Fully resolved scalar transport for high Prandtl number flows using adaptive mesh refinement

Fully resolved scalar transport for high Prandtl number flows using adaptive mesh refinement

Journal Pre-proofs Fully Resolved Scalar Transport for High Prandtl Number Flows using Adaptive Mesh Refinement A. Panda, E.A.J.F. Peters, M.W. Baltus...

10MB Sizes 0 Downloads 57 Views

Journal Pre-proofs Fully Resolved Scalar Transport for High Prandtl Number Flows using Adaptive Mesh Refinement A. Panda, E.A.J.F. Peters, M.W. Baltussen, J.A.M. Kuipers PII: DOI: Reference:

S2590-1400(19)30054-1 https://doi.org/10.1016/j.cesx.2019.100047 CESX 100047

To appear in:

Chemical Engineering Science: X

Received Date: Revised Date: Accepted Date:

2 August 2019 9 October 2019 14 October 2019

Please cite this article as: A. Panda, E.A.J.F. Peters, M.W. Baltussen, J.A.M. Kuipers, Fully Resolved Scalar Transport for High Prandtl Number Flows using Adaptive Mesh Refinement, Chemical Engineering Science: X (2019), doi: https://doi.org/10.1016/j.cesx.2019.100047

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.

Fully Resolved Scalar Transport for High Prandtl Number Flows using Adaptive Mesh Refinement A. Pandaa , E.A.J.F. Petersa,∗, M.W. Baltussena , J.A.M. Kuipersa a Multiphase Reactors Group, Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands.

Abstract In multiphase systems, boundary layers occur at fluid-fluid or fluid-solid interfaces. In a direct numerical simulation, the grid requirements are often dictated by the thickness of these boundary layers. Systems that are characterized by high Prandtl (or Schmidt number) exhibit temperature (or mass) boundary layers that are much thinner than the momentum boundary layers. In this paper, a hybrid computational approach is presented that uses a fixed Cartesian grid for the Navier-Stokes and continuity equations and an adaptive mesh for scalar transport, thus reducing the memory and CPU requirements tremendously while resolving all boundary layers. We describe the key aspects that need to be addressed in this hybrid approach, related to discretization, grid mapping, velocity interpolation along with detailed verification tests. Finally, the robustness and accuracy of our hybrid methodology is demonstrated for forced-convection heat transfer over stationary spherical particles at high Prandtl numbers. Keywords: adaptive mesh refinement, high Prandtl number, high Schmidt number, boundary layers

1

1. Introduction

2

Numerous examples exist in nature of fluid flow systems with associated

3

scalar transport such as dilution of pollutants in air (Zhong et al., 2016), nu∗ Corresponding

author Email address: [email protected] (E.A.J.F. Peters)

Preprint submitted to Chemical Engineering Science

October 18, 2019

4

trients dissolution in estuaries (Andriˇcevi´c and Galeˇsi´c, 2018), tracer studies

5

for reactor characterization (Dixon et al., 2006), mantle convection (Stadler

6

et al., 2010), heat transfer and dispersion in porous media (Xu et al., 2019).

7

Scalars can be classified as active or passive depending on whether they alter

8

the flow field or not. Numerical simulations of such systems is ubiquitous and

9

of paramount importance to gain fundamental understanding of the underlying

10

phenomena. Very often, the scalar quantity of interest is the concentration of a

11

certain species, or the temperature.

12

Direct Numerical Simulation (DNS) constitutes an important and powerful

13

tool to study complex multiphase flow systems. In DNS, the Navier-Stokes

14

equations are solved to obtain the velocity distribution for the fluid flow whereas,

15

an associated convection-diffusion equation is solved to obtain the scalar field.

16

To properly numerically resolve the boundary layers, an important requirement

17

for DNS is that the mesh size must be smaller than the smallest length scale,

18

which in turn depends on the molecular transport coefficients of the quantities

19

that are being solved. It follows from the boundary layer theory that, in the

20

laminar case, boundary layer thicknesses for momentum (δmom ), temperature

21

(δT ) and mass (δm ) are related as δT = P r−0.5 δmom and δm = Sc−0.5 δmom where

22

P r and Sc are the Prandtl and Schmidt numbers, respectively. Both numbers

23

are a ratio of diffusivities. When the scalar of interest is the temperature, the

24

Prandtl number is defined as P r = ν/α, where ν is the momentum diffusivity

25

(i.e. kinematic viscosity) and α is the thermal diffusivity. For many fluids e.g.

26

water, oils, polymer melts and earth mantle, P r > 1 and the thermal boundary

27

layer thickness is much smaller than the momentum boundary layer thickness.

28

Ostilla-Monico et al. (2015) demonstrated that even for P r ≈ 1 the practical grid

29

requirement for scalar transport is higher than that of momentum transport. For

30

mass transfer the Schmidt number is defined as Sc = ν/D, with D the diffusion

31

coefficient. In liquids, momentum can diffuse via intermolecular collisions, while

32

for mass diffusion, molecules need to physically move in a crowded environment.

33

This difference in diffusion mechanisms gives Sc  1.

34

Choosing a grid fine enough to resolve the scalar field would over-resolve 2

35

the momentum field. The situation becomes aggravated due to two reasons:

36

(a) solving for momentum transport requires a vector field (3 variables) and

37

an associated pressure field in comparison to a single variable needed for scalar

38

transport, (b) in many applications, for a given resolution, most of the CPU

39

time is spent on solving the momentum transport (≈ 90%). Given the above

40

arguments, it is evident that over-resolving the momentum field leads to higher

41

memory and CPU demands. To circumvent the above problems, two approaches

42

are prevalent in literature: 1) Solving the scalar equations on a coarse grid, but

43

use a sub-grid model for the unresolved scales. 2) Resolving the finest scales

44

fully only in regions where it is needed by way of multiple resolution grids.

45

The former approach involves using a subgrid scale model with a filter length

46

in the viscous-convective scale similar to the approach used in large eddy simula-

47

tions of turbulence which rely on the following principle: resolve the large eddies

48

and model the small eddies. Notable works in this direction has been carried

49

out for high Schmidt number flows in case of mass transfer from gas bubbles,

50

where the boundary layer solution is approximated by an analytical profile and

51

then added as a source term in the neighboring coarse grid after a certain cutoff

52

distance from the interface (Weiner and Bothe, 2017; Aboulhasanzadeh et al.,

53

2012). Verma and Blanquart (2013) adopted the subgrid scale model that uses

54

a filter length to include source terms accounting for the under-resolved scalar

55

turbulence. Though these methods are relatively simple to implement, they

56

are susceptible to the choice of analytical function used to model the boundary

57

layer. We thereby propose to use a computational approach in which different

58

spatial resolutions for momentum and scalar fields are used.

59

In the literature, several studies have been reported adopting such a hybrid

60

computational approach. Gotoh et al. (2012) used a pseudospectral method

61

to solve the Navier-Stokes equations on a coarse mesh and a finite difference

62

discretization of the scalar convection-diffusion equation on a (substantially)

63

finer mesh. They report a reduction of CPU time by 26% for Sc = 1 and 76%

64

for Sc = 50. While interpolating the velocity field from the coarser mesh to

65

the finer mesh, one of the key constraints that needs to be obeyed is the di3

66

vergence free condition. Ostilla-Monico et al. (2015) and Chong et al. (2018)

67

proposed a similar approach using a coarser grid for the Navier-Stokes equa-

68

tion and a finer grid for active/passive scalar transport using a finite difference

69

and finite volume approach respectively by also using divergence-free/solenoidal

70

interpolations. For interface capturing in multiphase flows, Gada and Sharma

71

(2011) used a dual-grid method where a finer grid was used to resolve the level

72

set and the temperature field while the coarser grid was used to resolve the

73

momentum/velocity field.

74

In the works cited above, the authors have used a uniformly refined grid

75

for scalar transport in the multiple resolution techniques or dual grid methods.

76

The uniform refinement method, though not very efficient, is useful at low to

77

moderate Schmidt/Prandtl numbers, but suffers from problems associated with

78

memory and CPU requirements, which grows exponentially with each level of

79

refinement beyond the coarse grid used to solve the Navier-Stokes equations.

80

Also, at higher values of P r and Sc, the solution fronts become sharper since

81

the molecular transport coefficient of the scalar quantity is decreased requiring

82

a higher refinement. Considering the targeted study of forced convection heat

83

transfer over solid particles at high Prandtl numbers, the actual location of the

84

sharp thermal boundary layer is known apriory i.e. at the solid surface.

85

Therefore, in the novel dual grid methodology proposed in this paper, an

86

adaptive mesh is used instead of a uniformly refined Cartesian mesh for the

87

scalar transport. Adaptive mesh refinement (AMR) has its own set of com-

88

putational challenges specific to different methods of mesh refinement. Among

89

many different AMR strategies, notable are block structured AMR, overset or

90

Chimera grids, tree-based AMR etc. A tree-based adaptive mesh framework is

91

chosen where, each grid cell is a structured square/cubical cell called quadrant

92

(octant in 3D). This simplifies the coarse-fine grid interpolation procedures.

