A Consistent Dual-Mesh Framework for Hybrid LES

0 downloads 0 Views 2MB Size Report
Nov 14, 2011 - Heng Xiao∗, Patrick Jenny∗∗. Institute of Fluid Dynamics, ETH Zürich, 8092 Zurich, Switzerland. Abstract. In this work, we propose a consistent ...
A Consistent Dual-Mesh Framework for Hybrid LES/RANS Modeling Heng Xiao∗, Patrick Jenny∗∗ Institute of Fluid Dynamics, ETH Z¨ urich, 8092 Zurich, Switzerland

Abstract In this work, we propose a consistent turbulence-modeling framework for hybrid LES/RANS modeling. In this framework, the filtered and Reynolds averaged Navier-Stokes (RANS) equations are solved simultaneously in the whole domain on their respective meshes. Consistency between the two solutions is achieved in terms of velocity, pressure, and turbulent quantities through additional drift terms in the corresponding equations. This approach leads to clean conditions at the LES/RANS interfaces. Note that this general framework does not depend on the specific choice of LES and RANS models. A hybrid LES/RANS solver is developed within this framework and used to simulate the flow in a plane channel and that in a channel with periodic hills. The results demonstrate that the hybrid solver leads to significantly improved results with moderate computational overhead compared to traditional LES, making it a promising candidate for industrial flow simulations. Keywords: hybrid LES/RANS, turbulent flow, wall-bounded flow



Corresponding author. Tel: +41 44 632 5189; Fax: +41 44 632 1147 Primary corresponding author Email addresses: [email protected] (Heng Xiao), [email protected] (Patrick Jenny) ∗∗

Preprint submitted to Journal of Computational Physics

November 14, 2011

1

1. Introduction

2

Large-Eddy Simulation (LES) has gained popularity and successes in the

3

past decades, particularly for free-shear flows, where the computational cost

4

of LES is only weakly dependent on the Reynolds number (∼ Re0.4 ) [1, 2].

5

In wall-bounded flows, however, to resolve the near-wall region (inner layer),

6

the computational cost of LES scales as Re1.8 , similar to Direct Numerical

7

Simulation (DNS), which scales as Re9/4 . Therefore, the high computational

8

cost in wall-bounded flows is still a major hurdle for the application of LES in

9

industrial and practical flows. In many LES conducted for wall-bounded flows

10

in practice, wall modeling based on logarithmic profile or other theoretical

11

models is used to avoid the high resolution [3]. However, in complex flows

12

the validity of these theoretical models is in question.

13

To overcome this difficulty, various hybrid LES/RANS approaches have been

14

developed, where RANS (Reynolds-averaged Navier Stokes) equations are

15

solved in the near-wall region instead of using wall functions. A recent re-

16

view by Fr¨ohlich and von Terzi [3] presented an excellent survey and proposed

17

a detailed classification of these approaches. They divided current hybrid

18

LES/RANS methods into the following categories: (1) blending turbulence

19

models, (2) interfacing turbulence models, (3) segregated turbulence models,

20

and (4) second generation unsteady RANS models (2G-URANS), includ-

21

ing partially averaged Navier-Stokes (PANS) and scale-adaptive simulation

22

(SAS), etc.

23

The segregated models do not mainly target the difficulty related to wall-

24

bounded flows mentioned above and are not further discussed here. An im-

25

portant feature of the 2G-URANS models is that the turbulent stress models

26

do not explicitly depend on the grid spacing, and thus they do not revert

27

to DNS no matter how fine the grid is. According to the definition in the

28

review [3], they do not strictly qualify as a hybrid LES/RANS method, since 2

29

they do not have the LES component (i.e., with a turbulence model depend-

30

ing on grid spacing and reverting to DNS with refined grids), and thus they

31

are referred to as 2G-URANS. In addition, most models in this category are

32

still young and only preliminary results are available at present. Another

33

approach that is omitted in the review of Fr¨ohlich and von Terzi [3] is the

34

method based on hybrid-filtered Navier Stokes (HFNS) equations, proposed

35

by Germano [4]. More recently, Rajamani and Kim [5] conducted some a

36

priori tests and preliminary simulations using this method. This method is

37

also in its early stage of development.

38

The introduction here focuses on two categories of hybrid LES/RANS meth-

39

ods: blending models and interfacing models, because at present the most

40

widely used and mature hybrid LES/RANS models belong to one of the two

41

categories. The blending turbulence models take advantage of the structural

42

similarity of LES and RANS equations. That is, the LES and RANS equa-

43

tions have similar momentum equations, and the only difference lies in how

44

turbulent stresses are modeled. In fact, many subgrid scale (SGS) turbulence

45

models are developed from the corresponding RANS turbulence models. In

46

the blending approach, the turbulent stress tensor is computed as follows: hybrid

τij

= f τijRANS + (1 − f )τijLES ,

(1)

47

where the blending factor f is a smooth function ranging from 0 to 1, de-

48

pending on the spatial location and possibly time as well. The work by Uribe

49

et al. [6] is an example that blends turbulent stresses. Furthermore, if both

50

the LES and the RANS models that are blended belong to the class of the

51

eddy viscosity models, which is often the case in practice, one could directly

52

blend the turbulent eddy viscosity in the same way as in Equation (1). The

53

blending functions are often chosen in an ad-hoc way and the coefficients are

54

then calibrated empirically. This feature leaves a lot of freedom for tuning

55

and thus impairs its predictive capabilities, although sometimes very good

56

results can be obtained with the tuning. 3

57

The interfacing approach is the limiting case of the blending approach with

58

vanishing blending region, or equivalently using a step function as blending

59

function. It thus avoids the ad hoc choice of blending functions. However,

60

the two categories of hybrid models share most other limitations, which will

61

be detailed below. Notable examples of the interfacing models include the

62

detached-eddy simulation (DES) method [7] and its improved variants such as

63

delayed DES [8], which are among the most widely used hybrid LES/RANS

64

methods. However, the DES is based on the Spalart-Allmaras RANS model,

65

which is a one-equation model mostly tuned for external aerodynamic flows

66

at high Reynolds numbers for applications such as flow around airfoils. Its

67

performance on internal wall-bounded flows is not clear in general, although

68

much research is going on [e.g. 9].

69

In general, the interfacing models suffer from the fundamental inconsistency

70

between LES and RANS. The blending models have the same issue, although

71

to a lesser extent. Specifically, even though the LES and RANS equations are

72

similar in structure, the quantities (velocity, pressure, etc.) in the LES equa-

73

tions are filtered quantities and in RANS equations ensemble- or Reynolds-

74

averaged quantities. Consequently, it is difficult for the LES to sustain proper

75

fluctuations near the interfaces (in the blending models) or for RANS to pro-

76

vide boundary conditions to the LES (in the interfacing models). This in-

77

consistency has important physical implications. For example, the blending

78

can lead to artificial “super-streaks” near walls due to the “modeled stress

79

depletion (MSD)”, and consequently spurious buffer layers can be observed

80

in the mean flow profiles near the LES/RANS interfaces [e.g. 3, 10].

81

Additional forcing near the interface is necessary to sustain the velocity fluc-

82

tuations in the LES region. Various attempts have been made to address this

83

problem. For example, Piomelli et al. [11] applied a stochastic forcing term to

84

the momentum equations. Davidson and Dahlstr¨om [12] added fluctuations

85

to the momentum equations according to a DNS of a generic boundary layer. 4

86

Davidson [13] used the backscatter from a scale-similarity SGS stress model

87

as the forcing. Despite some successes, many difficulties remain, which are

88

not detailed here.

89

In addition to the fundamental difficulty discussed above, the blending and

90

interfacing approaches also have the drawbacks of relying upon the existence

91

of clear LES/RANS interfaces, which are difficult to specify in complex ge-

92

ometries. In practical simulations, it is more desirable to dynamically and

93

individually define each cell to be an LES cell or a RANS cell depending on

94

certain criteria, that is, a non-zonal approach is preferred.

95

To avoid the fundamental inconsistencies mentioned above, we propose a con-

96

sistent framework for turbulence modeling, where the filtered and Reynolds-

97

averaged equations are solved simultaneously in the entire domain. That

98

is, both LES and RANS simulation are conducted, which results in some

99

redundancy. To ensure consistency between the two solutions in terms of

100

velocity, pressure, and turbulent quantities, additional drift terms are added

101

to the corresponding equations. This approach leads to very clean condi-

102

tions at the LES/RANS interfaces. By solving two sets of equations (i.e.

