total mass is conserved.

To determine the total momentum, the scalar product of velocity vector uh with the vector Ω1 is taken, where the constant vector now encompasses as many entries as there are control volumes for the discrete velocity components ui,j and vi,j). The evolution of total momentum straightforwardly follows from Eq. (25)

ddt1Ωuh=1Cuh+Duh+1Mph=0.E26

Thus, momentum conservation is ensured if Cuh+D1=0, and the law of mass conservation is consistently discretized as M1=0. The former condition can be divided into two conditions: one for convective discretization, Cuh1=0, and one for diffusive discretization, D1=0. Moreover, we can omit the 's, yielding Cuh1=0 and D1=0, since the convective matrix Cuh is skew-symmetry and the diffusive matrix D is symmetric. Thus, it suffices to ensure that the constant vector lies in the null space of the approximate, convective, and diffusive operators.

The discretization is configured to ensure that the evolution of the (kinetic) energy uhΩuh of any solution to Eq. (25) adheres to

ddtuhΩuh=uhCuh+Cuh)uhuhD+Duh+uhMph+Mphuh,

where the right-hand side is negative for all uh's, except those within the null space of D+D. The convective component cancels out due to Cuh‘s skew-symmetry, while the pressure terms nullify on the staggered grids (thus, cannot destabilize the spatial discretization) because the discrete pressure gradient correlates with the transpose of M, as seen in Eq. 22.

In summary, energy is conserved for inviscid flow, while for viscous flow, the energy uh2=uhΩuh does not escalate in time. This indicates that the symmetry-preserving discretization Eq. 25 is stable and upholds conservation laws for mass, momentum, and energy.

2.3.4 Addressing the pressure-velocity coupling on the collocated grid

n a collocated grid setup the actual velocity uc is stored in the cell center. The coupling between velocity and pressure introduces an additional error term proportional to the third-order derivative of pressure p˜c into the momentum equation. This issue is commonly referred to as the checkerboard problem [25]. Trias et al. [11] proposed a method to eliminate the checkerboard spurious mode without introducing any nonphysical dissipation. The concept behind this approach involves utilizing a linear shift operator to transform a cell-centered velocity uc into a staggered one us and employing a fully conservative regularization of the convective term to restrain the generation of small motions.

The linear shift operator serves to establish a relationship between the staggered velocity field and the cell-centered counterparts and vice versa. In this context, the subscript c denotes variables centered on the collated mesh, while s indicates variables staggered on the faces. The cell-to-face linear shift operator, denoted as ΓcsRm×3n, converts a cell-centered velocity field into a staggered one

us=Γcsuc.E27

Conversely, the cell-centered fields are linked to the staggered ones through the linear shift transformation ΓscR3n×m,

uc=Γscus.E28

Note the general relation ΓscΓcs=I only holds approximately, meaning ucΓscΓcsuc.

The face-to-cell shift operator Γsc is constrained by Eq. (22) to the contribution of the pressure gradient term to the global kinetic energy. The expression for is as follows

Γsc=I3Ωc1ΓcsΩs,E29

where I3R3×3 is the identity matrix.

The linear shift operator is given by

Γcs=NsΠ,E30

where matrices ΠR3m×3n and NsRm×3m are defined as Π=I3Πcs and Ns=Ns,1Ns,2Ns,3, where ΠcsRm×n is the operator that interpolates a cell-centered scalar field to the faces, and Ns,iRm×m are diagonal matrices containing the xi-spatial components of the face-normal vectors.

A classical fractional step projection method [26, 27, 28] is employed to address the velocity-pressure coupling. The staggered velocity field, denoted as us, a velocity usp can be uniquely decomposed into a curl-free vector represented as the gradient of a scalar field, Gp˜c and a solenoidal vector, usn+1. This decomposition is formulated as

usp=usn+1+Gp˜c,E31

taking the divergence of Eq. (31) yields a discrete Poisson equation for p˜c,

Musp=Musn+1+MGp˜cMusp=MGp˜c.E32

The cell-center predicted velocity field, ucp, is computed with the projection method and then corrected to obtain the velocity at the next time step, ucn+1. Assuming ucpΓscΓcsucp is satisfied, the overall correction algorithm can be explicitly written by combining the expression of Eq. (31) and the linear shift operator, i.e.,