93

Also a tree-based adaptive grid provides flexibility to adapt to any available

94

features in the solution domain for e.g. walls, deformable interfaces (fluid-fluid

95

interface), non-deformable interfaces (fluid-solid interface) or to the gradients

96

present in the computed fields. 4

97

In this work, we describe the methodology and implementation of the dual

98

grid method along with the divergence free velocity interpolations. The method-

99

ology is subsequently verified for different theoretical problems and validated

100

with empirical heat transfer correlations for forced convection over single and

101

multiple solid particles.

102

2. Model Description

103

2.1. Governing Equations

104

For the solution of a system with an associated passive scalar transport,

105

the equations for conservation of mass and momentum need to be solved along

106

with the scalar convection diffusion equation. The continuity and Navier-Stokes

107

equations describing the velocity field, u (u, v, w), and pressure, p, can be writ-

108

ten as:

ρ 109

∂u + ρ∇ · (uu) = −∇p + ρg + ∇ · µ(∇u + (∇u)T ) ∂t ∇·u=0

(1) (2)

110

where ρ, µ are the local density and viscosity, and g is the acceleration due to

111

gravity. A regular staggered Cartesian grid is used for the numerical solution

112

of the Navier-Stokes equations. The curved fluid-solid interface of immersed

113

objects such as particles does not conform to the Cartesian grid and to accu-

114

rately incorporate the solid-fluid coupling for momentum, an immersed bound-

115

ary method (IBM) is used. This results in a flow field that satisfies the no-slip

116

boundary condition at the solid-fluid interface. Details addressing the imple-

117

mentation, verification and validation of the current IBM method can be found

118

in the papers of Das et al. (2017b); Deen et al. (2012).

119

The generalized transport equations for a passive scalar, φ,is ∂φ + ∇ · (uφ) = α∇2 φ + Sφ , ∂t

5

(3)

120

where α is the diffusivity, which is the thermal diffusivity, k/ρCp , in case of

121

heat transfer (k and Cp represent the thermal conductivity and specific heat

122

capacity, respectively) and the molecular diffusivity, D, in case of mass transfer.

123

Although, the Navier-Stokes equations are solved on a regular Cartesian

124

grid, the scalar transport equations are solved on a grid that can be adaptively

125

refined as the solution proceeds in time. For the purpose of the work presented

126

here, we restrict ourselves to the problem of heat transfer, i.e. φ is temperature.

127

For the representation of the velocity field on this refined grid, divergence free

128

interpolation of the velocity field defined on the coarse ‘hydrodynamic’ grid is

129

required. For the heat transfer we restrict ourselves to the constant solid tem-

130

perature case for the immersed particle, which is represented on the adaptive

131

grid with a high enough resolution to have a sufficiently smooth staircase rep-

132

resentation of the particle interface. In the subsequent subsections we focus on

133

the discretization of the scalar convection diffusion equation on the adaptive

134

grid, velocity update on the adaptive grid from the Cartesian grid and the grid

135

adaptivity.

136

2.2. Spatial Discretization

137

In the current dual grid method the base hydrodynamic/momentum grid is

138

of static nature whereas the passive scalar grid is of dynamic nature featuring

139

grid adaptation. The underlying hydrodynamic grid is comprised of a staggered

140

Cartesian grid stored in arrays of fixed sizes, each grid point identified by a

141

triad of indices i, j, k for the x, y, z-direction respectively. For the adaptive grid,

142

the computational domain is represented in the form of control volumes which

143

are square in 2D (or cubic in 3D). These elementary square (or cubic) control

144

volumes are adapted using a quadtree (or octtree) data structure (Popinet,

145

2003). Figure 1(a) depicts a representation of such a discretization for a 2D

146

tree where an adaptive grid is initiated as one root cell uniformly refined twice

147

to give a 4 × 4 Cartesian grid. Each control volume (shown in green) is defined

148

as a cell and can have 4 descendants (8 in case of 3D) defined as children,

149

in which case, the cell is defined as the parent cell of these children. A cell 6

150

which has no parent cell is defined as the root cell (black cell in figure 1(d))

151

while cells that do not have any children are termed leaf cells (all numbered

152

cells in the figure 1(c,d)). The length of any cell edge is denoted by h. The

153

level of a cell is defined with the root cell as reference level 0 and children

154

being one level higher than the parents. The actual data structure used is a

155

so-called graded/balanced quadtree (or octree), which means that adjacent cell

156

levels are not allowed to differ by more than one, i.e. a maximum of one hanging

157

node is possible at any cell face. This 2:1 balance constraint is beneficial for

158

the spatial discretization of partial differential equations. Figure 1(b) shows a

159

quadtree that violates this criterion since there exists a face with two hanging

160

nodes (shown as yellow dots). The constraint restricts the number of possible

161

cases which is very convenient in devising efficient numerical schemes for such a

162

spatial representation. Each cell on the adaptive tree is identified by a unique

163

index called the Morton number derived by bit interleaving the i, j, k index of

164

the associated cell on an uniform grid at that level. For example, the quadrant

165

corresponding to 2D Cartesian control volume (2, 2) has a Morton index(Zid) of

166

3 in figure 1(a). It is worth mentioning that the Morton indexing (z-order) of the

167

constituent control volumes of an adaptive grid leads to a non-banded pattern

168

of matrix sparsity, even when all the cells are at the same level of refinement.

169

Though, the benefit of such an encoding is that the cells lying closer to each

170

other in the physical space, also lie closer to each other in the z-order space.

171

The base framework used for the implementation of the developed method-

172

ology is an open source library which provides efficient parallel tree data struc-

173

ture and functions to perform operations on this data structure, called p4est

174

(Burstedde et al., 2011; Isaac et al., 2015). The chronological order of operation

175

for the methodology explained here is described very well with the flow diagram

176

given in figure 2. The key building blocks are described in the next subsections.

7

177

178

179

2.3. Convection To discuss the treatment of convection in the scalar transport equation, we focus on the following reduced equation: ∂φ + ∇ · (uφ) = 0 ∂t

180

(4)

which can be discretized using a finite volume technique, to obtain: φn+1 = φn +

∆t X ( uf n+1 φnf Af ) ∆V

(5)

f aces

181

where ∆t is the time step whereas ∆V and Af respectively represent the volume

182

and face area. The velocity at the cell face, uf , is either directly available, or

183

is obtained from the coarse Cartesian grid after interpolation. The convective

184

flux at any cell face can be computed after estimating the value of the scalar

185

at the cell face, φf . Figure 3 (a) denotes a schematic representation of the cell

186

configuration in the case of same size cells, i.e. in the absence of adaptivity.

187

The face in consideration for the calculation of convective flux is shown with a

188

dashed line. Assuming that the velocity at this face is positive we adhere to the

189

nomenclature defined in figure 3, where the cell being emptied is termed as the

190

donor cell (D), the cell being filled is defined as the acceptor cell (A) and the

191

cell upstream to the donor cell is termed as the upwind cell (U). The respective

192

cell sizes are hD , hA and hU . The face scalar, i.e. the value of the scalar used at

193

the face to compute the flux through that face, which is depicted with a green

194

arrow in the figure can be calculated as a function of the scalar values φU , φD

195

and φA . Two schemes have been implemented here viz. the first order upwind

196

scheme and a total variation diminishing (TVD) min-mod scheme also referred

197

to as the Barton scheme (Centrella and Wilson, 1984).

198

Three different estimates of the face scalar are computed using linear upwind-

199

LI ing φLU f , linear interpolation/central differencing φf and first order upwinding

200

OU φF . For defining these scalar values at the cell face, the cell sizes or the f

201

associated distances are used. The notation dAf denotes the distance of the

202

cell centre of the acceptor cell from the face under consideration whereas dAD 8

203

denotes the distance between the cell centres of the acceptor and donor cells.

204

Hence, the face scalar values are defined as follows:

205

Linear Upwinding: hD hD )φD − 0.5 φU dU D dU D

(6)

dDf dAf φA + φD dDf + dAf dDf + dAf

(7)

φLU = (1 + 0.5 f 206

Linear Interpolation: φLI f =

207

First Order Upwind : OU φF = φD f

208

(8)

For the first order upwind scheme the face scalar, φf , is simply the value but, for the total variation diminishing scheme a unique value of φf is

209

OU φF f

210

determined by comparing the three distinct values, which for an outward velocity

211

at the cell face can be given as follows:   MIN(φF OU , MAX(φLU , φLI )), if φA ≤ φD f f f φf =  F OU LU MAX(φ , MIN(φf , φLI f f )), if φA > φD

(9)

212

For adaptive grids, as shown in figure 3 (b), the large dots for acceptor and

213

upwind cell correspond to the values required in the above discretization which,

214

may or may not always coincide with the node centre values (shown with small

215

dots). In that case the required values are computed using interpolation (in the

216

direction perpendicular to the face normal) from the neighboring cells. These

217

interpolations are also used for a Laplacian operator, which will be presented

218

in the subsequent subsection. For the case of same size cells neighboring a

219

face, the convective flux, uf φf Af,D is subtracted from the donor cell(s) and

220

uf φf Af,A is added to the acceptor cell(s). Here, Af,D and Af,A correspond to

221

the face area of the donor cell(s) and acceptor cell(s), respectively. In the case

222

of a hanging node at the face, there will be multiple donor/acceptor cells with

223

Af,D 6= Af,A . Considering the case of an outward convective flux at the cell

224

face, (∆t/∆VD ) φf uf Af,D is subtracted from the donor cell and correspondingly 9

225

added to the acceptor cell. If there are multiple acceptor cells then the flux is

226

divided equally among the half size cells across the face. As evident from equa-

227

tion 5, the term uf φf Af is multiplied by ∆t/∆V as a result of the explicit first

228

order Euler forward treatment of the temporal term. Hence, the time step ∆t is

229

chosen such that the Courant–Friedrichs–Lewy (CFL) number ( umax ∆t/∆xmin

230

) is less than 0.33, where ∆xmin is the cell size of the smallest quadrant/cell in

231

the domain and umax is the maximum velocity in the domain.