103

filtered and RANS) separately, the difficulty of “modeled stress depletion” is

104

avoided. There is no need to add random fluctuations into the LES. On the

105

other hand, we acknowledge that MSD is only one of the many difficulties

106

faced by hybrid methods. Other notable difficulties include the prediction of

107

laminar-turbulence transition and the generation of inflow boundary condi-

108

tions. Admittedly, in these aspects a hybrid solver per se (this one included)

109

does not gain any advantage compared to its individual components. This

110

framework does not specifically deal with these issues. Readers are referred

111

to the recent work of Sagaut and co-workers [14] for an example of such ef-

112

forts. Finally, we emphasize that this general framework does not depend on

113

the specific choice of LES and RANS models.

5

114

The rest of the paper is organized as follows. The details of the framework

115

and the algorithm are presented in Section 2. A hybrid LES/RANS solver

116

is developed as a specific implementation of this general framework. The

117

detailed implementation of the solver, the turbulence models, and the nu-

118

merical methods used in the simulations are presented in Section 3. The

119

developed solver is used to simulate the flow in a plane channel and a chan-

120

nel with periodic hills. The results from numerical simulations are presented

121

in Section 4. Additional issues and further research related to this framework

122

are discussed in Section 5. Section 6 concludes the paper.

123

2. Proposed Framework and Algorithm

124

2.1. Model Equations

125

For incompressible flows with constant density, in the context of LES the

126

filtered momentum and pressure equations can be written as follows:  ∂τijsgs ∂ U¯i ∂ U¯i U¯j 1 ∂ p¯ ∂ 2 U¯i + =− +ν − + QLi ∂t ∂xj ρ ∂xi ∂xj ∂xj ∂xj  ∂QLi ∂2 1 ∂ 2 p¯ U¯i U¯j + τijsgs + =− , and ρ ∂xi ∂xi ∂xi ∂xj ∂xi

(2a) (2b)

127

where U¯i , p¯, and τijsgs are filtered velocity, filtered pressure, and residual

128

stresses, respectively; t and xi are time and space coordinates, respectively;

129

ν is the kinematic viscosity; ρ is the fluid density, which is assumed constant;

130

QLi is the drift force in the filtered equations to ensure consistency between

131

LES and RANS and will be defined later in Equation (10).

132

Similarly, the Reynolds-averaged momentum and pressure equations are writ-

6

133

ten as: ∂hUi i ∂ (hUi ihUj i) 1 ∂hpi ∂ 2 hUi i ∂hui uj i + =− +ν − + QRi ∂t ∂xj ρ ∂xi ∂xj ∂xj ∂xj 2 2 1 ∂ hpi ∂ ∂QRi and =− (hUi ihUj i + hui uj i) + , ρ ∂xi ∂xi ∂xi ∂xj ∂xi

(3a) (3b)

134

where hUi i, hpi, and hui uj i are filtered velocity, filtered pressure, and Reynolds

135

stresses, respectively; QR i is the drift force in the RANS equations to ensure

136

consistency and will be defined later in Equation (15).

137

As a clarification of the terminology: the ensemble- or Reynolds-averaged

138

Navier-Stokes equations in the context of most hybrid LES/RANS studies

139

are generally unsteady. However, usually they are still referred to as RANS

140

instead of URANS equations. This convention of terminology is followed in

141

this paper.

142

2.2. Consistency and Drift Terms

143

To discuss the consistency issue, we first define an exponentially weighted

144

averaging operator, h•iAVG , applied on a time-dependent quantity φ(t): Z t 1 0 AVG hφi (t) = φi (t0 ) W (t − t0 )dt0 with W (t − t0 ) = e−(t−t )/T , (4) T −∞

145

where t is the current time, t0 denotes earlier times over which the averaging

146

is performed, and T is the time scale for the averaging. The exponential

147

weighting function ensures that the more recent time is weighted stronger

148

than earlier times. Exponential function is chosen because is particularly

149

convenient for implementation (see Equation (18) in Section 2.4). With this

150

definition, the exponentially weighted average velocity, turbulent stresses,

7

151

and dissipation rate in the LES are Z t 1 0 AVG hU¯i i (t) = U¯i (t0 )e−(t−t )/T dt0 , (5a) T −∞ Z  1 t  00 0 00 0 0 AVG hτij i (t) = ui (t )uj (t ) + τijsgs (t0 ) e−(t−t )/T dt0 , (5b) T −∞ Z  1 t  ¯ 0 ¯ 0 0 AVG 2ν Sij (t )Sij (t ) − τijsgs (t0 )S¯ij (t0 ) e−(t−t )/T dt0 , (5c) and hεi (t) = T −∞

152

respectively, where u00i = U¯i − hUi iAVG is the fluctuating velocity with respect

153

to the exponentially weighted average. In Equation (5b), the total turbulent

154

stress in the LES is defined as τij = u00i u00j + τijsgs ,

(6)

156

which includes the resolved part and the residual part. Similarly in Equation (5c), the total dissipation rate in LES, ε = 2ν S¯ij S¯ij −τ sgs S¯ij , also includes

157

the resolved part and the residual part. The resolved rate-of-strain tensor is

158

defined as

155

ij

1 S¯ij = 2 159 160

 ¯  ∂ Ui ∂ U¯j + . ∂xj ∂xi

(7)

One can think of the “exponentially weighted averaging operator” h(·)iAVG (t) = R 0 1 (·)(t0 )e−(t−t )/T dt0 as a linear operator, which operates on any instantaT −∞

161

neous field/quantity, and results in an corresponding exponentially weighted

162

average field/quantity. In Eq (5a), it operates on the velocity field and ob-

163

tains the EWA velocity; in (5b) and (5c) it operates on turbulent stresses

164

and on dissipation, obtaining EWA turbulent stresses and EWA dissipation,

165

respectively.

166

The legitimacy of the LES/RANS coupling shown above relies on the as-

167

sumption that the velocity (and other quantities) can be decomposed into the

168

following components: the mean flow, the resolved fluctuating flow, and the

169

unresolved fluctuating flow. This is a special case of the general framework

170

of multilevel turbulence decomposition [15, 16]. In addition, we assume that 8

171

the exponentially weighted averaging is approximately the same as ensemble-

172

or Reynolds-averaging, i.e., hφiAVG ≈ hφi.

(8)

173

This assumption in Eq. (8) is theoretically rigorous only if the filtering is re-

174

stricted to homogeneous direction of the flow with constant grid spacing [16],

175

which would certainly be violated in practice. However, compared to many

176

of the current hybrid methods where essentially the filtered quantities are re-

177

garded equal to the Reynolds averaged quantities at the interfaces or in the

178

blending regions, this approximation is a significant improvement in terms

179

of consistency.

180

With the two assumptions above, if both filtered and RANS equations are

181

solved for the same flow, the consistency requirements suggest that hU¯i iAVG ≈ hUi i,

(9a)

hτij iAVG ≈ hui uj i,

(9b)

and hεiAVG ≈ εR ,

(9c)

182

where εR is the modeled dissipation rate as in k–ε two-equation turbulence

183

models [17].

184

weighted average quantities, instead of the filtered quantities themselves,

185

are made consistent with their RANS counterparts. This is fundamentally

186

different from the LES/RANS coupling in previous hybrid methods. For this

It is emphasized that in this framework the exponentially

188

reason, we shall also refer to the Exponentially Weighted Average quantities as averaged quantities or EWA quantities for simplicity, i.e., hU¯i iAVG , hτij iAVG ,

189

and hεiAVG are called EWA velocity, EWA turbulent stress tensor, and EWA

190

dissipation rate, respectively. For statistically stationary flows, the consis-

191

tency is rigorous by using a large averaging time scale, although a finite value

192

is usually taken for convergence reasons. For transient flows with coherent

193

structures, however, the averaging time scale should be smaller than the time

194

scale of these structures. See Section 4.4 for more discussions.