ucn+1=ucp+ΓscΩs1ML1MΓcsucpE33
=ucp+I3Ωc1ΓcsML1MΓcsucp,E34

where the ΩsRm×m and ΩcRn×n are diagonal matrices with staggered and cell-centered control volumes, respectively. The discrete Laplacian operator LRn×n, defined asLMΩ1M, is symmetric and negative-definite. The checkerboard problem stems from the unrealistic component of the cell-centered velocity field that cannot be eliminated by the pseudo-projection matrix,

uc=uc+I3Ωc1ΓcsML1MΓcsuc,E35

where uc denotes the spurious modes. These’unrealistic’ velocity modes persist unless explicitly removed. Trias et al. [27] suggested regularizing (smoothing approximation) the convective term to address the origin of these ‘unrealistic’ velocity modes while maintaining the numerical solution devoid of nonphysical oscillations.

2.3.5 Constructing the discrete operators on unstructured collocated grids

Skew-symmetry convective operator Cus is obtained in two stages [10]. Firstly, attention is given to the off-diagonal elements. The matrix Cus-dia Cus is skew-symmetry when the interpolation weights of the adjacent discrete variables are set equal to 1/2 hence the discrete normal velocity usfufnf, situated at the centroid of the cell faces f, is expressed as

usf=Γcsucf=12ucc1+ucc2nf,E36

where c1 and c2 represent the cells neighboring the face f. The entries of the matrix Cusk correspond to half of the flux through face f between the adjacent cells i and j, formulated as

Cusk=12usfAf,E37

where Af denotes the face area of f. In addition, for skew-symmetry, the diagonal elements should be zero,

Cusk,k=12fFfiusfAf=0,E38

where the Ffi represents the set of faces neighboring face f. This criterion is met due to the vanishing discrete divergence of us. Thus, the convective operator unstructured collocated grids

Cusuck=fFfk12ucc1+ucc2usfAf,E39

By integrating the continuity equation given in Eq. (16) over an arbitrary cell k with volume Ωckk, we obtain

Ωck,kudV=Ωck,kundS=fFfkSfundS.E40

A discretization of Eq. (40) that is second-order accurate is

fFfkSfundSfFfkusfAf.E41

Hence, the divergence operator is as

MuskfFfkusfAf=0.E42

According to Eq. (22) the integrated pressure gradient operator, ΩsG, equals the negative of the transpose of the divergence operator M. Hence, the discretization of the pressure gradient at the face f follows from Eq. (42)

ΩGpcf=pc1pc2Af,E43

The discrete gradient inherits boundary conditions from the discrete divergence, thus negating the need for specifying boundary conditions for the pressure. Finally, we compute the pressure from a Poisson equation that from the incompressibility constraint.

The Laplacian operator is represented by the matrix,

L=MΩs1M.E44

negative-definite and symmetric matrix resembles the continuous Laplacian operator Δ.

The approach used to discretize the Laplacian operator in the pressure Poisson equation is also employed to discretize the diffusive term in NS equations. The diffusive operator is seen as the product of two first-order differential operators, a divergence and a gradient. The divergence operator is discretized and the gradient operator becomes the transpose of the discrete divergence

Dc=1ReMG=1ReMΩs1M.E45

This procedure results in a symmetric, positive-definite, approximation of the diffusive operator . The collocated diffusive operator for a cell-centered variable ϕc is expressed as,

Dcϕc=1RefFfkϕc1ϕc2Afδnf,E46

where the length δnf approximates the distance between the centroid of cells c1 and c2, calculated as δnf=nfc1c2. Afterwards, the volume of the cell associated with the face-normal velocity at face f is denoted as Ωsf=δnfAf, where the Ffk is the set of faces boarding the face f.

2.3.6 Solver

The RKSymFoam solver’s primary algorithm comprises three hierarchical iterative levels in the case of implicit time discretization [12, 13]:

  1. The outer loop iterates over each RungeKutta (RK) stage i, ranging from 1 to s (see Section 2.3.8 for details).

  2. A loop manages to update the nonlinear convective term.

  3. An inner PISO loop handles the pressure-velocity coupling.

For explicit temporal discretization, a single projection step suffices for the pressure-velocity coupling, and there’s no need to update the convective term.