232

2.4. Diffusion

233

234

To discuss the treatment of diffusion on the scalar transport equation, we focus on the following reduced equation: ∂φ = ∇ · (α∇φ) + Sφ ∂t

(10)

235

where, Sφ represents the volumetric source term. Discretization of equation 10

236

with a finite volume technique and a semi-implicit scheme, leads to the following

237

representation: φn+1 − φn = (1 − β)∇ · (αn ∇φn ) + β∇ · (αn ∇φn+1 ) + Sφn ∆t

(11)

238

where, n + 1 and n corresponds to values at new time step and old time step,

239

respectively. To calculate the diffusive flux at each face, a numerical representa-

240

241

tion of the Laplacian operator ∇·(∇φ) or ∇2 φ is required, which in it’s discrete P form is given as f aces ∇f φAf , where the ∇f φ is the gradient of φ defined at

242

the cell face. For a conventional Cartesian grid, this is normally evaluated by

243

central differencing, which is second order accurate for cells of the same size.

244

However, when this scheme is applied for a cell face with a hanging node, the

245

scheme is first order accurate (Losasso et al., 2006). This type of differencing

246

would result in schemes with inconsistent order of convergence, across the whole

247

domain. To circumvent this problem, one needs to specifically adopt schemes

248

that are overall second order accurate. In this paper, two different Laplacian

249

operators viz. ∇21 and ∇22 have been implemented that are both overall second

250

order accurate but, result into different global matrix representations. 10

251

While the computation of the gradient at a cell face with second order ac-

252

curacy is simple on a regular Cartesian grid, this is more difficult within the

253

adaptive framework. In practice, only two configurations are possible at a given

254

cell face for the calculation of face centered gradients (figure 4):

255

256

(I) Both sides of the face are at the same level. (II) One side of the face is at one level lower than the other one.

257

In case (I), the stencil reduces to a configuration observed in conventional Carte-

258

sian grid and thus one can use the standard second order accurate central dif-

259

ference approximation. In case (II), the approaches for ∇21 and ∇22 differ and

260

will be discussed subsequently.

261

2.4.1. Laplacian Operator 1

262

This Laplacian operator was first given by Popinet (2003) and forms an

263

ingredient of the open source CFD code named Gerris Flow Solver. For the

264

Laplacian operator 1, the face of interest in the direction d is the one shared

265

between the cell (C) and the neighbor (Nb ) as shown in figure 5. In the case (II)

266

as described above, a second order polynomial is fitted passing through the cell,

267

C, the neighbor Nb and the opposite cell OC. The gradient at the face center

268

can thus be calculated using the derivative of the polynomial evaluated at the

269

examined face. Based on the configuration of the opposite cell, only two subcases

270

can arise due to the 2:1 restriction: an undivided leaf cell at the same level of cell

271

C or a parent cell with 4 leaf cells (8 in 3D). If the opposite cell is divided, which

272

happens to be the case shown in the figure 5, the interpolated value φ7 is created

273

from the 2 (or 4 in 3D) children across the opposite face by linear averaging.

274

Similarly, the value φ6 in the double sized neighbor Nb is also interpolated

275

using a second order polynomial fitted through its neighbors in the direction

276

perpendicular to d. The configuration of the neighbor of the neighbor across the

277

lower perpendicular(N⊥L ) and the upper perpendicular (N⊥U ) can again give

278

rise to four different combinations out of which one sample configuration has

11

279

been sketched in figure 5. With the definition of φ6 and φ7 from figure 5, the

280

expression for an undivided and divided opposite cell are respectively:

281

282

283

1 1 8 h∇fd φ = − φC − φOC + φ6 3 5 15

(12)

2 8 14 h∇fd φ = − φC − φ7 + φ6 9 27 27

(13)

where h is the size of reference cell C and ∇fd φ is the gradient at face f in the

direction d. Subsequently for the four different combinations possible across the perpendicular faces of the neighbor Nb are:  5 3 15   if N⊥L  16 φN b + 32 φN⊥L − 32 φN⊥U     5φ + 5 φ − 1 φ if N⊥U 6 Nb 21 3 14 N⊥U φ6 = 1 1   φN b + 7 φN⊥L − 7 φ4 if N⊥L      8φ + 2φ − 1φ if N⊥L 9 Nb 9 3 9 4

& N⊥U are undivided is undivided is undivided

(14)

& N⊥U are divided

284

where, φ3 and φ4 are constructed using linear averaging of the children values

285

across the lower and upper perpendicular faces respectively. Additional issues

286

arise when extending this discretization to 3D since there are two directions

287

perpendicular to the direction d that needs to be traversed to obtain the value

288

of φ6 . Moreover, these perpendiculars are non-intersecting unlike the case in

289

2D, which means neither of the two interpolated values(φ6,1 and φ6,2 ) lie on

290

the axis passing through the cell C and OC. In this case a unique value φ6

291

is computed using the relation φ6 = φ6,1 + φ6,2 − φN b . The gradient for the

292

left face of cell Nb is evaluated as minus the average gradients of the coinciding

293

smaller faces (Popinet, 2003).

294

2.4.2. Laplacian Operator 2

295

For the Laplacian operator 2, where the face of interest has a hanging node

296

as shown in figure 4 (II). The face has been sketched in a detached manner for

297

the purpose of discussion. Any face with a hanging node will be shared by 2 (4

298

in 3D) small size cells on one side of the face denoted in the figure as cell S1

299

and S2 and one large cell denoted as cell L. This operator was first proposed by 12

300

Losasso et al. (2006) where central differencing can be used to define a gradient,

301

∇fd = (φL − φS1 )/0.75 hL where, hL is the size of the large cell. This gradient

302

is evaluated at the mid point of the line joining φS1 and φL . On basis of the

303

assumption that O(∆x) perturbations in the scalar value location would still

304

yield consistent approximations as elucidated in Gibou et al. (2002), one can

305

then conclude that setting ∇fd φS1 = (φL − φS1 )/0.75 hL still leads to convergent

306

solutions. These authors report that the discretization is found to be convergent

307

and of first order. More importantly this discretization suffers from inconsistent

308

gradient approximation at the three coincidental faces since the gradient at the

309

large face is given by, ∇fd φL = As /AL × ∇fd φS1 + As /AL × ∇fd φS2 where As is

310

311

312

313

the area of the smaller face and AL is the area of the larger face.

Thus the gradient at the larger face in the direction, d, can be written as ∇fd φL AL and for the smaller as ∇fd φS1 As and ∇fd φS2 As ,

faces corresponding to S1 and S2 can be written respectively. A possible way to circumvent the

314

inconsistent gradient evaluation is to impose that the gradients at the small faces

315

must be equal to that of the large face, i.e. ∇fd φS1 = ∇fd φL and ∇fd φS2 = ∇fd φL .

316

Care has to be taken with respect to the sign of ∇fd as for the small face the

317

normal is in the opposite direction to the normal defined at the large face.

318

Although the discretization is formally first order accurate, Losasso et al. (2006)

319

report a second order convergence with the support of arguments from Lipnikov

320

et al. (2004).

321

2.5. Boundary Conditions

322

At the domain boundaries, both Dirichlet and Neumann boundary condi-

323

tions are applied with second order accuracy by use of ghost cell values. For

324

the boundary cells, a ficitious neighbor of the same size is constructed and the

325

earlier described discretization methods are implemented. For the case of Lapla-

326

cian operator 1, the boundary conditions are factored in while interpolating the

327

value φ6 (see figure 5) since the neighbor across the lower perpendicular N⊥L

328

and the upper perpendicular N⊥U can also be a boundary cell.

13

329

330

331

2.6. Linear solver When solving for both diffusion and convection using equation 3, the resultant set of simultaneous linear equations can be written as: ac φc +

X

anb φnb = bc

(15)

nb 332

where the subscript ‘c’ corresponds to the cell center and subscript ‘nb’ cor-

333

responds to the neighboring cells (neighbors of neighbors as well in case of

334

the Laplacian operator described in section 2.4.1) arising from the state of

335

quadtree/octtree. The left hand side of equation 15 contains the coefficients

336

and the unknowns φ which arise out of the implicit treatment (unknown values

337

at the current time step) of the diffusive flux and the right hand side of this

338

equation contains the terms arising from the explicit treatment (using the known

339

values from the previous time step) of the convective flux. The resulting set of

340

simultaneous linear equations is solved using robust matrix solvers available

341

from an open source library called HYPRE (Falgout et al., 2002). Laplacian oper-

342

ator 1 leads to non-symmetric matrices that are then solved using an algebraic

343

multigrid (AMG) solver called boomerAMG. On the contrary, Laplacian opera-

344

tor 2 results in symmetric matrices, which can be efficiently solved using the

345

PCG/BiCGSTAB/FlexGMRES class of solvers with preconditioning by boomerAMG.