187

9

195 196

To fulfill the consistency requirements above, the drift force in the filtered equations is formulated as  (hUi i − hU¯i iAVG )/τl + Gij (hU¯j iAVG − U¯j )/τg L Qi = 0

in RANS regions in LES regions, (10)

197

where Gij =

198

hτij iAVG − hui uj i . hτkk iAVG + huk uk i

Equation (10) can be equivalently written as QLi = QL,u + QL,g i i

199

(11)

in RANS regions,

(12)

where ≡ (hUi i − hUi iAVG )/τl QL,u i and QL,g ≡ Gij (hUj iAVG − U¯j )/τg . i

(13a) (13b)

200

We will interpret QL,u and QL,g shortly. LES and RANS regions refer to i i

201

the sub-domains which are well resolved and under-resolved for LES, respec-

202

tively; τl and τg are relaxation time scales related to the EWA velocity and

203

EWA turbulent stresses, respectively. The drift term QLi is only active in the

204

RANS region, where the LES is under-resolved. The first part, QL,u i , lets the

205

average filtered velocity relax towards the RANS velocity. The second part,

206 207

QL,g i , lets the turbulent stresses relax towards the RANS modeled turbulent stresses. Note that the term QL,u only changes the average of U¯i and not the i

QL,g i

208

fluctuations, while

209

The term QL,g can be interpreted as follows. First, it is a linear function i

210

of the tensor Gij , defined as the difference between hτij iAVG and hui uj i (nor-

211 212 213

only changes the fluctuations and not the average.

malized by the sum of their traces). This serves to let hτij iAVG relax towards hui uj i, which is achieved by scaling the fluctuation u00j = U¯j − hUj iAVG of the filtered velocity with respect to the averaged velocity. 10

214

For many RANS models, the reconstructed Reynolds stresses may not be

215

reliable, particularly for separate flows. For example, in the standard k–ε

216

model, the reconstructed Reynolds stresses do not account for the anisotropy

217

near the wall. In these scenarios, it may be better to only require consistency

218

between total turbulent kinetic energy in LES and RANS, that is hτii /2iAVG ≈

219

k R , instead of requiring the full turbulent stress tensors to be consistent

220

(k R being the turbulent kinetic energy in RANS; usually one of the solved

221

quantities). Therefore, QL,g is reformulated as i QL,g = G(hUi iAVG − U¯i )/τg , i where G =

222

223

(14)

R

hτii /2i −k AVG hτii /2i + kR AVG

is a scalar instead of a tensor as in Equation (11). Similarly, the drift force in the RANS equations is:  (hUi iAVG − hUi i)/τr in LES regions R Qi = 0 in RANS region,

(15)

224

where τr is the relaxation time scale for the RANS velocity. Also note that

225

QR i is only active in the LES region, where the LES is well resolved. Drift

226

terms are also added in a similar way to RANS model equations for the

227

turbulent quantities. For example, in k–ε turbulence models, the drift terms

228

for k- and ε- equation are  Qk = hτii /2iAVG − k R /τr  and Qε = hεiAVG − εR /τr .

(16a) (16b)

229

Both source terms are active only in the LES region. Here, the same re-

230

laxation time scale τr as in the velocity equations is adopted, although it is

231

possible to use different time scales.

232

It can be seen that there are several algorithmic parameters in this formu-

233

lation, i.e., T , τl , τg , and τr . It will be discussed in Section 4.4 how these 11

234

algorithmic parameters should be chosen and that the computational results

235

are not very sensitive on these parameters, as long as they are within a rea-

236

sonable range. The determination of whether a region is well resolved or not

237

in LES can be done on a cell-by-cell basis. Therefore, a clear segregation of

238

the LES and RANS region is not essential for the proposed framework.

239

2.3. Dual-mesh approach

240

In this framework, both the filtered equations and the RANS equations are

241

solved in the whole domain. While one could use the same mesh for the two

242

solvers, it is concluded, however, that using different and separate meshes is

243

in general more efficient, due to the following reasons:

244

(1) The RANS mesh should be refined in the near-wall region in the wall-

245

normal direction. If a low Reynolds number turbulence model is used

246

(without using wall functions), the first grid point should be no more

247

than one wall-unit away from the walls. On the other hand, the LES

248

mesh should well resolve the free shear flow region away from the walls.

249

It does not need to resolve the near wall region, since this would be too

250

expensive; this region would be resolved by RANS mesh. This near-wall

251

region becomes the RANS region, and thus the average quantities in the

252

LES relax towards the RANS quantities. Clearly, a dual-mesh approach

253

is optimal since the different resolution requirements for RANS and LES

254

can be honored.

255

(2) For the RANS mesh, particularly in relatively simple flows or in simple

256

blocks of a complex geometry, the resolutions in streamwise and spanwise

257

directions are less important than they are in wall-normal direction, and

258

the overall required resolution only weakly depends on the Reynolds

259

number (∼ log(Re), see Ref [18]). On the contrary, the LES mesh needs

260

to resolve the streamwise and spanwise directions, and the required grid

261

spacing is proportional to the wall-units (and thus strongly depends on 12

262

the Reynolds number). The different scaling behaviors of RANS and

263

LES meshes justify different meshes in the cases mentioned above.

264

An example of such a dual-mesh is shown in Figure 3. With the dual-mesh

265

setup, when calculating the drift terms, for example QLi in Equation (10),

267

the RANS velocity hUi i needs to be interpolated from the RANS mesh to the LES mesh, and vice versa for hU¯i iAVG in Equation (15), which has to

268

be interpolated from the LES mesh to the RANS mesh for the calculation

269

of QR i . It is emphasized that the interpolated quantities are only used to

270

compute the drift terms, which are eventually divided by relaxation time

271

scales. On the other hand, the primary variables (e.g., velocity, pressure,

272

turbulent kinetic energy) are never directly manipulated or overwritten by

273

these interpolated quantities.

274

2.4. Solution Algorithm

275

Summarizing the presentation of the modeling framework, the overall algo-

276

rithm of a hybrid LES/RANS solver is presented in Algorithm 1.

277

For the computation of exponentially weighted average quantities in Equa-

278

tion (5), time integration is needed. To reduce storage and to facilitate

279

numerical computation, it is recognized that Equation (4) is the solution of

280

following equation:

281

dhφiAVG 1 = (φ − hφiAVG ) . (17) dt T Hence, the exponentially weighted averaging can be approximated up to first

282

order as [19]

266

hφiAVG, n =(1 − α)φn + αhφiAVG, n−1 1 with α = , 1 + ∆t/T

13

(18) (19)

283

where n indicates the current time level, and (1 − α) is the weight of the

284

current value φn . With this implementation, the fields from previous steps

285

earlier than (n − 1) do not need to be stored.

286

3. Implementation and Numerical Methods

287

The framework and algorithm presented in Section 2.4 do not depend on

288

specific LES and RANS models or on specific flow solvers. In this section,

289

we present an implementation of the framework and algorithm in an in-

290

compressible turbulent flow solver. This implementation is presented in a

291

separate section in order to emphasize the fact that the framework and al-

292

gorithm of the previous section are general. Similarly, one could implement

293

this algorithm in a compressible solver or in other simulation platforms. The

294

implementation details are presented in Section 3.1. The turbulence mod-

295

els and the numerical methods used for the simulations in this study are

296

discussed in Sections 3.2 and 3.3, respectively.

297

3.1. Solver Implementation

298

According to the framework and algorithm proposed above, a hybrid solver

299

for incompressible turbulent flows has been implemented based on the open

300

source CFD platform OpenFOAM [20, 21], taking advantage of its existing

301

LES and RANS solvers as well as its field operation capabilities. The finite

302

volume method with unstructured meshes is used for both LES and RANS

303

solvers, but LES and RANS may use different meshes, which is motivated

304

in Section 2.3. The PISO (Pressure Implicit with Splitting of Operators)

305

algorithm [22] is used to solve the coupled momentum and pressure equations

306

in both LES and RANS solvers. Collocated grids are used and the Rhie and

307

Chow interpolation is used to prevent the pressure–velocity decoupling [23]. 14

308

At each time step, the PISO algorithm as used in OpenFOAM can be sum-

309

marized as follows [24]:

310

(1) Momentum prediction: solve the fluid momentum equation using the

311

312 313

314

pressure field and the external forcing from the previous step. (2) Pressure solution: solve the pressure equation for the updated pressure field. (3) Velocity correction: correct the velocity field using the updated pressure

315

field with the momentum equation.

