You are on page 1of 17

Computers and Fluids 156 (2017) 247–263

Contents lists available at ScienceDirect

Computers and Fluids


journal homepage: www.elsevier.com/locate/compfluid

Scrutinizing lattice Boltzmann methods for direct numerical


simulations of turbulent channel flows
M. Gehrke∗, C.F. Janßen, T. Rung
Institute for Fluid Dynamics and Ship Theory (M-8), Hamburg University of Technology (TUHH), Am Schwarzenberg-Campus 4 (C), 21073 Hamburg,
Germany

a r t i c l e i n f o a b s t r a c t

Article history: The paper reports on the predictive performance of Lattice Boltzmann methods in turbulent channel
Received 31 March 2017 flows. Attention is confined to model-free (direct) numerical simulations at Reτ = 180 using essentially
Revised 1 June 2017
different collision models, i.e. the Bhatnagar–Gross–Krook (BGK), the Multiple-Relaxation-Time (MRT) and
Accepted 4 July 2017
the Cumulant model. The three approaches are assessed by a comparison of the predicted mean flow and
Available online 13 July 2017
turbulence statistics against benchmark Navier–Stokes solutions. Initial studies employ a fine isotropic
Keywords: grid which resolves all relevant scales. Subsequently, a sequence of four gradually decreasing resolutions
Lattice Boltzmann method is utilized to analyze the sensitivity of the collision models for an inadequate resolution. Moreover, the
Scale-resolving simulation influence of the Mach number and the discretization on the predictive accuracy of the weakly compress-
Turbulent channel flow ible LBM framework is briefly addressed.
Collision model
Whilst the Mach number influence is negligible below Ma < 0.1, and the predictive agreement with ref-
erence data is generally satisfactory in conjunction with all employed discretizations for the fine grid,
significant disparities between the collision models are observed when the mesh is coarsened. The BGK
approach offers an improved predictive agreement with Navier–Stokes results for an adequate resolution,
which comes at the expense of severe (formerly reported) stability issues that emanate from the inter-
face between the buffer and the log-layer for an under-resolved flow. Both the MRT and the Cumulant
model are more demanding with respect to grid convergence. As opposed to the BGK and the MRT model,
the Cumulant model remains stable over a wide range of resolutions. The reason can be attributed to a
rigid alignment between predicted streamwise two-point correlation lengths and the grid spacing, which
augments the damping of sub-grid scales. The latter is deemed to be an inherent feature of the model.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction When attention is given to complex domains requiring inho-


mogeneous and anisotropic grids or the formulation of physi-
The Lattice Boltzmann Method (LBM) is a promising alternative cally complex boundary conditions for dynamically coupled multi-
for scale-resolving flow simulations. Related benefits of the method physics simulations, LBM appears less handy as classical contin-
are associated with its favorable computational efficiency, which is uum mechanics strategies. Confining the interest to scale resolv-
greatly supported by the strict separation of nonlocal and nonlin- ing simulations of isotropic small-scale phenomena, however, LBM
ear terms. Moreover, LBM is known to provide an excellent frame- seems very attractive. Direct numerical simulations of turbulent
work for parallel applications on shared and distributed memory shear flows are a relevant case in point and the ability of an LBM
CPU and GPU systems [1–4]. Hence, massively parallel LBM appli- approach to accurately mimic the related physics is of interest.
cations often reflect both, an impressive scalability up to many ten A number of theoretical and numerical investigations on scale-
thousand cores and short turnaround times due to large node up- resolving simulations of turbulence with an LBM framework have
date rates per second in the order of 107 /core for a CPU application been published, e.g. [5–13]. Although many studies reported a fair
or 5 × 105 /core for a GPU application. predictive accuracy, a significant lack of clarity remains as regards
the influence of the collision model. The analogies between kinetic
theories and turbulence modeling are outlined e.g. in [14,15] and
revealed the significance of an effective relaxation in the frame-

Corresponding author. work of Boltzmann kinetic equations as a replacement of the tradi-
E-mail addresses: martin.gehrke@tuhh.de (M. Gehrke), christian.janssen@tuhh.de tional effective viscosity concept. Moreover, the sensitivity of both,
(C.F. Janßen), thomas.rung@tuhh.de (T. Rung).