346

It is worth mentioning that the resulting matrix from Laplacian operator 2 is

347

less dense than that obtained from Laplacian operator 1 and is easier to invert

348

with fewer iterations. Also, the Laplacian operator 1 is only valid for a tree

349

which is 2:1 corner balanced (corner cells’ level do not differ by more than one)

350

whereas, Laplacian operator 2 has no 2:1 restrictions. In both cases, the solver

351

scales to 103 processors based on the hybrid OpenMP+MPI model and uses parallel

352

graph partitioning methods for load balancing.

353

2.7. Staggered Velocity Field and Communicator

354

An essential element of this novel dual grid methodology is the velocity

355

update on the adaptive grid from the underlying fixed Cartesian grid. The ve-

356

locity needed to calculate the convective fluxes of the scalar is available on the 14

357

hydrodynamics grid and needs to be transferred to the adaptive grid at each

358

time step. Hence, the update process needs to be time efficient. Therefore, a

359

communicator is implemented such that the velocity matrices are allocated in a

360

memory address space, shared between the hydrodynamics code and the scalar

361

transport code. This memory can be modified by the hydrodynamics code for

362

updating the velocity field and subsequently read by the scalar transport code.

363

The read-write synchronization between the two codes for this memory block

364

is facilitated by POSIX semaphores. These syncronisation semaphores can be

365

considered as locks depicted in figure 2 that lock & unlock either the hydrody-

366

namics code or the scalar transport code for operations. Since, the semaphores

367

and the shared memory are available on the physical memory employed by both

368

parts of the numerical code, it provides instantaneous communication between

369

the two grids for velocity matrix access. Another consideration in the numeri-

370

cal implementation is the choice of velocity field discretization on the adaptive

371

grids. Since, the underlying Cartesian grid contains staggered velocity field en-

372

coding, the same staggering is maintained for the adaptive grid. In the current

373

implementation, this choice leads to duplicate storage of face values among cells

374

sharing a common face but it simplifies the velocity update process.

375

The velocity update on the adaptive grid has two building blocks: (a) map-

376

ping of the adaptive grid to the underlying Cartesian grid at the beginning of

377

the simulation (b) Velocity update for cells smaller than the base Cartesian grid

378

by employing interpolations. For the first building block, a ‘forest’ (Burstedde

379

et al., 2011) consisting of one root cell (also called as tree) or multiple trees is

380

initiated and then refined uniformly such that the resultant mesh maps exactly

381

with the Cartesian grid used for the flow field computation. This uniform level

382

of refinement will be referred as the base level. Next, a two index identifier(three

383

in case of 3D) is attached to each cell on the adaptive grid, which stores it’s

384

corresponding i and j index (and also k in case of 3D) in the Cartesian grid.

385

This index once initiated, is kept intact throughout the adaptation process which

386

means that all children of a parent originating due to grid refinement beyond the

387

base level will inherit the i and j index as it is. In the current implementation, 15

388

coarsening is restricted to the base level with which the simulation was started.

389

Therefore, it is not possible to have a cell size larger than the cell size of the

390

hydrodynamics grid. For cells which are at the base level, the face velocities are

391

updated by using the index [i j]. For cells smaller than the base level, inter-

392

polations are used such that the resultant velocity adheres to the divergence free

393

criteria (equation 2). To ensure this, a piecewise linear interpolation method

394

is used to interpolate a velocity component in its flow-direction, while staircase

395

interpolation is used in perpendicular directions. This interpolation is visualized

396

in figure 6(b). For example, the −x face velocity of the scalar cell (shown in red),

397

ux− , is computed using the relative distance from the bounding hydrodynamic

398

cell’s (shown in blue) faces in x−direction i.e. d0 and d1 with their respective

399

velocities Ux− and Ux+ as: ux− = d1 /(d0 + d1 ) Ux− + d0 /(d0 + d1 ) Ux+ .

400

2.8. Grid Adaptation

401

402

403

404

405

406

407

408

One of the most important aspects in the solution process is the grid adaptation. Grid adaptation involves three steps in the following order • Refinement : Flag cells for refinement and replace the flagged parent with children.

• Coarsening : Flag cells for coarsening and replace the flagged children with their parent.

• Balancing: After the above two operations, refine relevant cells to ensure that the forest meets the 2:1 balance criteria.

409

The first two steps have two user inputs that depend on the problem at hand:

410

which cells to refine/coarsen and how to update values in the new cells upon

411

refinement/coarsening. The latter forms the only user input for the third step,

412

i.e. balancing. The decision on which cells need to be refined and coarsened

413

depends on the problem features such as the presence of an interface, which can

414

be used as an indicator to flag cells in its vicinity. Another flagging criterion

415

used in the current simulations is an error estimate, proportional to the sum of 16

P3

2

416

the square of the gradients in each cell as:

417

are the cell size and cell volume, respectively whereas ∇ci is the cell-centered

i=1

(h∇ci φ) ∆V , where h and ∆V

418

gradient in the direction i. The cell-centered gradients are computed using a

419

min-mod approach where for a cardinal direction, the cell-centered gradient is

420

the minimum of the two bounding face-centered gradients or if the two face-

421

centered gradients have opposite sign then the cell centered gradient is taken

422

to be 0. For example the gradient in the x− direction is bounded by the face

423

centered gradients calculated at the −x and +x face, respectively. For refine-

424

ment, those cells are flagged for which the error estimate is more than a set

425

value. On the other hand, for coarsening if all the children have their respective

426

error estimates lower than the critical value then the family of cells is marked

427

for coarsening. When the refinement occurs the scalar value of each child is

428

updated as φchild = φparent ±

h c h ∇ φparent ± ∇cy φparent 4 x 4

(16)

where ∇cx φparent is the cell-centered gradient in the x direction, h the size of

parent cell. The critical value of the error estimate for each cell is taken to be 10−5 ∆V in all the simulations where on the fly grid adaptivity is enabled. During refinement the velocity is updated using the divergence-free interpolation method. Referring to figure 6 (a), for a cell that is under refinement, the face velocities (shown in red) are computed from the parent cell’s face velocities(shown in brown), which for the x face velocities can be written as ui− 12 ,j− 14 = ui− 12 ,j+ 14 = ui− 12 ,j ui,j− 14 = ui,j+ 14 =

1 2

ui− 12 ,j +

(17) 1 2

ui+ 12 ,j− 14 = ui+ 12 ,j+ 14 = ui+ 12 ,j

ui+ 12 ,j

(18) (19)

429

During coarsening, the scalar value of the parent cell is computed from the

430

average of the children scalar values which can be written for a 2D case as

431

follows: φparent =

1 4

X all children

17

φchild,i

(20)

432

While updating the velocities during coarsening, the cell face velocities on the

433

outer edges/faces of the children are retained.

434

3. Verification

435

3.1. Convection

436

To test the implementation of the convection of a scalar, a step profile is

437

initiated inside a square domain. For this, two test cases are considered: (a)

438

a uniformly refined grid (b) adaptive grid. The uniform grid considered here

439

is generated out of one root cell uniformly refined to level 7 and 8, thereby

440

resulting into a grid of 128 × 128 and 256 × 256, respectively. The adaptive

441

grids on the other hand, have been generated by first taking the uniform grid

442

and then refining near the discontinuity of the step profile with maximum re-

443

finement restricted to 8, 9, 11 and 13 levels. A constant velocity u = 1 m/s

444

is prescribed over the whole domain in only one cardinal direction. The step

445

profile is initiated at a non-dimensional distance of 0.0390625 from the origin

446

(−x boundary of domain) which corresponds to a length of 5 grid cells and 10

447

grid cells for level 7 and level 8 refinement respectively. The time step is chosen

448

such that the Courant–Friedrichs–Lewy (CFL) number with the minimum grid

449

size (corresponding to refinement level 13) is equal to 0.3. A step profile when

450

advected with a constant velocity should theoretically remain a step profile hav-

451

ing travelled a distance u(t2 − t1 ) where t2 and t1 correspond to different time

452

instances. It is however well-known that, all numerical convection schemes ex-

453

hibit a certain degree of numerical diffusion thus reducing the sharpness of the

454

step profile. Figure 7(a) shows the performance of the two schemes explained in

455

section 2.3 from the obtained non-dimensional scalar (˜ c) profiles at 0.4 s, where

456

one can see that the Barton scheme on an uniform grid is less diffusive than the

457

first order upwind scheme. Figure 7(b) shows the results obtained from both

458

schemes on an adaptive grid. In this test case the refinement and coarsening is

459

performed after each solution time step based on the gradient of the solution.

460

As the maximum allowed refinement at the discontinuity is increased, the solu18

461

tion approaches the analytical solution, which is shown as a black dashed line.

462

Comparison of the results of the adaptive grid with those obtained from the uni-

463

form grid at a refinement level of 7, reveals that the numerical diffusion on an

464

adaptive grid becomes less prominent as the refinement near the discontinuity

465

is increased.