316

Repeat steps 2 and 3 until convergence.

317

Parallel processing is a built-in capability of OpenFOAM through domain

318

decomposition, where the mesh (and all the fields associated with the mesh,

319

the same hereafter) are decomposed and distributed to all processors. The

320

processors communicate with each other using MPI (Message Passing Inter-

321

face). Since a dual-mesh approach is used in this study, the LES and the

322

RANS meshes are decomposed separately, both meshes are distributed to all

323

processors. Thus, each processor holds part of the LES mesh and part of the

324

RANS mesh. For the convenience of implementation and to minimize com-

325

munication, we have ensured that the LES and RANS meshes on the same

326

processor approximately cover the same domain. This is easily achieved be-

327

cause we assume that the LES and RANS meshes cover the same domain,

328

although the meshes are different. This implementation decision is made for

329

coding convenience and computational efficiency only. It is not an essential

330

part of the algorithm, and one could certainly implement differently.

331

3.2. Turbulence Models

332

In the filtered momentum equation (2a), the residual stress term τijsgs needs

333

to be modeled. Various SGS turbulence models exist. In this study, a one-

334

equation eddy (OEE) viscosity model is used, where the equation for the 15

335

SGS kinetic energy ksgs is solved [25, 26], i.e.,   ∂ksgs ∂(ksgs U¯i ) ∂ ∂ksgs sgs ¯ + = − τij Sij + νsgs − εsgs , ∂t ∂xi ∂xi ∂xi 2 1 with τijsgs − ksgs δij = − 2νsgs (S¯ij − S¯kk δij ), 3 3 1/2 νsgs = Ck ksgs ∆ , 3/2 −1 ∆ , and εsgs = Cε ksgs

(20a) (20b) (20c) (20d)

336

where ∆ is the filter size. The rate-of-strain tensor S¯ij is defined in Equa-

337

tion (7); Ck and Cε are model constants; εsgs is the SGS dissipation rate;

338

νsgs is the subgrid scale eddy viscosity. Yoshizawa [25] observed that the

339

standard Smagorinsky model would be recovered from this model if produc-

340

tion equals dissipation, where the correspondence between model constants

341

in Equation (20) and the Smagorinsky constant is Cs = Ck /Cε . The stan-

342

dard coefficients are used, Ck = 0.094 and Cε = 1.048, which corresponds to

343

Cs = 0.167. Finally, van-Driest damping l = Cs ∆(1 − exp(−y + /25)3 )0.5 is

344

applied on the length scale near the wall.

345

Similarly, in the RANS momentum equation (3a), the Reynolds stress term

346

hui uj i also needs to be closed. The low-Reynolds number k–ε turbulence

347

model by Launder and Sharma is used [27]. Details of the model can be

348

found, for example, in reference [17]. The same standard coefficients as in

349

that reference are employed here. Since a damping function is included in

350

the modeling equations, no wall functions are used, but such low-Reynolds

351

number models require that the first grid points are located approximately

352

within one wall-unit away from the wall.

353

3.3. Numerical Methods

354

In the LES and the RANS simulations, second order central schemes are used

355

for both convective and diffusion terms. A second-order backward scheme is

3/4

16

1/4

356

used for the time integration of all variables including filtered velocities U¯i ,

357

SGS kinetic energy ksgs , RANS velocities hUi i, the turbulent kinetic energy

358

k R , and the dissipation rate εR . A multi-grid method is used to solve the

359

linear equations obtained from the finite volume discretizations.

360

4. Numerical Simulations

361

The developed hybrid LES/RANS solver as described in Section 3 is used

362

to simulate two representative cases: (1) the plane channel flow and (2) the

363

flow in a channel with periodic hills. The first case is a classical benchmark

364

case of wall-bounded flows with simple geometry. The second case features

365

a massive separation due to the wall curvature. It has been reported that

366

RANS models, even with second-order closure, have difficulties predicting

367

the mean flow profiles correctly [9]. On the other hand, properly resolved

368

LES shows much better performance.

369

Before presenting the results, the general issues are explained in Section 4.1,

370

including guidelines for choosing mesh resolutions, the objectives of the sim-

371

ulations, and the methods used to assess the results. The simulation results

372

are shown in Sections 4.2 and 4.3. Finally, the sensitivity of the results on

373

the algorithmic parameters is studied in Section 4.4.

374

4.1. General remarks on the simulations

375

In this study, the LES meshes are designed to be adequate to resolve the

376

free-shear region, but with little refinement near the wall. As a result, in the

377

near-wall region the mesh is coarse in all directions in terms of wall-units. It is

378

expected that the mesh in the near-wall region is at least fine enough to allow

379

for enough fluctuations, if forced by the term QL,g as in Equation (12) in order i

380

to achieve consistency with the TKE in RANS. On the other hand, the RANS 17

381

meshes are designed such that the first grid point is located approximately

382

one wall unit from the wall. The grids in other directions, particularly for

383

the homogeneous directions, are very coarse. The LES and RANS meshes for

384

the test case are shown in Figure 3. More discussion on grid spacing follows

385

in Section 2.3.

386

Compared to the fine meshes used in previous studies, the LES mesh adopted

387

here is very coarse. For example, for the periodic hill case of Re=10595, ap-

388

proximately five million cells are used for a properly revolved LES and one

389

million cells are used for DES (see reference [9]). Simulations presented here

390

employ a grid with less than 105 cells. However, it is fine enough to resolve the

391

free shear region. The purpose of this study is not to obtain the best results

392

possible or to provide benchmark solutions. Instead, we intend to demon-

393

strate that the proposed method can lead to improved results compared to

394

standalone LES on the relatively coarse grids with moderate computational

395

overhead. Therefore, it is reasonable to assess the performance of the hy-

396

brid method by examining quality of the results against those obtained with

397

standalone LES on the same mesh.

398

In addition, particular emphasis is placed on the internal consistency of the

399

LES and RANS solutions, which forms the foundation of the proposed frame-

400

work. For the case presented here, it is challenging due to the transient co-

401

herent structures, and thus smaller relaxation time scales are used. Detailed

402

consistency study is presented.

403

As demonstrated by the studies of Davidson [28, 29], evaluating whether the

404

resolution of an LES is adequate is a challenging task. Considering the fact

405

that determining appropriate LES/RANS regions is a separate issue from

406

the coupling algorithm itself, for simplicity, pre-specified LES/RANS regions

407

according to the cell distances to the nearest wall are used for the simulations

408

presented here. This issue is further discussed in Section 5.2. 18

409

Compared to the LES on the same grid, the hybrid method also needs to

410

compute the drift terms and to solve the RANS equations additionally (usu-

411

ally on a coarser grid). The overhead due to these operations is approxi-

412

mately 30% to 50% of the total computational cost for the simulations pre-

413

sented here. However, since the RANS meshes are almost independent of the

414

Reynolds number, the fraction of the overhead becomes negligible for high

415

Reynolds number flows encountered in practice. Future development will

416

consider larger time-steps for the RANS solver (compared to the small time-

417

steps in LES), and the computational overhead compared to standalone LES

418

would be further reduced. For high Reynolds number flows, theoretically the

419

computational cost of the hybrid method should depend on the Reynolds

420

number in the same way as LES does for free-shear flows, probably with a

421

slightly larger proportionality constant.

422

4.2. Flow in a plane channel

423

In this case, the fully developed turbulent flow in a plane channel is simulated

424

using the hybrid method. The domain size, meshes, and resolutions are pre-

426

sented in Table 1. The nominal Reynolds number based on friction velocity p uτ and half channel width δ is Reτ = uτ δ/ν = 395, where uτ = τw /ρ, and

427

τw is the wall shear stress. The actual Reτ is a result of the computation.

428

The streamwise (x) and the spanwise (z) directions have periodic boundary

429

conditions. Non-slip boundary conditions are applied at the wall (and y

430

coordinate is aligned with the wall-normal direction). A pressure gradient is

431

applied on the whole domain to keep the mass flux constant in x direction.

432

The flow-through time is thus defined as Tthr = 2πδ/Ub , and the algorithmic

433

parameters are: T = 11δ/Ub , τl = 1.3δ/Ub , τr = 1.3δ/Ub , and τg = 0.67δ/Ub .