2.3.7 Temporal discretization

To maintain the favorable conservation and stability properties in discrete time, the time discretization in Eq. (25) is a skew-symmetric operator, implicit temporal integration. In high-Reynolds-number flow simulations, the implicit method may incur higher computational costs compared to the explicit counterpart. Yet in an explicit temporal discretization scheme for convection-diffusion equations, the time step Δt is typically constrained by convective stability conditions such as UΔt<Δy (where U represents the maximum velocity magnitude and Δy denotes the spatial grid size), diffusive stability conditions akin to 2Δt<ReΔy2. The implicit method does not impose such stringent restrictions, allowing for a time step that can be larger. To strike a balance between accuracy and stability, ensuring that the simulation remains globally second-order accurate, the Crank-Nicolson and diagonal-implicit Runge–Kutta methods are chosen.

2.3.8 Butcher tables specify Runge-Kutta methods

After applying the spatial discretization the momentum equations Eq. (16) become

ducdt=Fctucpcwithuctn=ucn.E47

Here, the subscript c denotes the cell-centered values, while the pressure gradient term remains to be discretized. Runge-Kutta (RK) schemes, specified by a Butcher table [29], are used to discretize Eq. (47) in time. Using aij to denote the stage weights for stage i, and ci as the quadrature nodes for the schemes with ci=j=1saij, where i=1,,s, and ti=tn+ciΔt, s denotes the number of stages, and bj denotes the primary weights of the utilized RK scheme, satisfying j=1sbj=1. Explicit schemes are characterized by having only non-zero entries in the lower triangle of the ai,j portion of the table. Entries on or above the diagonal lead to implicit methods as they involve ucn+1 on the right-hand side of Eq. (47).

The intermediate solution u˜ci for stage i at time ti is expressed as

u˜ci=ucn+Δtj=1iaijFctju˜cjcip˜ci,E48

and the final solution ucn+1 at time tn+1=tn+Δt by

ucn+1=ucn+Δtj=1ibjFctju˜cjpcn+1.E49

Here we concentrate on explaining the forward-Euler method as an example, as higher-order explicit RK methods essentially involve a series of forward-Euler stages. We compute an intermediate velocity field ucp through the following predictor step

ucpucnΔtn=HusnucnGcpcp.E50

Here HusnucnΩc1Cusnucn+Ducn, where HusnR3n×3n, and Δtn denotes the time difference between time levels n and n+1. As the intermediate velocity is not divergence-free, we correct the final values for the time step n+1 by adding the following corrections to the intermediate values: ucn+1=ucp+uc and pcn+1=pcp+p˜c, the velocity and pressure correction

uc=ΔtGcp˜c.E51

After obtaining the pressure correction p˜ from the Poisson equation Eq. (32), we can calculate the velocity correction u using Eq. (51). Subsequently, the new velocity and pressure fields for the next time step can be computed using the velocity and pressure corrections uc and p˜c.

The RKSymFoam solver employs the PISO approach for implicit time integration. Let us illustrate this with the backward Euler method. The PISO algorithm comprises a predictor step followed by ncorr corrector steps (or inner iterations), as outlined below [30]:

  1. Predictor step: The initial intermediate velocity uc1 is calculated using the predictor equation

    uc1ucnΔtn=HuspucpGcpcp,E52

    where Huspucp=Husnucn is utilized to refrain from an implicit treatment of the convective and diffusive terms. The resulting initial intermediate velocity may not be divergence-free, necessitating subsequent corrector steps.

  2. Corrector steps: In each correct step, a new pressure field pck and a corresponding revised velocity uck+1 (which is divergence-free MΓcsuck+1=0c,n) are. To enhance the stability of the momentum equation during the corrector step, the operator is partitioned into a diagonal component Adusn acting on uck+1 and an off-diagonal component Hodusn acting on uck, resulting in

uck+1=B1ΔtHodusnuck+ucnΔtB1Gcpck,withk=1,2,,ncorr,E53

where B=IΔtnAdusnR3n×3n. A Poisson equation for pressure pck by taking the divergence of Eq. (53) and employing MΓcsuck+1=0c,n