466

Since in the aforementioned test case the prescribed velocity is in only one

467

cardinal direction, another test was performed where a circular blob of diam-

468

eter D = 1 mm is advected with a velocity field inclined to control volume

469

faces at 45◦ , ux = uy = 1 m/s. Here, the accuracy of the method is demon-

470

strated by comparing the results with that obtained from open source Gerris

471

flow solver, which uses a Godunov method for the evaluation of the convective

472

fluxes (Popinet, 2003). The circular blob with it’s center located at (0.25, 0.25)

473

is initiated in a domain of edge length 1. The grid is initially at a uniform

474

refinement of level 7 and refinement is restricted to a maximum level of 10 at

475

the solid surface. Refinement and coarsening is carried out by using the error

476

estimate mentioned in section 2.8. Figure 8 (a) shows the scalar profile and the

477

associated grid at the start of the simulation. Figure 8(b) shows the solution

478

and the grid obtained after 0.5 s from the Gerris flow solver whereas figure 8(c)

479

shows the solution and the grid obtained at the same time using the methodol-

480

ogy presented in this paper. The image of the grid shows the solution marked

481

at a scalar contour of 0.5 (using a yellow line) whereas the numerical diffusion is

482

visualized by the contour of 0.001 shown with the blue line. It can be observed

483

that the convection scheme presented here exhibits a similar degree of numerical

484

diffusion as obtained from the Gerris flow solver.

485

3.2. Grid convergence of Laplace operators

486

In section 2.4, the implementation of two schemes for the Laplace operator,

487

needed in the convection-diffusion equation, was discussed. To compare the

488

order of convergence of the two operators without the influence of errors arising

489

from the discretization of the time derivative, a convergence test is performed

490

on Cartesian as well as on adaptive grids by solving a pressure Poisson equation 19

491

on a unit square domain centered around the origin. The divergence of the

492

intermediate/projected velocity field is given by the equation ∇ · u∗ (x, y) = −π 2 (k 2 + l2 ) sin(πkx) sin(πly)

(21)

493

where, u∗ (x, y) is the projected velocity, which is non-solenoidal. For a pressure

494

field, p, the pressure Poisson equation is of the form ∇2 p = ∇ · u∗ . When

495

k = l = 3, the analytical solution is given by the pressure field equation using

496

Neumann boundary conditions on all sides as p(x, y) = sin(πkx) sin(πly) + κ

(22)

497

where, κ is an arbitrary constant. Numerically, the system is not fully specified

498

with the Neumann boundary condition and hence an approach suggested by

499

Popinet (2003) is used, whereby the boundaries are specified with a pressure

500

value: p = sin(3πx) sin(3πy) as a Dirichlet boundary condition. Under these

501

conditions, κ reduces to the average value of the computed pressure over the

502

entire domain which is zero. The initial guess for the pressure field is a constant

503

value p0 .

504

Since the Laplace operator 1 and 2 both simplify to the same conventional

505

central differing scheme for Cartesian grids, only one of the operators is tested

506

on a Cartesian grid, while both of them are tested on adaptive grids. To study

507

the convergence for Cartesian grids, the grid is initiated out of one root cell

508

of unit length refined uniformly to a level L (thus giving 2L cells along one

509

direction) whereas for the adaptive grid an additional refinement operation is

510

performed, such that the cells within a circular patch of radius 0.25 are refined

511

to a level L + 2 (see figure 9). The 2:1 balance constraint thus creates a small

512

buffer zone of cells at level L + 1 as well. The grid convergence results are

513

reported in terms of L1 and the L2 error norms. A volume-averaged pth norm

514

for error can be defined as:

P p i |ei | vi Lp = P i vi

(23)

515

where, the summation is performed over all the cells in the domain, ei is the

516

difference between the analytical value and the value obatained from the simu20

517

lations whereas vi is the volume of the cell i. The order of convergence can then

518

be computed using the slope of the error norms. Figure 10 shows the evolution

519

of error when, successively, all cells are refined once. It can be clearly observed

520

that the order of convergence of both the Laplacian operators is indeed 2, which

521

is denoted by the guiding dashed lines in the graphs.

522

3.3. Staggered Velocity Interpolation

523

For testing of divergence free (or solenoidal) velocity interpolation as de-

524

scribed in section 2.7, an initial velocity distribution is prescribed in an unit

525

cube on a staggered (coarse) hydrodynamic grid. Correspondingly, for the adap-

526

tive grid, a unit cube root cell is uniformly refined to a level L such that the

527

control volumes on the adaptive grid coincide with the control volumes on the

528

hydrodynamics grid. The velocity values at the control volume faces follow from

529

a solenoidal analytical flow field defined as: u(x, y, z) = (0, −2 sin2 (πy) sin(πz) cos(πz), 2 sin2 (πz) sin(πy) cos(πy)) (24)

530

All the cells on the adaptive grid are flagged to recursively refine three times

531

and use the solenoidal velocity interpolation to define the values on the faces of

532

the newly created children. The errors arising from the difference in the inter-

533

polated face velocities and those available from equation 24 are then quantified.

534

Figure 11 shows the error norms with respect to the analytical solution after the

535

interpolation of the velocity field. The L∞ norm signifies the maximum error

536

in the domain. Since, the initial level of refinement influences the error in the

537

interpolated solution, the figure shows the error with respect to the level L at

538

which the original solution/grid was created. The figure also shows the maxi-

539

mum divergence encountered in the interpolation step which is essentially zero

540

considering the machine precision of a double precision floating point number.

541

Figure 12 shows the flow field obtained using interpolation for a grid at the 10th

542

level of refinement derived from a grid initiated at the 8th level of refinement.

543

It can be inferred that if the hydrodynamics grid is fine enough to resolve the

544

momentum boundary layers and the re-circulation zones, then the adaptive grid 21

545

preserves these flow features while using the solenoidal interpolation developed

546

in section 2.7.

547

4. Results

548

4.1. Heat conduction from a hot stationary spherical particle kept at a fixed

549

temperature

550

To test the implementation of transient scalar diffusion, the heat conduction

551

from a stationary sphere in a stationary fluid is simulated. The spherical particle

552

of radius Rp at constant temperature Ts is positioned at the center of a cubical

553

domain containing a fluid with initial temperature Tf where Ts > Tf . The

554

simulation results will be compared with the analytical solution of the unsteady

555

state heat conduction equation in an infinite domain. In an infinite domain

556

radial symmetry can be assumed. The governing PDE for the temperature, T

557

as a function of time t and the radial distance from the sphere surface r for this

558

system is given by:   ∂T α ∂ 2 ∂T = 2 r ∂t r ∂r ∂r subject to the initial and boundary conditions: at t = 0

at t > 0

(25)

T = Tf

for r > Rp

(26)

T = Ts

for r ≤ Rp

(27)

T = Ts

for r ≤ Rp

(28)

T = Tf

for r = ∞

(29)

with the solution for temperature distribution and Nusselt number respectively given by    T (r, t) − Tf Rp r − Rp Θ(r, t) = = 1 − erf √ Ts − Tf r 4αt 2 1 αt N u(t) = 2 + √ √ with F o = 2 Rp π Fo

(30) (31)

559

The left hand side of equation (30) signifies the non-dimensional temperature

560

Θ. In equation (31), F o represents the Fourier number. To neglect any bound-

561

ary effects, the domain length is kept as 16 Rp with the particle radius being 22

562

0.5 mm. The thermal diffusivity, α, was set equal to 10−8 m2 /s whereas the time

563

step ∆t was fixed at 10−5 s. The grid was initiated with a root cell uniformly

564

refined to a level 6 (base level, BL) whereas at the sphere surface, refinement

565

levels 9 (BL + 3) to 11 (BL + 5) were used. The Laplace operator 2 (section

566

2.4.2) was used in this case. A Neumann condition was prescribed at the do-

567

main boundaries. For the case of sphere surface at refinement level 10, Figure

568

13(a) shows the profile of the non-dimensional temperature at 0.5 s and Figure

569

13(b) shows the corresponding grid with the particle surface shown as yellow

570

line. Figure 14 shows the radial temperature profiles for three different time

571

instances corresponding to 0.1 s, 0.25 s and 0.5 s. The lines signify the analyti-

572

cal temperature profile obtained from equation 30 whereas the markers signify

573

the simulation results. Table 1 shows the obtained Nusselt number from three

574

different simulations corresponding to maximum refinements near the vicinity

575

of the sphere surface to be BL + 3, BL + 4 and BL + 5 respectively with the

576

associated percentage error from the analytical value derived from equation 31.

577

4.2. Forced convection heat transfer over an in-line array of three spherical par-

578

ticles at low Prandtl number

579

In this section, additional computations are performed to validate the nu-

580

merical model for systems involving both convection and diffusion at moderate

581

Prandtl numbers. Specifically we will focus on the convective heat transfer for

582

an inline array of three spheres. Figure 15 shows the schematic representation

583

of the computational domain for three spherical particles of equal diameter Dp

584

that are separated from each other by a distance s. The parameters used for

585

the simulation are provided in table 2. The entry and the exit regions are char-

586

acterized by distances r1 and r2 respectively. Simulations are performed for two

587

values of center-to-center spacing between the sphere (s/Dp = 2 and 4) and

588

for three different Reynolds number (Re= 1, 10, and 50). The hydrodynamics

589

are solved on a Cartesian grid with (256 × 128 × 128) number of cells in x, y

590

and z direction, respectively. Hence the corresponding adaptive grid for scalar

591

transport is initiated with two trees or root cells in the x−direction uniformly 23

592

refined at level 7 (27 cells in each direction along the edge of one root cell/tree).