434

The RANS region consists of all cells with distance smaller than D = 0.2δ to

435

the nearest wall. To achieve stable coupling, a linear ramp function F (t) is

425

19

436

multiplied on all drift terms during an initial simulation period of 2T , where

437

F (0) = 0 and F (t ≥ 2T ) = 1.

438

The computed Reτ , which is indicative of the prediction of wall shear stress,

439

is presented in Table 2. It can be seen that the predicted wall shear stress is

440

significantly improved in the hybrid method compared to the standalone LES

441

result on the same mesh (The LES mesh is intentionally made uniform in

442

all directions without attempting to resolve the wall). The wall shear stress

443

prediction is improved because the averaged velocity relaxes towards the

444

RANS velocity in the near-wall region. The mean velocity obtained through

445

averaging in time and spanwise/streamwise directions are shown in Figure 1.

446

The improvement over LES is appreciable.

447

4.3. Flow in a channel with periodic hills

448

The benchmark case of the flow in a channel with periodic hills is based

449

on the original experiment by Almeida et al. [30]. Recently this case has

450

been modified to make it more convenient for numerical simulations. New

451

experiments and benchmark numerical simulations (including LES and DNS)

452

have been conducted for a wide range of Reynolds numbers within the French-

453

German research group on Large-Eddy Simulation of Complex Flows [9, 31].

454

In this work, the results from the high Reynolds number case Re = 10595 is

455

presented.

456

The geometry of the computational domain is shown in Figure 2 and spec-

457

ifications are given in Table 1. The exact shape of the hill is described by

458

piecewise continuous polynomials [32]. Similar to the case above, periodic

459

boundary conditions are used in the streamwise and spanwise directions, and

460

a pressure gradient is applied on the whole domain to keep the mass flux con-

461

stant. The Reynolds number based on crest height and bulk velocity (at the

462

crest) is 10595. 20

463

Meshes with resolutions, simulation time-span, and time step size are shown

464

in Table 1. In this case, all lengths and times are normalized with H and

465

H/Ub , respectively. The flow-through time is defined as Tthr = Lx /Ub =

466

9H/Ub , and the algorithmic parameters are: T = 2.2H/Ub , τl = 0.28H/Ub ,

467

τr = 0.28H/Ub , and τg = 0.07H/Ub . The RANS region consists of all cells

468

with distance smaller than D = 0.2H from the nearest wall. The same linear

469

ramp function as above is applied.

471

The internal consistency between LES and RANS is investigated. Time series of the filtered velocity component U¯1 , the exponentially weighted average

472

velocity component hU¯1 iAVG , and the RANS velocity component hU1 iAVG are

473

depicted in Figure 4 at four different locations. The same comparison for k,

474

hτii /2iAVG , and k R is presented in Figure 5 at the same four locations.

475

From Figures 4 and 5, it can be observed that the filtered signals fluctuate

476

with high frequencies, while the averaged LES and the RANS quantities

477

are much smoother. The internal consistency, i.e., the agreement between

478

exponentially weighted average quantities and RANS quantities, is achieved

479

reasonably well. The fluctuations of RANS quantities are still smaller than

480

the EWA quantities (which are obtained by averaging the LES quantities;

481

see Equation (5)). This is true even when a much larger averaging time-

482

scale is used and is due to the very coarse RANS mesh, and probably more

483

important, due to the high dissipation from the RANS turbulence model. The

484

differences in the amount of fluctuations that can be sustained by RANS and

485

LES lead to some transient inconsistencies, e.g., t/Tthr = 14 at points (b) and

486

(c), but the durations are rather short. Since the consistency is enforced by

487

relaxation models, if so desired, the consistency can be further improved by

488

decreasing relaxation time scales. As can be seen from Figure 6(a), the mean

489

streamwise velocity profiles of the RANS and those of the LES agree very

490

well at all locations. For comparison, the same velocity profiles obtained by

491

standalone LES and RANS solvers are shown in Figure 6(b), which displays

470

21

492

a much larger discrepancy. This comparison confirms the effectiveness of the

493

drift terms.

494

In Figure 7(a), the mean streamwise velocities (time- and spanwise-averaged

495

filtered velocity) are presented at nine locations (from x/H = 0 to 8, with an

496

interval of 1); plotted at the corresponding locations in the domain. The hy-

497

brid results are compared with the standalone LES (the same grid is used as

498

for the hybrid LES solver) and the benchmark solutions [31]. The improve-

499

ment of the hybrid results compared to LES is significant, particularly from

500

y/H = 1.8 to y/H = 2.8 and at the bottom of the channel. The shaded re-

501

gion covering the recirculation and reattachment regions is enlarged to show

502

more details in Figure 7(b). The improvement in the near-wall region is due

503

to the good wall-resolution of RANS mesh, which in turn provides guidance

504

to the LES mean velocity via the drift term. It is noted that the standalone

505

RANS performs poorly in this region, partly because the spanwise velocity

506

fluctuations are not well represented in RANS [9]. The RANS results are

507

shown in Figure 4(b). From our experience with this study, in pure RANS

508

simulations the fluctuations gradually damp out and the solutions eventually

509

approach steady state after a few flow-through times. However, when cou-

510

pled with the LES solver, the hybrid RANS solutions are made consistent

511

with the exponentially weighted average filtered velocities (in the LES re-

512

gion). Consequently, the hybrid RANS simulations are unsteady in general.

513

This explains why the RANS solver predictions are poor, but that RANS

514

can still help to improve the LES results in the hybrid solver.

515

To further illustrate the improvements of the hybrid simulation results over

516

the pure LES, the mean turbulent kinetic energy are presented in Figure 8

517

for nine locations along the channel (same as above). It is evident from Fig-

518

ure 8(a) that the mean TKE in the free-shear region (between y/H = 0.5

519

and y/H = 2) is better predicted by the hybrid solver, particularly down-

520

stream of x/H = 3. The shaded region is zoomed in and highlighted in 22

521

Figure 8(b), which suggest that the improvement in the near-wall region is

522

equally good. The comparison of turbulent shear stress component u0 v 0 is

523

shown in Figure 9 for four locations (x/H = 1, 3, 5, 7). The hybrid simulation

524

results again show better agreement with the benchmark solution than the

525

pure LES results.

526

The friction coefficient Cf at the bottom wall predicted by standalone LES

527

and RANS are compared to benchmark results in Figure 10(a). The hybrid

528

solutions are presented in Figure 10(b), which show significant improvement

529

in the Cf prediction, compared to standalone LES and RANS. It can be seen

530

that in the whole region the prediction of Cf by pure LES is too low in mag-

531

nitude compared to the reference results, while RANS simulation predicts a

532

wrong reattachment point and thus a different overall behavior. In summary,

533

compared to the standalone LES and RANS solvers, the hybrid method leads

534

to improved wall-shear stress predictions at the bottom wall, which can be

535

explained similarly as above.

536

4.4. Parameter choices and sensitivity study

537

In this study, we have three groups of algorithmic parameters: (1) the lo-

538

cation D of the LES/RANS interface; (2) the averaging time scale T ; and

539

(3) the relaxation time scales for the average filtered velocities (τl ), for the

540

RANS velocity (τr ), and for the fluctuations in the filtered velocity (τg ).

541

In the future LES and RANS regions shall be detected dynamically based on

542

cell quantities, and thus this free parameter eventually will be eliminated. In

543

addition, clear guidelines have been proposed for the choice of this parameter

544

in the hybrid methods such as DES [33]. Therefore, choosing this parameter

545

is not of concern here, although studies of the sensitivity on this parameter

546

are presented.

23

547

The averaging time scale T and the relaxation time scales τl , τl , and τg are

548

specific to this framework, and will thus be discussed in more detail. The

549

purpose of the averaging on filtered quantities is to obtain appropriate quan-

550

tities that can be correctly made consistent with the corresponding RANS

551

quantities (since the filtered quantities in LES are fundamentally different

552

from the RANS quantities). Therefore, the averaging should smooth out all

553

the small-scale turbulent structures, while keeping the coherent structure.

554

Consequently, the time scale T should be larger than the turnover time of

555

most eddies, but smaller than the time scale of the large coherent structures.