MBS1Ωs1Mpck=1ΔtMΓcsB1ΔtHodusnuck+ucn,k=1,2,,ncorr,E54

where Bs1ΓcsB1Γsc, is a square diagonal matrix. The implicit scheme isL=MBs1Ωs1M. The pressure correction p˜ck is obtained from the following Poisson equation

MBS1Ωs1Mpck=1ΔtMΓcsB1ΔtHodusnuck+ucnGpcn,k=1,..,ncorr,E55

After obtaining the pressure field pck from the Poisson equation Eq. (54), the corresponding updated velocity can be computed using Eq. (53). This iterative process continues for ncorr iterations until an inner iteration criterion is met. The inner PISO iteration procedure concludes by updating the velocity, which is utilized to solve the non-linearity in the convective the operator H, with the new velocity ucncorr for the subsequent outer iteration.

Advertisement

3. Verification and validation

The test cases encompass channel flow across a range of friction Reynolds numbers Reτ, including 180, 550, 1000, and 2000, along with flow over periodic hills and flow past a circular cylinder. Both static and dynamic minimum-dissipation models are subjected to numerical examinations in channel flow scenarios. The investigation delves into the sensitivity of the QR model concerning spatial discretization schemes, mesh resolution, and Reynolds number variations, comparing standard OpenFOAM discretization against symmetry-preserving discretization methods. This section focuses on the main findings; for comprehensive details, see [31].

3.1 Optimal QR model coefficient

The simulations were performed using the QR model with different model constants for a channel flow at Reτ=180 and periodic hills, aiming to identify the optimal model constant. Figure 1 depicts the normalized mean streamwise velocity against the wall. It demonstrates that the model coefficient C = 0.024 aligns precisely with DNS data. Additionally, utilizing five different error measurements (mean square, absolute, maximum absolute, and slope errors), the constant C = 0.024 was identified as yielding the smallest error. This model coefficient yields accurate results under mesh convergence. Furthermore, the various comparisons carried out for flows over periodic hills demonstrate that the model coefficient of C = 0.024 is again the best one. In literature, the Smagorinsky model coefficient is typically considered optimal within the range of Cs=0.1 to Cs=0.2. Consequently, the ideal coefficient for the QR model falls within the range of Cs2. Note that the model coefficient utilized in the QR model is the Smagorinsky.

Figure 1.

Normalized mean streamwise velocity versus wall-normal distance in channel flow simulation at Reτ=180.

3.2 Static QR model compared to dynamic models in channel flow at Reτ=180

The static QR model is compared with four dynamic models. The dynamic procedure is based on the Lilly-Germano identity [2]. All subgrid models depicted in Figure 2 closely approximate the DNS results in the near-wall region, particularly the dynamic QR model, which agrees with the DNS data. A detailed examination of the figure highlights that in the logarithmic wall range y+>30, the static QR, dynamic minimum-dissipation (DMD), and dynamic QR (DQR) models closely match the DNS results. Conversely, the hybrid dynamic (HDM) and dynamic Smagorinsky (DSM) models nearly overlap and deviate from the DNS data in the logarithmic wall region.

Figure 2.

Average streamwise velocity u+ obtained from static QR and dynamic models compared to DNS data.

, both the static QR and DQR models exhibit similar performance, with negligible differences in errors. On the other hand, the DSM and HDM demonstrate comparatively lower accuracy, while the DMD model exhibits the largest error.

In summary, this study underscores the reliability of the static QR model. A suitably chosen model coefficient can yield results akin to a dynamic model while cutting computational costs by a factor of three.

3.3 Sensitivity to expansion ratio in channel flow at Reτ=180

This investigation focuses on the expansion ratio of the mesh in the wall-normal direction. The expansion ratio is defined as follows

Expansion  ratio=Δyc+Δyw+,E56

where Δyc+ represents the size of the cell at the channel center, while Δyw+ denotes the size of the first cell adjacent to the wall. Figure 3 illustrates the normalized mean streamwise and rms velocity against wall distance in wall unit at Reτ=180, employing a grid resolution of 40×50×30. The normalized uniform grid spacing in the streamwise direction is Δx+=18 and in spanwise direction is Δz+=12. The expansion ratios in the wall-normal direction are (approximately) 5, 10, 15, 20, and 25. The first normalized wall-normal grid height is Δyw+=1.728. The time step for simulation is Δt+=36. As the expansion ratio increases, the near-wall mesh, while the mesh in the channel center.