593

A constant particle temperature, Ts is enforced using a staircase represen-

594

tation by initially refining the grid in the vicinity of the particle surface to a

595

maximum level of 10. These simulations are performed with the Laplace opera-

596

tor presented in section 2.4.2. On the fly grid adaptivity is thus constrained by

597

level 7 for coarsening and level 10 for refinement. The initial refined cells are

598

also exempted from any coarsening throughout the simulation. The simulation

599

domain is bounded by free slip walls with zero heat flux in the y − z directions.

600

The fluid enters the domain with a constant inlet velocity, U∞ , and constant

601

inlet temperature, T∞ , whereas at the outlet zero heat flux is prescribed. At

602

the surface of the sphere a no slip boundary condition is imposed using the

603

2nd order accurate IBM. Earlier grid independence tests (Deen et al., 2012; Das

604

et al., 2017a) using the same IBM technique have revealed that, the results are

605

essentially grid independent for Dp /∆x = 15 and above, where ∆x is the grid

606

size. Hence, a grid resolution of Dp /∆x = 20 is used for the results presented

607

here. The physical properties of the fluid are chosen such that the resulting

608

Prandtl number is 0.74. This allows for the results to be compared with the

609

values reported in the literature, quantified by the Nusselt number defined as

610

N u = hf Dp /kf . Maheshwari et al. (2006) performed similar simulations with a

611

body fitted grid method using FLUENT and our results are found to be closer to

612

the body fitted grid results (1 − 5%) as compared to those reported by Tavassoli

613

et al. (2013) (5 − 7%) (see table 3). Tavassoli et al. (2013) found the results

614

obtained from 1st and the 3rd sphere to be sensitive (variations up to 15%) to

615

the choice of entrance/exit lengths and the grid resolution needed to resolve the

616

thin boundary layers, especially present near the first sphere.

617

4.3. Forced convection heat transfer for a stationary sphere at high Prandtl num-

618

bers

619

Based on the previous results, it can be concluded that the proposed method-

620

ology produces accurate results at low/moderate Prandtl numbers. In this sub-

621

section, the method’s potential for resolving the heat transfer with thin bound24

622

ary layers at high Prandtl number flows are explored. For this, flow over a hot

623

isothermal sphere is considered, as shown in figure 17. Simulations are per-

624

formed for three different Reynolds number, Re: 20, 100 and 500, defined on

625

the basis of the inlet velocity, U∞ for a particle diameter Dp = 1 mm. The

626

Prandtl numbers considered for each of the aforementioned particle Reynolds

627

numbers are 1, 10, 100 and 500. The change in Reynolds number is achieved by

628

changing the inlet velocity whereas the change in Prandtl number is achieved

629

by changing the thermal diffusivity. Since, the target is to obtain the Nusselt

630

number for this system without confinement effects, a large domain is needed

631

such that the boundary effects are negligible. The boundary conditions in this

632

problem are similar to those for the in-line array of spheres where the walls in

633

the y and z direction are set to free slip with zero heat flux. At the −x boundary

634

the inlet velocity, U∞ , and inlet temperature T∞ , are prescribed whereas the +x

635

boundary is defined as the outlet with zero derivatives in the normal direction

636

for velocity and temperature. Guardo et al. (2006) studied a similar problem

637

with different inlet sizes with respect to the particle diameter and concluded

638

that wall effects over velocity profile are present up to a domain inlet size of

639

4Dp × 4Dp whereas the wall effects on the temperature profile are observed up

640

641

to a domain inlet size of 2Dp × 2Dp . These inlet sizes were deduced for the Re ≈ 300 and subsequently used for obtaining results for 0.33 < Re < 3300.

642

The particle Reynolds number in this study lies well within this range, hence, for

643

the simulations performed in this subsection a domain size of 6Dp × 6Dp × 18Dp

644

645

is taken.

Thus, the hydrodynamics is solved on a Cartesian grid with 128 × 128 × 384

646

grid cells in the y, z and x directions respectively. Hence, the corresponding

647

adaptive grid has a configuration of three root cells/trees sewed together in the

648

x−direction each with an uniform refinement of level 7. The initial adaptive grid

649

is subsequently refined in the vicinity of the surface of the solid particle such

650

that a zone of thickness δ = Dp /2 is at the finest level which helps in resolving

651

the boundary layers forming at the particle surface. Similar to the previous case,

652

a grid resolution of Dp /∆x = 20 is taken for the Cartesian grid. With the base 25

653

level on the adaptive grid fixed at level 7, the scalar transport is solved for two

654

different maximum levels of refinement at the particle surface, thus resulting in

655

two cases: (a) coarsest level 7 and finest level 9 and (b) coarsest level 7 and

656

finest level 11. For each of the two cases and the sub cases therein, simulations

657

are performed using both the Laplace operators developed in section 2.4.

658

For the case of forced convection, there exists well–established empirical

659

correlations in the literature quantifying the heat transfer in terms of N u as a

660

function of Re and P r. Notable among these are the ones proposed by Ranz and

661

Marshall (1952) and Whitaker (1972) for a stationary spherical particle given

662

respectively as N u = 2.0 + 0.6Re0.66 P r0.33

for

10 < Re < 104 , 0.6 < P r < 380

(32)

N u = 2.0 + (0.4Re0.5 + 0.06Re2/3 )P r0.4 for

3.5 < Re < 7.6 × 104 , 0.7 < P r < 380

(33)

663

Figure 18 shows the results obtained for the two different operators in compar-

664

ison to the Ranz and Marshall (1952) correlation. The points corresponding to

665

Prandtl number of 500 are not included in the graph as it falls outside the range

666

for which the correlations were devised. Table 4 shows in detail the obtained

667

Nusselt number for the two grids and the two Laplace operators for various com-

668

binations of Reynolds and Prandtl numbers. It can be inferred that the results

669

are in good agreement with the correlation(s). It is worth mentioning that the

670

results presented in table 4 should not be comprehended as a grid convergence

671

study since, only the maximum refinement level is increased while keeping the

672

base level refinement fixed. In a true grid convergence analysis for adaptive

673

grids all cells must be flagged for refinement thus keeping the same refinement

674

level difference among the neighbors as done in section 3.2.

675

5. Conclusions

676

In this paper a novel computational strategy is presented to resolve the scalar

677

transport using a dual grid/multiple resolution approach that minimizes the 26

678

computational and memory overhead, required to accurately determine heat and

679

mass transfer coefficients in fluid-particle systems with high Prandtl numbers

680

or Schmidt numbers. Unlike earlier studies (Ostilla-Monico et al., 2015; Gotoh

681

et al., 2012; Chong et al., 2018), our method uses an adaptive grid for the scalar

682

transport. A quadtree (octree in 3D) distributed data structure is used to

683

discretize the domain for scalar transport while a base Cartesian mesh (coarser

684

in size) is used to solve for the hydrodynamics. Different discretizations for the

685

convective and the diffusive fluxes have been presented in this work as they

686

differ from those for a conventional Cartesian mesh with respect to the local

687

neighborhood of a face. A verification test for the convective scheme along with

688

the estimate for numerical diffusion is presented. Subsequently, convergence

689

tests for the two Laplace operators are presented where one of the operators

690

results in symmetric matrices whereas the other results in asymmetric matrices.

691

The corresponding matrix solvers that can be used to solve such a system of

692

simultaneous linear equations are described and it was found that the symmetric

693

Laplace operator produces linear systems that are faster and easier to solve than

694

the asymmetric one.

695

Since, in the current methodology, the solution of the momentum on the

696

adaptive grid is avoided, the details of the initial mapping of the dual grids and

697

its subsequent use in the staggered velocity interpolation under the divergence

698

free constraint are presented. The shared memory communicator facilitates the

699

update of velocity from the hydrodynamics grid to the scalar transport grid with

700

close to zero latency. The verification case for an analytical solenoidal field was

701

performed, from which it can be concluded that, the key flow features which

702

are captured by the hydrodynamic grid are maintained by the interpolation

703

on the refined scalar grid with divergence being close to machine precision.

704

Additionally, the conditions for the adaptation criteria and the update of the

705

children data from the parent data in case of refinement, or vice-versa in case

706

of coarsening was also discussed.

707

With the proposed methodology, validation cases were performed for the

708

case of heat conduction from a hot stationary sphere kept at a fixed tempera27

709

ture where the obtained radial temperature profiles and Nusselt numbers where

710

compared with the analytical solutions available for this system. Subsequently,

711

validations were also performed for the study of blockage effects in forced con-

712

vection heat transfer for an in-line array of three spheres for a low Prandtl

713

number of 0.74. The Nusselt number obtained from the simulations were com-

714

pared with those published in literature using an Immersed Boundary Method

715

on Cartesian grid and from a body-fitted unstructured grid. It was observed

716

that the results obtained from the current methodology were closer to the re-

717

sults published using the body fitted grid in comparison to those published

718

with IBM. Subsequently, simulations were performed for the case of forced con-

719

vection heat transfer for a single stationary sphere at high Prandtl numbers.