556

Based on the same reasoning, the time scales τl and τr should be fractions

557

of T (e.g., between 1/4 and 1/20). Values of τl and τr comparable to or

558

even larger than T may lead to large internal inconsistencies. On the other

559

hand, using too small relaxation time scales (for example, two or more orders

560

of magnitude smaller) may cause convergence difficulties and deteriorated

561

results. The time scale τg is related to the QL,g term, which adjusts the i

562

fluctuation levels of the filter velocity to achieve consistency with the modeled

563

turbulent stresses in the RANS calculations. The time scale of these actions

564

has to be comparable to the smallest resolved eddies. Hence, τg should be

565

smaller than τl and τr . Using a too large τg would lead to large inconsistencies

566

in the Reynolds stresses or TKE, while a too small τg may cause convergence

567

difficulties. From our experience, if the time scales are chosen based on the

568

reasoning discussed above, good results are usually observed.

569

Parametric studies are conducted to assess the influence of the averaging

570

and relaxation time-scales and of the LES/RANS interface location on the

571

computational results. The simulation presented in Section 4.3 is considered

572

as base case and the parameters are varied as follows:

573

(a) the relaxation time scales in the base case are scaled by factors of 0.75,

574 575

1.5, and 2; (b) the averaging time scale is varied, i.e., T Ub /H is set to 1.7, 2.8, and 4.5; 24

576

577 578

and (c) the LES/RANS interface location is placed at 0.1H, 0.15H, and 0.25H away from the walls.

579

Simulations are conducted with the varied parameters (nine cases in total, in

580

addition to the base case), and the results are presented in Figures 11(a)–(c),

581

respectively, and are summarized in Figure 11(d). In all the plots, the results

582

from the standalone LES are shown as a reference to quantify the variations of

583

the results due to perturbed parameters. It can be seen that the results seem

584

to be relatively more sensitively to the perturbation of LES/RANS location

585

D, as shown in Figure 11(c). However, overall the differences between the

586

results with the perturbed parameters are small compared to the difference

587

to the standalone LES results.

588

Another parameter of interest, although not intrinsic to the hybrid frame-

589

work, is the Smagorinsky constant Cs . The sensitivity of the hybrid simula-

590

tion results to the perturbation of Cs is studied. In Figure 12, the mean

591

velocity at four locations, (x/H = 1, 3, 5, 7) with standard and reduced

592

Smagorinsky constants (Cs = 0.167 and 0.1, respectively) are compared.

593

The results obtained from pure LES with Cs = 0.167 are also presented as

594

above. It can be seen that differences due to perturbed Cs are relatively small

595

compared to the differences between hybrid simulations results and pure LES

596

results. This observation further demonstrates the robustness of the hybrid

597

framework.

598

5. Discussions

599

In this study, we propose a novel hybrid framework and algorithm for tur-

600

bulent flow simulations. At this stage, we focus on the most fundamental

601

aspects of the method to investigate and demonstrate its feasibility and po25

602

tential. Other non-essential features, possible improvements, variants, and

603

extensions of the algorithm are left for future research, but are briefly dis-

604

cussed in this section.

605

5.1. Using other turbulence models

606

As emphasized in Section 3, the hybrid solver used in this study is one pos-

607

sible implementation. Considering the purpose of this study as mentioned

608

above, we choose to use simple LES and RANS models, which are not nec-

609

essarily state-of-the-art according to the most recent literature. By using

610

modern subgrid-scale models for LES (e.g., dynamic models, scale-similarity

611

models, or mixed SGS models) and/or RANS models with better perfor-

612

mance near the wall (e.g., k–ω [17] or Reynolds stress models with elliptic

613

relaxation [34]), better performance can be expected from the hybrid method.

614

Other models will be implemented and investigated in future studies.

615

5.2. Dynamic evaluation of resolution

616

Although in the current study the LES/RANS regions are pre-specified, the

617

framework in principle can accommodate for cell-based dynamic adaptation

618

of these regions. The main difficulty, however, is a general and reliable cri-

619

terion. Davidson conducted extensive studies on the evaluation of appropri-

620

ate resolution for LES of simple wall-bounded flows [28] and recirculating

621

flows [29]. It is concluded that none of the criteria found in the literature

622

consistently give reliable results and that the two-point correlations are con-

623

sidered to be the most reliable. However, the two-point correlations criterion

624

is not practical in the context of hybrid LES/RANS simulations because res-

625

olution evaluations need to be performed on the fly and based on individual

626

cell quantities. In addition, the numerical dissipation, which is significant

26

627

in solvers using low-order schemes [35, 36], further complicates the prob-

628

lem. This issue is further investigated and the results will be published in a

629

separate work.

630

6. Conclusion

631

In this work, we propose a novel consistent framework for hybrid LES/RANS

632

modeling, where the filtered and RANS equation are solved simultaneously in

633

the whole domain on their respective meshes. Consistency between the LES

634

and RANS solutions is enforced via drift terms in the corresponding equa-

635

tions. This framework leads to very clean conditions at the LES/RANS in-

636

terfaces and allows for individual cell-based determination of LES and RANS

637

regions.

638

As a specific implementation, a hybrid solver has been developed according

639

to the proposed framework and algorithm using the open-source CFD plat-

640

form OpenFOAM. The developed solver is used to simulate a representative

641

case of the flow in a plane channel and that in a channel with periodic hills.

642

Results demonstrate that internal consistency is honored faithfully and that

643

the improvements over standalone LES on the same grid are appreciable.

644

Parametric studies suggest that the sensitivity of the results on algorith-

645

mic parameters is minor. Therefore, the proposed framework is a promising

646

candidate for hybrid LES/RANS simulations.

647

Further improvements are likely by using dynamically adjusted averaging

648

time scales, automatically detected LES/RANS regions based on cell quan-

649

tities and more advanced turbulence models. This, however, is subject of

650

future investigations.

27

651

Acknowledgement

652

We acknowledge the financial support from the Swiss Commission for Tech-

653

nology and Innovation (CTI), and computational resources provided by ETH

654

Z¨ urich. We thank Dr. M. Breuer at the University of Erlangen-Nuremberg

655

for providing us the benchmark data for comparison. Helpful discussions

656

with Prof. L. Kleiser at ETH Z¨ urich are appreciated. The first author would

657

like to thank Michael Wild for the fruitful discussions on numerous occasions

658

throughout this work. His technical advice during the development of the

659

hybrid solver is gratefully acknowledged.

660

References

661 662

663 664

[1] D. R. Chapman, Computational aerodynamics: development and outlook, AIAA Journal 17 (1979) 1293–313. [2] U. Piomelli, E. Balaras, Wall-layer models for large-eddy simulations, Annual Review of Fluid Mechanics 34 (2002) 349–374.

665

[3] J. Fr¨ohlich, D. von Terzi, Hybrid LES/RANS methods for the simulation

666

of turbulent flows, Progress in Aerospace Sciences 44 (5) (2008) 349–377.

667

[4] M. Germano, Properties of the hybrid RANS/LES filter, Theoretical

668

669 670

and Computational Fluid Dynamics 17 (2004) 225–331. [5] B. Rajamani, J. Kim, A hybrid-filter approach to turbulence simulation, Flow Turbulence Combustion 85 (2010) 421–441.

671

[6] J. C. Uribe, N. Jarrin, R. Prosser, D. Laurence, Development of a two-

672

velocities hybrid RANS-LES model and its application to a trailing edge

673

flow, Flow Turbulence Combustion 85 (2) (2010) 181–197.

28

674

[7] P. R. Spalart, W. H. Jou, M. Strelets, S. R. Allmaras, Comments

675

on the feasibility of LES for wings, and on a hybrid RANS/LES ap-

676

proach, in: First AFOSR International Conference on DNS/LES, Rus-

677

ton, Louisiana, 1997.

678

[8] P. Spalart, S. Deck, M. Shur, K. Squires, M. Strelets, A. Travin, A new

679

version of detached-eddy simulation, resistant to ambiguous grid densi-

680

ties, Theoretical and Computational Fluid Dynamics 20 (2006) 181–195.

681

[9] S. Saric, S. Jakirlic, M. Breuer, B. Jaffrezic, G. Deng, O. Chikhaoui,

682

J. Fr¨ohlich, D. von Terzi, N. Manhart, M. andPeller, Evaluation

683

of detached eddy simulations for predicting the flow over periodic

684

hills, in:

685

10.1051/proc:2007016.