Figure 3.

Normalized mean streamwise and rms velocity using QR model with C = 0.023 at Reτ=180 expansion ratios: 5,10,15, 20, and 25.

For the mean streamwise velocity u+, reducing the expansion ratio from 15 to 5 leads to increasingly pronounced overestimation of the mean velocity beyond y+>10. However, this overestimation exacerbates further with expansion ratios of 20 and 25.

For the velocity fluctuation in the streamwise direction uu, lowering the expansion ratio from 25 to 5 results in notably exaggerated peaks in the near-wall region and relatively diminished overestimation in the central channel. For the velocity fluctuation in the wall-normal υυ and spanwise ww directions (not shown here), reducing the expansion ratio from 20 to 5 causes overestimation of peak values near the wall and comparatively smaller underestimations in the central channel. The highest expansion ratio, 25, underestimates the near-wall peak but provides more accurate values in the channel center.

3.4 Grid Independence study in channel flow at Reτ=550

This section of the survey investigates how different grid configurations affect the mean and root-mean-square velocities at a Reynolds number of Reτ=550, utilizing a QR model with C = 0.023. Four grid setups are considered: 1283, 643, 323, and 163. Figure 4 illustrates the normalized mean streamwise velocity as a function of wall distance in wall units, with the channel dimensions being 2π×2×π. The expansion in the wall-normal direction is approximately 10, and the simulation employs a time step of Δt+=55.

Figure 4.

Comparison of streamwise velocities using QR Model (C = 0.023) at Reτ=550 on various mesh resolutions (1283, 643, 323, 163) with DNS results at Reτ=500.

Initially, the analysis focuses on the results of the simulation conducted using the 163 grid points (depicted by the blue line on the left side of Figure 4). The first mesh point is situated around y+=7, indicating a significant distance from the wall. Despite the coarse mesh, the obtained results surprisingly exhibit reasonable agreement with the overall trend observed in the other simulations.

A noteworthy observation is the discernible trend that emerges with increasing mesh resolution: the mean velocity profile progressively aligns more closely with DNS data, particularly evident when utilizing 1283 mesh points. As resolution increases, the proximity of the initial mesh points to the wall decreases, resulting in a greater concentration of mesh in the near-wall region. Consequently, the 1283 case demonstrates complete agreement with DNS data in the near-wall region.

3.5 Symmetry-preserving compared to standard OpenFOAM discretization

In this section, we illustrate the effect of the numerical schemes by comparing the symmetry-preserving discretization with the Pimple algorithm conjugated with Gauss linear schemes in three benchmarks, that is, periodic channel flow, periodic hill flow, and flow over a cylinder.

The channel flow outcomes depicted in Figure 5, the results obtained symmetry-preserving techniques closely resemble those of the DNS reference. Notably, in the region where y+10, the symmetry-preserving scheme with the QR model precisely aligns with the DNS results.

Figure 5.

The mean streamwise velocity obtained from the QR model with standard OpenFOAM and symmetry-preserving (SP) schemes for fully developed turbulent channel flow at Reτ=1000.

The symmetry-preserving method enhances velocity fluctuation predictions in the region where y+220. In other areas of the computational domain, both methods yield comparable results. For the Reynolds shear stress uυ, the two methods give the same accurate prediction.

Additionally, it has been observed that the contribution of the subgrid-scale model to the diffusive flux in both methods is lower than 20% of the contribution of molecular viscosity [31].

3.6 Application to periodic hill flow

The phenomenon of flow separation from curved surfaces followed by reattachment is frequently encountered in engineering applications. To evaluate the effectiveness of the proposed minimum-dissipation model in OpenFOAM for simulating separated flows, three-dimensional flow simulations over periodic hills at Re = 10,595 as, illustrated in Figure 6, have been conducted.

Figure 6.

The side view of the periodic hill flow.

The mean velocity and Reynolds stress calculated by the QR model, without incorporating a wall damping function, are compared with experimental data [32] and three numerical experiments [33]. It is important to note that all large-eddy simulations are based on second-order accurate numerical techniques both in space and time. Additionally, the QR model grid comprises six times fewer grid points compared to the grid used in the reference CFD data. The QR simulation mesh comprises 2.56 million grid points and necessitates 6 hours for simulation on two nodes, each equipped with 128 cores. This mesh size proves adequate for modeling this case, as further refinement does not yield improved results.