720

The Nusselt number obtained from these simulations were compared with the

721

well–established empirical correlations and an excellent agreement was found.

722

Finally we conclude with some comments on the limitations of the method-

723

ology and scope of further improvements. Any distributed dynamic data struc-

724

ture such as tree-based approaches, suffers from limitations related to oblivious

725

cache use, inefficient vectorization, complex load balancing with graph partition-

726

ing methods and performance of matrix solvers. Some of these limitations have

727

been alleviated by using state-of-the-art, well documented open-source packages

728

providing optimal capabilities with respect to the aforementioned aspects.

729

A second drawback arises from the complexity in devising higher order dis-

730

cretization schemes that could be tackled to some extent by using aggressive

731

refinement. Quadtree based grid adaptation falls under the category of isotropic

732

grid refinement, i.e. when a cell is marked for refinement it is refined in all di-

733

rections. Hence, it is not the optimal method to resolve boundary layers which

734

develop near solid boundaries aligned to any of the cardinal directions such as,

735

walls which coincide with the computational domain boundaries. Also, non-

736

deformable and fixed particles presented in this work do not pose an ideal test

737

case problem because competing methods such as body-fitted grids and overset

738

grids take benefit from the static nature of the immersed body and can actually

739

produce excellent grids which is a one-time affair. Vreman (2016) and Vreman 28

740

and Kuerten (2018) propose a methodology using an overset grid for fluid turbu-

741

lence over static and moving no-deformable spherical particles. An interesting

742

problem class where this method will show its full potential involves multiphase

743

flows with moving and deformable interfaces, specifically for gas-liquid systems

744

(drops and bubbles) where the Schmidt numbers are high. A quadtree based

745

grid adaptation would provide cheaper and more efficient remeshing along with

746

properly resolving the extremely thin mass boundary layers occuring at such

747

interfaces in multicomponent chemical systems.

748

Acknowledgments

749

This work is part of the Industrial Partnership Programme i36 Dense Bubbly

750

Flows that is carried out under an agreement between Akzo Nobel Chemicals

751

International B.V., DSM Innovation Center B.V., Sabic Global Technologies

752

B.V., Shell Global Solutions B.V., Tata Steel Nederland Technology B.V. and

753

the Netherlands Organisation for Scientific Research (NWO).

754

This work was carried out on the Dutch national e-infrastructure with the

755

support of SURF Cooperative.

756

References

757

Aboulhasanzadeh, B., Thomas, S., Taeibi-Rahni, M., Tryggvason, G., jun 2012.

758

Multiscale computations of mass transfer from buoyant bubbles. Chemical

759

Engineering Science 75, 456–467.

760

761

Andriˇcevi´c, R., Galeˇsi´c, M., 2018. Contaminant dilution measure for the solute transport in an estuary. Advances in Water Resources 117, 65 – 74.

762

Burstedde, C., Wilcox, L. C., Ghattas, O., 2011. p4est: Scalable algorithms

763

for parallel adaptive mesh refinement on forests of octrees. SIAM Journal on

764

Scientific Computing 33 (3), 1103–1133.

29

765

Centrella, J., Wilson, J. R., 1984. Planar numerical cosmology. II - the difference

766

equations and numerical tests. Astrophysical Journal Supplement Series 54,

767

229–249.

768

Chong, K. L., Ding, G., Xia, K., 2018. Multiple-resolution scheme in finite-

769

volume code for active or passive scalar turbulence. Journal of Computational

770

Physics 375, 1045 – 1058.

771

Das, S., Deen, N. G., Kuipers, J. A. M., 2017a. A DNS study of flow and heat

772

transfer through slender fixed-bed reactors randomly packed with spherical

773

particles. Chemical Engineering Science 160, 1 – 19.

774

Das, S., Deen, N. G., Kuipers, J. A. M., 2017b. Immersed boundary method

775

(IBM) based direct numerical simulation of open-cell solid foams: Hydrody-

776

namics. AIChE Journal 63 (3), 1152–1173.

777

Deen, N. G., Kriebitzsch, S. H., van der Hoef, M. A., Kuipers, J. A. M., 2012.

778

Direct numerical simulation of flow and heat transfer in dense fluidparticle

779

systems. Chemical Engineering Science 81, 329 – 344.

780

Dixon, A. G., Nijemeisland, M., Stitt, E. H., 2006. Packed tubular reactor

781

modeling and catalyst design using computational fluid dynamics. In: Marin,

782

G. B. (Ed.), Computational Fluid Dynamics. Vol. 31 of Advances in Chemical

783

Engineering. Academic Press, pp. 307 – 389.

784

Falgout, R. D., Yang, U. M., G., H. A., Tan, C. J. K., Dongarra, J. J., 2002.

785

hypre: A library of high performance preconditioners. In: Computational

786

Science — ICCS 2002. Springer Berlin Heidelberg, Berlin, Heidelberg, pp.

787

632–641.

788

Gada, V. H., Sharma, A., 2011. On a novel dual-grid level-set method for two-

789

phase flow simulation. Numerical Heat Transfer, Part B: Fundamentals 59 (1),

790

26–57.

30

791

Gibou, F., Fedkiw, R. P., Cheng, L., Kang, M., 2002. A second–order–accurate

792

symmetric discretization of the poisson equation on irregular domains. Jour-

793

nal of Computational Physics 176 (1), 205 – 227.

794

Gotoh, T., Hatanaka, S., Miura, H., 2012. Spectral compact difference hybrid

795

computation of passive scalar in isotropic turbulence. Journal of Computa-

796

tional Physics 231 (21), 7398 – 7414.

797

Guardo, A., Coussirat, M., Recasens, F., Larrayoz, M., Escaler, X., 2006. CFD

798

study on particle–to–fluid heat transfer in fixed bed reactors: Convective

799

heat transfer at low and high pressure. Chemical Engineering Science 61 (13),

800

4341–4353.

801

Isaac, T., Burstedde, C., Wilcox, L. C., Ghattas, O., 2015. Recursive algo-

802

rithms for distributed forests of octrees. SIAM Journal on Scientific Comput-

803

ing 37 (5), C497–C531.

804

Lipnikov, K., Morel, J., Shashkov, M., 2004. Mimetic finite difference methods

805

for diffusion equations on non–orthogonal non–conformal meshes. Journal of

806

Computational Physics 199 (2), 589 – 597.

807

Losasso, F., Fedkiw, R., Osher, S., 2006. Spatially adaptive techniques for level

808

set methods and incompressible flow. Computers & Fluids 35 (10), 995 – 1010.

809

Maheshwari, A., Chhabra, R., Biswas, G., 2006. Effect of blockage on drag

810

and heat transfer from a single sphere and an in-line array of three spheres.

811

Powder Technology 168 (2), 74 – 83.

812

Ostilla-Monico, R., Yang, Y., van der Poel, E., Lohse, D., Verzicco, R., 2015. A

813

multiple resolution strategy for direct numerical simulation of scalar turbu-

814

lence. Journal of Computational Physics 301, 308 – 321.

815

Popinet, S., 2003. Gerris: a tree-based adaptive solver for the incompress-

816

ible euler equations in complex geometries. Journal of Computational Physics

817

190 (2), 572 – 600.

31

818

819

Ranz, W. E., Marshall, W. R., 1952. Evaporation from drops. Chemical Engineering Progress 48 (3), 141–146.

820

Stadler, G., Gurnis, M., Burstedde, C., Wilcox, L. C., Alisic, L., Ghattas, O.,

821

2010. The dynamics of plate tectonics and mantle flow: From local to global

822

scales. Science 329, 1033–1038.

823

Tavassoli, H., Kriebitzsch, S., van der Hoef, M., Peters, E., Kuipers, J., 2013. Di-

824

rect numerical simulation of particulate flow with heat transfer. International

825

Journal of Multiphase Flow 57, 29 – 37.

826

Verma, S., Blanquart, G., 2013. On filtering in the viscous-convective subrange

827

for turbulent mixing of high schmidt number passive scalars. Physics of Fluids

828

25 (5), 055104.

829

Vreman, A. W., 2016. Particle-resolved direct numerical simulation of homoge-

830

neous isotropic turbulence modified by small fixed spheres. Journal of Fluid

831

Mechanics 796, 40–85.

832

833

Vreman, A. W., Kuerten, J. G., 2018. Turbulent channel flow past a moving array of spheres. Journal of Fluid Mechanics 856, 580–632.

834

Weiner, A., Bothe, D., 2017. Advanced subgrid-scale modeling for convection-

835

dominated species transport at fluid interfaces with application to mass trans-

836

fer from rising bubbles. Journal of Computational Physics 347, 261 – 289.

837

Whitaker, S., 1972. Forced convection heat transfer correlations for flow in pipes,

838

past flat plates, single cylinders, single spheres, and for flow in packed beds

839

and tube bundles. AIChE Journal 18 (2), 361–371.

840

Xu, H., Xing, Z., Wang, F., Cheng, Z., 2019. Review on heat conduction, heat

841

convection, thermal radiation and phase change heat transfer of nanofluids in

842

porous media: Fundamentals and applications. Chemical Engineering Science

843

195, 462 – 483.