ESAIM: Proceedings, Vol. 16, 2007, pp. 133–145, doi:

686

[10] J. S. Baggett, On the feasibility of merging LES with RANS for the

687

near-wall region of attached turbulent flows, in: Annual Research Briefs,

688

Center for Turbulence Research, 1998, pp. 267–277.

689

[11] U. Piomelli, E. Balaras, H. Pasinato, S. K. D., P. Spalart, The inner-

690

outer layer interface in large-eddy simulations with wall-layer models,

691

International Journal of Heat and Fluid Flow 24 (2003) 538–550.

692

[12] L. Davidson, Hybrid LES-RANS: An approach to make LES applicable

693

at high Reynolds number, International Journal of Computational Fluid

694

Dynamics 19 (6) (2005) 415–427.

695

[13] L. Davidson, Hybrid LES-RANS: back scatter from a scale-similarity

696

model used as forcing, Philosophical Transactions of the Royal Society

697

A 367 (2009) 2905–2915.

698

[14] R. Laraufie, S. Deck, P. Sagaut, A dynamic forcing method for un-

699

steady turbulent inflow conditions, Journal of Computational Physics

700

230 (2011) 8647–8663. 29

701

[15] P. J. Morris, L. N. Long, A. Bangalore, Q. Wang, A parallel three-

702

dimensional computational aeroacoustics method using nonlinear dis-

703

turbance equations, Journal of Computational Physics 133 (1997) 56–

704

74.

705

[16] E. Labourasse, P. Sagaut, Reconstruction of turbulent fluctuations using

706

a hybrid RANS/LES approach, Journal of Computational Physics 182

707

(2002) 301–336.

708 709

710 711

712 713

714 715

[17] D. C. Wilcox, Turbulence Modeling for CFD, 2nd Edition, DCW Industries, 1998, Ch. 4, pp. 138–141. [18] K. Hanjalic, Will RANS survive LES? A view of perspectives, Journal of Fluids Engineering 127 (2005) 831–839. [19] T. S. L. C. Meneveau, W. H. Cabot, A lagrangian dynamic subgrid-scale model of turbulence, Journal of Fluid Mechanics 319 (1996) 353–385. [20] OpenCFD Ltd., The open source CFD toolbox. URL www.openfoam.com

716

[21] H. G. Weller, G. Tabor, H. Jasak, C. Fureby, A tensorial approach to

717

computational continuum mechanics using object-oriented techniques,

718

Computers in Physics 12 (6) (1998) 620–631. doi:10.1063/1.168744.

719 720

[22] R. I. Issa, Solution of the implicitly discretised fluid flow equations by operator-splitting, Journal of Computational Physics 62 (1986) 40–65.

721

[23] C. M. Rhie, W. L. Chow, A numerical study of the turbulent flow past

722

an isolated airfoil with trailing edge separation, AIAA 21 (11) (1983)

723

1525–1532.

724 725

[24] J. H. Ferziger, M. Peric, Computational Methods for Fluid Dynamics, 3rd Edition, Springer, New York, 2001.

30

726

[25] A. Yoshizawa, K. Horiuti, A statistically-derived subgrid-scale kinetic

727

energy model for the large-eddy-simulation of turbulent flows, Journal

728

of Physical Society of Japan 54 (1985) 2834–2839.

729

[26] C. Fureby, G. Tabor, H. G. Weller, A. D. Gosman, Comparative study

730

of subgrid scale models in homogeneous isotropic turbulence, Physics of

731

Fluids 9 (5).

732

[27] B. E. Launder, B. I. Sharma, Application of the energy-dissipation

733

model of turbulence to the calculation of flow near a spinning disc, Let-

734

ters in Heat and Mass Transfer 1 (2) (1974) 131–138.

735 736

[28] L. Davidson, Large eddy simulations: How to evaluate resolution, International Journal of Heat and Fluid Flow 30 (2009) 1016–1025.

737

[29] L. Davidson, How to estimate the resolution of an LES of recirculating

738

flow, in: Quality and Reliability of Large-Eddy Simulations II, Springer,

739

2010, pp. 269–286.

740

[30] G. P. Almeida, D. F. G. Durao, M. V. Heitor, Wake flows behind two-

741

dimensional model hills, Experimental Thermal and Fluid Science 7 (1)

742

(1993) 87–101.

743

[31] M. Breuer, N. Peller, C. Rapp, M. Manhart, Flow over periodic hills:

744

Numerical and experimental study in a wide range of reynolds numbers,

745

Computers & Fluids 38 (2) (2009) 433–457.

746 747

748 749

750 751

[32] L. Temmerman, M. A. Leschziner, Flow over 2D periodic hills, http://cfd.mace.manchester.ac.uk/twiki/bin/view/CfdTm/TestCase014. [33] P. R. Spalart, Young person’s guide to detached-eddy simulation grids, Tech. Rep. CR-2001-211032, NASA (2001). [34] S. B. Pope, Turbulent Flows, Cambridge University Press, , Cambridge, 2000. 31

752

[35] Y. Addad, D. Laurence, C. Talotte, M. C. Jacob, Large eddy simula-

753

tion of a forward-backward facing step for acoustic source identification,

754

International Journal of Heat and Fluid Flow 24 (4) (2003) 562–571.

755 756

[36] I. B. Celik, Z. N. Cehreli, I. Yavuz, Index of resolution quality for large eddy simulations, Journal of Fluids Engineering (2005) 949–958.

757

[37] R. D. Moser, J. D. Kim, N. N. Mansour, Direct numerical simulation

758

of turbulent channel flow up to Reτ =590, Physics of Fluids 11 (1999)

759

943–945.

32

1. Choose LES and RANS models and other parameters ; 2. Initialize the fields for LES and RANS ; 3. Initialize the fields for Exponentially Weighted Average (EWA) quantities ; 4. Initialize all drift terms to zeros ; for each time step do 1. For each grid cell determine whether it is well resolved (in LES), and assign it to LES or RANS region accordingly ; 2. Solve filtered momentum and pressure Poisson equations (2) for U¯i and p¯ ; 3. Solve averaged momentum and pressure Poisson equations (3) for hUi i and hpi ; 4. Update EWA quantities hU¯i iAVG , hτij iAVG and hεiAVG , according to Equation (5) ; 5. Interpolate the quantities needed for the calculation of drift terms ; 6. Update drift terms according to Equations (10), (15), and (16) ; end Algorithm 1: Overall algorithm of the hybrid solver as implemented in this work.

33

Table 1: Domain and mesh parameters for the test case. x, y, z are aligned with streamwise, wall-normal, and spanwise directions, respectively.

cases

plane channel

periodic hill

domain size (Lx × Ly × Lz )

2πδ × 2δ × πδ

9H × 3.036H × 4.5H

simulation time-span

800(≈ 120Tthr )

460Ub /H (≈ 50 Tthr )

Nx × Ny × Nz (LES)

50 × 60 × 30

74 × 37 × 36

Nx × Ny × Nz (RANS)

10 × 80 × 10

74 × 37 × 18

∆x × ∆y × ∆z in y + (LES) 50 × 12 × 41

76 × [30, 72] × 78

(1)

first grid point (RANS)

0.65y +

below 2y + at most areas

time-step size

6.68 × 10−3 Ub /δ

2.8 × 10−3 Ub /H

(2)

(1)Numbers in the brackets indicate the range of ∆y (smallest for the cells next to the wall and largest for those at the center line). (2) The wall unit is defined p as y + = ν/uτ = ν/ τ /ρ, where τ is the shear stress. Table 2: Comparison of computed Reτ (Reynolds number based on friction velocity and half channel height), which is indicative of the wall shear stress predictions.

Reτ nominal

395

DNS (reference [37])

392

pure LES

339

hybrid LES

371

34

25 20 U/uτ

15 10

hybrid LES pure LES DNS (Moser et al. 1999)

5 00.0

0.2

0.4

x/δ

0.6

0.8

1.0

Figure 1: Mean velocity profile of the hybrid method compared to the LES results on the same mesh and with DNS data [37]. The LES mesh is uniform in all directions with mesh density (in y + ) presented in Table 1.