3.6.1 Separation and reattachment lengths

The separation and reattachment points are obtained at Re = 11,230, that is, the bulk velocity ub=1.06, using Gauss linear spatial discretization and backward temporal discretization. The separation point is accurately pinpointed by numerically solving boundary layer equations under pressure-adverse conditions. The QR model predicts the separation point, where the wall shear stress reaches zero, to be approximately x/H0.175, slightly smaller than the reference value of x/H0.19. This deviation is expected, as the separation point tends to shift upstream with increasing Reynolds numbers. The precise location of the separation point significantly influences the point of reattachment. Recirculation initiates around x/H0.27 and extends until approximately x/H4.71, with the main recirculation bubble spanning a length of roughly 4.48. The reattachment position, where the dividing streamline reconnects with the wall, is determined to be at x/H=4.71, closely aligning with the reference value of 4.69 [32].

3.6.2 Sensitivity to the numerical schemes of standard OpenFOAM

A significant discrepancy has been identified in the prediction of the streamwise velocity component u/ub within the range 1.5<y/H<2.8, despite employing the same Reynolds number as the reference case (Re = 10, 595).

In an effort to understand the cause of this discrepancy, various interpolation numerical schemes for solving the divergence term and pressure gradient have been evaluated. These schemes include central difference, filtered and limited central difference, linear upwind blended with central difference, least squares, and detached eddy discretization schemes. However, the outcomes of these different numerical schemes fail to accurately predict the mean velocity, as illustrated in Figure 7. Consequently, it has been determined that increasing the Reynolds number to better align with the reference data is necessary. Furthermore, despite minor differences observed between the simulations, the central difference schemes are ultimately selected due to their satisfactory performance in previous validation cases. These discrepancies, while present, are deemed insignificant in distinguishing one technique from another.

Figure 7.

Streamwise velocity at five different positions with varying finite volume discretization schemes.

3.6.3 Sensitivity to the Reynolds number

In order to address the underprediction of the streamwise velocity component u/ub within the range of 1.5<y/H<2.8 while maintaining the Reynolds number equal to the reference value of Re = 10,595, simulations at higher Reynolds numbers were explored. The most notable revelation from the analysis, as depicted in Figure 8, is that the mean velocity and Reynolds stress obtained with a 6.5% higher Reynolds number align with the B-spline reference. Furthermore, a 7% increase in Reynolds number corresponds to agreement with experimental data (not illustrated here). These findings confirm that augmenting the Reynolds number by 3−7% successfully aligns the simulations with the reference data. Note that despite the experimental data and three LES simulations used for comparison being conducted at the same Reynolds number of Re = 10,595, they exhibit discrepancies among themselves.

Figure 8.

Streamwise velocity (upper) and Reynolds stress (lower) at ten distinct locations, with a 6.5% increment in Reynolds number compared to the B-spline data. 11,284 refers to the QR results and LES-1 indicates the reference B-spline results.

3.7 Application to flow over a cylinder

3.7.1 Comparison between standard OpenFOAM spatial discretization

To examine the impact of different numerical schemes employed in standard OpenFOAM spatial discretization on turbulence, simulations are conducted for both periodic hills and flow over a cylinder.

Figure 9 displays the mean streamwise velocity u resulting from three distinct numerical discretizations applied to flow over a cylinder. Notably, discrepancies in mean streamwise velocity u are observed in the vicinity of the wake x/D<2.02. Specifically, the upwind-blended scheme predicts a U-shaped mean velocity profile at x/D=1.06, transitioning to a markedly lower V-shaped profile at x/D=1.54 and 2.02, compared to the other two schemes.

Figure 9.

The mean streamwise velocity at six locations (left) and the Reynolds stress in the cross-section at x/D=1.54 (right) are depicted for flow over a cylinder at ReD=3900. The QR model is utilized with filtered central difference (Flinear), pure central difference (Linear), and upwind-biased (LUST) schemes.