32

844

Zhong, J., Cai, X.-M., Bloss, W. J., 2016. Coupling dynamics and chemistry

845

in the air pollution modelling of street canyons: A review. Environmental

846

Pollution 214, 690 – 704.

33

Table 1: Computed instantaneous Nusselt numbers for the case of heat conduction from a stationary sphere for three different spatial resolutions with the coarsest level of refinement (base level, BL) is 6 and the finest level of refinement at the vicinity of the sphere surface is 9 (BL + 3), 10 (BL + 4) and 11 (BL + 5).

t (s)

Fo

N u (BL + 3)

N u (BL + 4)

N u (BL + 5)

N u analytical

0.01

0.0004

64.23 (+9.96%)

59.89 (+2.53%)

58.78 (+0.65%)

58.41

0.02

0.0008

44.58 (+6.42%)

42.46 (+1.36%)

42.04 (+0.38%)

41.89

0.03

0.0012

35.87 (+3.76%)

35.0 (+0.69%)

34.62 (+0.16%)

34.57

34

Table 2: Parameters used in the simulation of heat transfer from an inline array of three spheres kept at a constant temperature.

s/Dp 2 4

nx × ny × nz

Dp (m)

Dt /Dp

r1 /Dp

r2 /Dp

Pr

−4

16

5

5

0.74

256 × 128 × 128

10−4

16

5

5

0.74

256 × 128 × 128

10

35

Table 3: Nusselt number obtained for the case of forced convection over in-line array of three spheres.

36

847

Table 4: Nusselt number obtained from simualations for forced convection heat transfer from a stationary single sphere at different combinations of Reynolds and Prandtl numbers compared with the empirical correlations derived from experiments. L denotes the coarsest level of refinement (level 7) corresponding to the Cartesian grid on which hydrodynamics is solved whereas L + 2 and L + 4 denotes the finest level of refinement which is in the vicinity of the particle surface.

Obtained Nusselt Number, N u Pr

1

10

100

500∗



Re

Laplace Operator 1

Laplace Operator 2

Correlations Nu

Nu

(eq. 32)

(eq. 33)

L+2

L+4

L+2

L+4

(∇2(1,2) )

(∇2(1,4) )

(∇2(2,2) )

(∇2(2,4) )

20

4.77

4.76

4.75

4.73

4.68

4.23

100

8.16

8.15

8.17

8.16

8.00

7.29

500

15.89

15.87

15.91

15.84

15.41

14.72

20

8.19

7.89

8.18

7.89

7.73

7.60

100

15.75

15.14

15.64

15.13

14.82

15.29

500

32.86

31.55

32.81

31.48

30.68

33.96

20

15.45

14.74

15.42

14.68

14.26

16.07

100

32.10

30.42

32.15

30.36

29.42

35.39

500

69.34

65.52

69.41

65.47

63.32

82.28

20

25.12

23.66

25.09

23.64

22.86

28.79

100

53.56

50.37

53.59

50.33

48.64

65.57

500

118.26

110.94

117.97

110.86

106.30

154.83

outside the limit of correlations

37

Ghost cell centre

X Face Velocity,ux

Internal cell centre,ϕ

Y Face Velocity,uy

Invalid Tree

5

Level 0 4

10

11

14

15

3

8

9

12

13

2

2

3

6

Level 1

(b)

0

0

1

4

0

y 0

1

(a)

2

3

Level 2

9

4

6 1

6

1

5

0

x

9

7

8 1

8

7

7 4 2

Level 3

5 3

2

3

4

5

Valid Tree

5

(c)

(d)

Figure 1: Schematic diagram of spatial discretization (a) Dual grid mapping: background Cartesian grid with ghost cells and overlapping uniformly refined quadtree grid at initial time step (b) Tree violating 2:1 balance criteria (c) Balanced tree with quadrant numbers (d) Hierarchical layout of the tree showing the quadrant numbers and the corresponding levels. The dashed zig-zag line in (a) and (c) show the sequence in the cell numbering provided by the z-ordering.

38

Start

Hydrodynamics (Cartesian Grid)

Scalar Transport (Adaptive Grid)

Cartesian Grid Memory allocation

Adaptive Grid Memory allocation µ

Initialisation Project non-solenoidal velocity u∗x , u∗y , u∗z Calculate pressure corrections & Solve Pressure Poisson eq.

c

Initialise uniform refinement

µ

Mapping Cartesian ò Quadtree Zid → [i, j, k] µ

c

c

µ

Initial Adaptivity Read ux , uy , uz

Use solenoidal interpolation to update velocities on quadtree

Calculate solenoidal velocity & Update ux , uy , uz I/O Operations & Update time, t

no

Is t ≥ tend ? yes

Hydrodynamics (Other V ariables) Shared Memory ux [ ], uy [ ], uz [ ] Scalar Transport (Other V ariables) µ

c

subcycling

M emory M ap

Update gradients of scalar field Adapt the grid Calculate the convective fluxes & fill implicit diffusion matrix & solve I/O Operations & Update time, t

Exit µ: Lock for reading shared memory, when locked scalar transport code cannot read the velocity field µ: Lock for writing shared memory, when locked hydrodynamics code cannot write the velocity field

Figure 2: Flowsheet showing the processing sequence and the memory map related to the dual grid method.

39

Figure 3: Computing the face value of scalar for (a) uniform Cartesian grid (b) adaptive quadtree grid.

40

φS2 f S2 ∇d φS2

φC1 C1

∇fd φS1

∇fd φ φC2 C2

∇fd φL

φS1 S1

(I)

L (II)

Figure 4: Schematic showing two face morphologies that arise for the discretization of gradient at a face (a) face being shared by same size cells (b) face being shared by different size cells thus having a hanging node at the face.

41

φ3 N⊥U

OC

Nb

C φ7

∇fd φ

φ

φ6

φ4

N⊥L

Figure 5: Sample stencil for face centered gradient calculation in the case where opposite cell(OC), lower perpendicular cell(N⊥L ) and upper perpendicular cell(N⊥U ) is not a leaf cell.

42

Figure 6: Divergence free velocity interpolation during (a) Grid adaptation (b) Velocity update from hydrodynamics grid.

43

Figure 7: Convection of a step profile using First Order Upwind (shown in green) and Barton (shown in red) scheme on a (a) uniform grid at refinement level 7 ( (b) adaptive grid with coarsest level 7 and finest level 8 ( (

), 9 (

) and level 8 ( ), 10 (

) at the discontinuity. Exact solution is shown as black dash dotted line (

44

),

) and 11 ).

Figure 8: Convection of a circular blob in an inclined flow field (a) initial scalar distribution and the associated grid. The scalar distribution and the associated grid with the contours for dimensionless scalar value of 0.5 (yellow) and 0.001 (blue) obtained at 0.5 s from (b) Gerris flow solver and (c) Current implementation.

45

Figure 9: Domain setup for the convergence test of the Laplace operators for the pressure Poisson problem along with the solution(pressure) field (a) Cartesian grid of level L = 7 comprising of 27 × 27 cells (b) Coarse-fine grid obtained by successively refining a circular zone twice to obtain a grid with maximum refinement level of L = 9 and minimum refinement level of L = 7.

46

Figure 10: Order of convergence for the Laplacian operator for the case of (a) operator 1 on adaptive (b) operator 2 on adaptive grid (c) central differencing scheme on uniform grids. Dashed line (

) represents the convergence of order 2.

47

Figure 11: Error norms and divergence for staggered grid interpolation corresponding to different levels of refinement ((

) corresponds to order 2 and (

1).

48

) corresponds to order

Figure 12: Velocity vector plot at 10th level of refinement obtained from staggered grid interpolation of analytical flow field at 8th level of refinement.

49

Figure 13: (a) Non-dimensional temperature (Θ) profile of heat conduction from a static particle at 0.5 s and (b) the corresponding adaptive grid with the particle surface shown with yellow line.

50

Figure 14: Non-dimensional temperature (Θ) profile in the vicinity of a stationary particle along the x-axis for three different times (inset graph provides a zoomed view). Lines denote the exact solution whereas markers denote the numerical result.

51

z

r1

s

s

r2

Outlet

y Dt

Inlet

x

Figure 15: Schematic of case setup for forced convection heat transfer over an inline array of spheres.

52

Figure 16: Non dimensional fluid temperature around spheres with a constant surface temperature for the case of Re = 10 and (a) s/Dp = 2 (b) s/Dp = 4, (c) the associated adaptive grid corresponding to the case of s/Dp = 2.

53

Figure 17: Schematic of case setup for forced convection heat transfer over a solid stationary sphere.

54

Figure 18: Nusselt number obtained from single particle simulations for various cases benchmarked with the Ranz-Marshall (Ranz and Marshall, 1952) correlation for forced convection over a heated sphere.

55

Fully Resolved Scalar Transport for High Prandtl Number Flows using Adaptive Mesh Refinement Highlights    

Novel multiple resolution technique is proposed to resolve scalar boundary layers Comparison of second order Laplacian operators for quadtrees Solenoidal velocity interpolations and Z-order curve for dual grid communications Accurate prediction of Nusselt number is obtained for single/multiple particles