Lx general flow direction

Ly y z

H x

Figure 2: Domain shape for the flow in the channel with periodic hills. The square and the circle denote the approximate locations of the separation and reattachment, respectively. x, y, and z coordinates are aligned with the streamwise, wall-normal, and spanwise directions, respectively. The dimensions of the domain are: Lx = 9H, Ly = 3.036H, and Lz = 4.5H.

35

Figure 3: The meshes used for (a) the LES and (b) the RANS in the simulation of the flow over periodic hills. The LES mesh is designed to resolve the free-shear region and with little stretching towards the wall. The RANS mesh is refined in the near-wall region in the wall-normal direction.

36

Normalized velocity

1.2 1.1

(a) filtered velocity consistent velocity RANS velocity

(b) 1.2

1.0 0.9 0.8 0.7 0.6 0.5 0

5

10

t/T0

4 3 2 1 0 1 0 2 4 6 8 10 15 20

Normalized velocity

1.3

0.6 0.4 0

5

10

t/T0

4 3 2 1 0 1 0 2 4 6 8 10 15 20

(d) 0.6

0.0 0.1 0.2 0.3 0.4 0.5 5

10

t/T0

4 3 2 1 0 1 0 2 4 6 8 10 15 20

Normalized velocity

0.1 Normalized velocity

0.8

(c)

0.2

0.6 0

1.0

0.4 0.2 0.0 0.2 0

5

10

t/T0

4 3 2 1 0 1 0 2 4 6 8 10 15 20

Figure 4: Time series of filtered velocity, exponentially weighted average (EWA) velocity, and RANS velocity (all normalized by Ub ). The time series are presented for four locations: (a) x/H = 3, y/H = 1.9; (b) x/H = 7, y/H = 1.9; (c) x/H = 3, y/H = 0.1; and (d) x/H = 7, y/H = 1.9. The spanwise coordinates are z/H = 2 for all points. Plots (a) and (b) correspond to points located in the LES region, and plots (c) and (d) correspond to points located in the RANS region.

37

(a)

Normalized TKE

0.06

(b) 4 3 2 1 0 1 0 2 4 6 8 10

total TKE EWA TKE RANS TKE

0.05 0.04 0.03 0.02

0.06 0.04 0.02

0.01 0

4 3 2 1 0 1 0 2 4 6 8 10

0.08 Normalized TKE

0.07

5

10

t/T0

15

0.00 0

20

5

(c)

0.06 0.05 0.04 0.03

0.05 0.04 0.03 0.02

0.02 0.01 0

20

4 3 2 1 0 1 0 2 4 6 8 10

0.06

Normalized TKE

Normalized TKE

0.07

15

(d) 4 3 2 1 0 1 0 2 4 6 8 10

0.08

10

t/T0

5

10

t/T0

15

0.01 0

20

5

10

t/T0

15

20

Figure 5: Time series of total turbulent kinetic energy in LES (i.e., τii /2; see Equation (6)), exponentially weighted average (EWA) TKE hτii /2iAVG , and RANS turbulent kinetic energy k R . The coordinates of the four points are the same as in Figure 4.

38

y/H y/H

3.0 2.5 2.0 1.5 1.0 0.5 0.0 0

3.0 2.5 2.0 1.5 1.0 0.5 0.0 0

(a)

LES (hybrid)

RANS (hybrid)

outline of domain

2

4

LES (standalone)

6

x/H; 3U/Ub + x/H

(b)

8

outline of domain 10

12

RANS (standalone)

outline of domain

2

4

6

x/H; 3U/Ub + x/H

8

outline of domain 10

12

Figure 6: Consistency between the LES and the RANS: mean streamwise velocity from the LES and that from the RANS are compared. (a) LES and RANS velocities from the hybrid solver with coupling; (b) LES and RANS velocities from standalone LES and RANS solvers without coupling. For the LES results, the lines pass through all data points, but markers are only shown for every seventh points for clarity.

39

y/H

3.0 2.5 2.0 1.5 1.0 0.5 0.0 0

(a) outline of domain

4

2

6

DNS (Breuer et al.)

8

outline of domain 10

pure LES

12

hybrid

(b)

1.0 0.8 y/H

0.6 0.4 0.2 0.01

2

3

4

5

6

x/H; 3U/Ub + x/H

7

8

9

Figure 7: Comparison of mean streamwise velocities between the results from the benchmark simulations [31], the standalone LES, and the current hybrid method. The velocities are all normalized by the bulk velocity Ub (at the crest), presented for all locations between x/H = 0–8 with an interval of 1. Plot (a) shows the velocities plotted on the shape of the domain, at the corresponding locations. Plot (b) is the enlargement of the shaded region in plot (a).

40

(a)

3.0 2.5

y/H

2.0 1.5 1.0 0.5 0.0

0

2

x/H;

Breuer et al.

4

−20k/Ub2

+ x/H

6

8

pure LES

hybrid

(b)

3.0 2.9 y/H

2.8 2.7 2.6 2.5 2.43

4

5

6

−40k/Ub2

+ x/H

7

8

9

Figure 8: Comparison of mean turbulent kinetic energy (TKE) between the results from the benchmark solution [31], the standalone LES, and the current hybrid method. The energy is normalized by Ub2 , presented for all locations between x/H = 0–8 with an interval of 1. Plot (a) shows the TKE plotted on the shape of the domain, at the corresponding locations. Plot (b) is the enlargement of the shaded region in plot (a).

41

(a)

3.0 2.5

y/H

2.0 1.5 1.0 0.5 0.00

2

Breuer et al.

x/H;

4

­

−90

®

6

u0 v0 /Ub2 + x/H

pure LES

8

hybrid

Figure 9: Comparison of shear stress between the results from the benchmark solution [31], the standalone LES, and the current hybrid method. The shear stresses are normalized by Ub2 , presented for four locations (x/H = 1, 3, 5, 7).

42

(a) DNS (Breuer et al. 2009) pure LES Pure RANS

0.03

Cf

0.02 0.01 0.00 0.010

1

3

4

x/H

5

6

7

8

9

6

7

8

9

(b) DNS (Breuer et al. 2009) hybrid

0.03 0.02 Cf

2

0.01 0.00 0.010

1

2

3

4

x/H

5

Figure 10: Comparison of friction coefficients Cf = 2τw /(ρUb2 ) on the bottom wall, obtained from the benchmark simulation [31], pure LES, pure RANS, and the current hybrid method. Pure LES/RANS results are presented in separate plots from the hybrid results for clarity. (a) Friction coefficient obtained using pure LES and pure RANS, compared to benchmark results. (b) Friction coefficient obtained from the hybrid method. The RANS velocities are used for the calculation of wall shear stress since it is based on a finer near-wall mesh.

43

3.0 2.5 2.0 1.5 1.0 0.5 0.0

(a)

10

3.0 2.5 2.0 1.5 1.0 0.5 0.0

y/H

10

3.0 2.5 2.0 1.5 1.0 0.5 0.0

2

0

2

4

6

x/H; 3U/Ub + x/H

8

(c) y/H

y/H y/H

3.0 2.5 2.0 1.5 1.0 0.5 0.0

2

0

2

4

6

x/H; 3U/Ub + x/H

8

(b)

2

0

4

2

6

8

10

6

8

10

x/H; 3U/Ub + x/H

(d)

2

0

2

4

x/H; 3U/Ub + x/H

Figure 11: Parameter sensitivity study: mean streamwise velocity of the base case compared to the cases with perturbed algorithmic parameters. Legend:

pure LES; solid

lines of various colors (or gray scales): hybrid LES in the base case and in those with perturbed parameters. The plots show the scattering of the velocity profiles with (a) the relaxation times scaled by factors of 0.75, 1.5, and 2; (b) varied averaging time scales, T Ub /H = 1.7, 2.8, and 4.5; and (c) varied LES/RANS interface locations: D/H = 0.1, 0.15, and 0.25 away from the walls. Plot (d) summarizes all the profiles in (a)–(c). In each plot, a part of the profile for one location is enlarged. The base case result is shown on all the plots.

44

3.0 2.5

y/H

2.0 1.5

Pure LES Hybrid Hybrid, CS =0.1

1.0 0.5 0.00

2

4

6

6U/Ub + x/H

8

10

12

Figure 12: The results obtained from the current hybrid method using standard Smagorinsky constant (Cs = 0.167) and reduced value Cs = 0.1 are compared with the standalone LES results.

45

Suggest Documents