Analysis from the central difference simulation reveals that the onset of turbulence in the separating shear layers occurs nearer to the cylinder, leading to the formation of a shorter vortex region and the development of a V-shaped velocity profile. Consequently, the shear layers exhibit reduced length, and the recirculation region diminishes in size [31]. In the downstream region, variations in numerical schemes appear to have minimal impact on mean velocity. However, simulations of periodic hills demonstrate that the selection of numerical schemes substantially influences both mean and root-mean-square variables.

Advertisement

4. Conclusions

A comprehensive comparison between the minimum-dissipation model (QR) and experimental/DNS/LES results across channel flow, flow past a circular cylinder, and flow over periodic hills reveals generally favorable agreement, particularly on coarse meshes.

The findings suggest that the static QR model is comparable to dynamic models while reducing computational costs. Notably, a model coefficient of C = 0.024 produces the most accurate predictions, with the contribution of the subgrid model diminishing to less than 20% of the molecular viscosity as mesh resolution increases to its finest level.

The characteristics of turbulence exhibit dependence on spatial discretization methods. Various comparisons highlight the superiority of the QR model when paired with symmetry-preserving discretization over standard OpenFOAM discretization techniques. Among standard OpenFOAM discretization methods, central difference schemes emerge as the most effective. Finally, for the periodic hill flow, augmenting the Reynolds number by 3−7% successfully aligns the simulations with the reference data.

Advertisement

Acknowledgments

This work was supported by the Chinese Scholarship Council and the University of Groningen under Grant (CSC No. 202006870021). The computing time was granted by the Dutch Research Council (NWO) under project 2022.009.

The first version was published as a preprint titled ‘Minimum-dissipation model for large-eddy simulation in OpenFOAM – A study on channel flow, periodic hills, and flow over a cylinder’ on arXiv.org in 2023.

References

  1. 1. Smagorinsky J. General circulation experiments with the primitive equations: I. The basic experiment. Monthly Weather Review. 1963;91(3):99-164
  2. 2. Lilly D, K. On the application of eddy viscosity concept in the inertial sub-range of turbulence. NCAR Manuscript. 1966;123:1-19
  3. 3. Ducros Frédéric, Nicoud Franck, Poinsot Thierry. Wall-Adapting Local Eddy-Viscosity Models for Simulations in Complex Geometries. Numerical Methods for Fluid Dynamics VI. 1998;6:293-299
  4. 4. Vreman AW. An eddy-viscosity subgrid-scale model for turbulent shear flow: Algebraic theory and applications. Physics of Fluids. 2004;16(10):3670-3681. DOI: 10.1063/1.1785131. ISSN 1070-6631
  5. 5. Verstappen R. When does eddy viscosity damp subfilter scales sufficiently? Journal of Scientific Computing. 2011;49(1):94-110
  6. 6. Verstappen R. How much eddy dissipation is needed to counterbalance the nonlinear production of small, unresolved scales in a large-eddy simulation of turbulence? Computers & Fluids. 2018;176:276-284
  7. 7. Rozema W, Bae HJ, Moin P, Verstappen R. Minimum-dissipation models for large-eddy simulation. Physics of Fluids. 2015;27(8):085107
  8. 8. Abkar M, Bae HJ, Moin P. Minimum-dissipation scalar transport model for large-eddy simulation of turbulent flows. Physical Review Fluids. 2016;1(4):041701
  9. 9. Zahiri A-P, Roohi E. Anisotropic minimum-dissipation (AMD) subgrid-scale model implemented in OpenFOAM: Verification and assessment in single-phase and multi-phase flows. Computers & Fluids. 2019;180:190-205
  10. 10. Verstappen RWCP, Veldman AEP. Symmetry-preserving discretization of turbulent flow. Journal of Computational Physics. 2003;187(1):343-368
  11. 11. Trias FX, Lehmkuhl O, Oliva A, Pérez-Segarra CD, Verstappen RWCP. Symmetry-preserving discretization of Navier–stokes equations on collocated unstructured grids. Journal of Computational Physics. 2014;258:246-267. ISSN 0021-9991
  12. 12. Komen EMJ, Hopman JA, Frederix EMA, Trias FX, Verstappen RWCP. A symmetry-preserving second-order time-accurate PISO-based method. Computers & Fluids. 2021;225:104979
  13. 13. Hopman Jannes, Edo Frederix. A RKSymFoam Github page. Netherlands. 2023. Available from: https://github.com/janneshopman/RKSymFoam [Accessed: June 13, 2023]
  14. 14. Hoyas S, Jiménez J. Scaling of the velocity fluctuations in turbulent channels up to Reτ=2003. Physics of Fluids. 2006;18(1):011702
  15. 15. Hoyas S, Jiménez J. Reynolds number effects on the Reynolds-stress budgets in turbulent channels. Physics of Fluids. 2008;20(10):101511
  16. 16. Jimenez J, Hoyas S. Turbulent fluctuations above the buffer layer of wall-bounded flows. Journal of Fluid Mechanics. 2008;611:215-236
  17. 17. Del Juan Carlos Álamo and Javier Jiménez. Direct numerical simulation of the very large anisotropic scales (vlas) in a turbulent channel. San Diego. In: APS Division of Fluid Dynamics Meeting Abstracts. 54, KF–003, 2001. 54, KF–003, 2001
  18. 18. Del Alamo JC, Jiménez J. Spectra of the very large anisotropic scales in turbulent channels. Physics of Fluids. 2003;15(6):L41-L44
  19. 19. Del Alamo JC, Jiménez J, Zandonade P, Moser RD. Scaling of the energy spectra of turbulent channels. Journal of Fluid Mechanics. 2004;500:135-144
  20. 20. Ong L, Wallace J. The velocity field of the turbulent very near wake of a circular cylinder. Experiments in Fluids. 1996;20(6):441-453
  21. 21. Kravchenko AG, Moin P. Numerical studies of flow over a circular cylinder at ReD=3900. New York: Physics of Fluids. 2000;12(2):403-417
  22. 22. Courant R, Hilbert D. Methods of Mathematical Physics: Partial Differential Equations. New York: John Wiley & Sons; 2008
  23. 23. Payne LE, Weinberger HF. An optimal poincaré inequality for convex domains. Archive for Rational Mechanics and Analysis. 1960;5(1):286-292
  24. 24. Manteufel TA, White Jr AB. The numerical solution of second-order boundary value problems on nonuniform meshes. Mathematics of Computation. 1986;47:511
  25. 25. Shashank JL, Iaccarino G. A co-located incompressible Navier-stokes solver with exact mass, momentum and kinetic energy conservation in the inviscid limit. Journal of Computational Physics. 2010;229(12):4425-4430
  26. 26. Chorin AJ. Numerical solution of the Navier-stokes equations. Mathematics of Computation. 1968;22(104):745-762
  27. 27. Yanenko NN. Economical implicit schemes (method of fractional steps). In: Doklady Akademii Nauk. Vol. 134. Moscow: Russian Academy of Sciences; 1960. pp. 1034-1036
  28. 28. Blair J, Perot. An analysis of the fractional step method. Journal of Computational Physics. 1993;108(1):51-58
  29. 29. Butcher JC. A history of Runge-Kutta methods. Applied Numerical Mathematics. 1996;20(3):247-260
  30. 30. Issa RI. Solution of the implicitly discretized fluid flow equations by operator-splitting. Journal of Computational Physics. 1986;62(1):40-65
  31. 31. Sun J, Verstappen R. Minimum-dissipation model for large-eddy simulation in OpenFOAM-A study on channel flow, periodic hills and flow over cylinder, Netherlands. arXiv. 2023, arXiv:2309.04415. DOI: 1048550/arXiv.2309.04415
  32. 32. Ch R, Pfleger F, Manhart M. New experimental results for a LES benchmark case. In: Direct and Large-Eddy Simulation VII: Proceedings of the Seventh International ERCOFTAC Workshop on Direct and Large-Eddy Simulation, Held at the University of Trieste. September 8–10, 2008. Netherlands: Springer; 2010. pp. 69-74
  33. 33. Temmerman L, Leschziner MA. Large eddy simulation of separated flow in a streamwise periodic channel constriction. In: Second Symposium on Turbulence and Shear Flow Phenomena. London, UK: Begel House Inc.; 2001

Written By

Jing Sun and Roel Verstappen

Submitted: 13 February 2024 Reviewed: 08 May 2024 Published: 22 July 2024