http://dx.doi.org/10.1016/j.compfluid.2017.07.005
0045-7930/© 2017 Elsevier Ltd. All rights reserved.
248 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

the quality of the results and the robustness of the algorithm to tions by LBM simulations. It is pointed out that almost all previ-
an insufficient coarse lattice spacing, is of importance for practi- ous LBM studies employed much smaller domains in streamwise
cal simulations and seems debatable depending on the choice of and spanwise direction, when compared to the reference simula-
the collision model. Both aspects become intertwined in many for- tions MKM-99, to reduce the computational cost. However, the re-
mer studies, which motivates the present effort. This study aims duction in the size of the domain affects the turbulence statistics,
to extent the existing body of research by a detailed analysis of especially for the pressure which reflects fairly large spanwise cor-
the performance of three different types of collision models in re- relation lengths, as the MKM-99 data demonstrates. Bespalko et al.
solved and under-resolved conditions. Emphasis is given to turbu- employed a computational domain that spans 12H in streamwise
lent channel flows at Reτ = 180, which carries relevant engineering direction and 4H in spanwise direction. An adequate lattice spac-
physics and can be studied on desktop GPU-systems without com- ing of x+ i
= 2 was used together with a somewhat large Mach
promising on the resolution and the extent of the domain. The in- number of Ma = 0.2. The assessment of results is fairly detailed. It
vestigated collision models refer to the standard single-relaxation- also inheres two-point correlation based data in streamwise and
time model of Bhatnagar–Gross–Krook (BGK) [16], the more so- spanwise direction for different wall distances and supports the
phisticated Multiple-Relaxation-Time (MRT) [17,18] and the Cumu- importance of an adequate resolution and domain size at low Mach
lant model (CUM) [19]. The BGK model is perhaps the most widely number. Another DNS and large eddy simulations (LES) effort for
used standard approach. The MRT model has been suggested in re- wall bounded flows was conducted Wang et al. [9], who compared
sponse to stability issues of the BGK model and the Cumulant ap- the results of their GPGPU implementation of a BGK model with
proach aims to avoid recently reported formal deficits of the MRT the KMM-87 reference data for a channel flow at Reτ = 180. The
model, e.g. violation of Galilean invariance or an insufficient level sensitivity of the results with respect to a variation of the stream-
of hyperviscosity, that seem to deteriorate the predictive perfor- wise and spanwise extent of the computed domain was particu-
mance. Similar to previous studies, the results are assessed by a larly investigated and confirmed the above mentioned results of
comparison with benchmark Navier–Stokes solutions reported by Bespalko et al. [13]. A fair predictive agreement of the averaged tur-
Moser et al. [20] (MKM-99), which were obtained from a spec- bulence intensities, Reynolds shear stress and mean velocity for a
tral numerical method developed by Kim et al. [21] (KMM-87). As resolution of x+ i
= 1.41 was reported, provided that the spanwise
opposed to most former studies, detailed turbulence statistics, in- extent is large enough. Results improved when the streamwise ex-
cluding one-dimensional two-point correlations in streamwise and tent was increased. Unfortunately, no two-point correlations or re-
spanwise direction and the related energy spectra and integral lated data were shown to reveal the reasons for the response to
length scales, are employed for the assessment and supplement the changes of the domain size in detail.
the analysis of conventional averages of mean velocity and turbu- The present effort accordingly complies with the findings of
lence intensities. former studies. The computed domain agrees with the reference
A review of former studies on direct numerical simulations Navier–Stokes simulations KMM-87 (4π H × 2π H × 2H). To ensure
(DNS) of wall-bounded turbulent flows using LBM leads to the small compressibility effects, a baseline Mach number of Ma =
work of Lammers et al. [12], who used a single-relaxation-time 0.05 is used. A sufficiently small lattice spacing is employed on
BGK approach to simulate a turbulent channel flow of height 2H the fine grid that agrees with the Kolmogorov length of x+ i
=2
at Reτ = 180. A compact lateral domain extension of H in con- near the wall [20]. Moreover, different types of collision models
junction with a dense lattice spacing (x+ i
= 1.5) below the Kol- are investigated and the assessment involves a more detailed anal-
mogorov length was used together with a slightly high Mach num- ysis for a sequence of successively coarsened grids. A D3Q19 dis-
ber of Ma ≈ 0.3. LBM results showed impressive agreement with cretization seems reasonable to perform the study with a resolu-
Navier–Stokes data of Moser et al. [20]. A companion study of Lam- tion assigned to the Kolmogorov scale. Mind that the Cumulant
mers [22] also analyzed the influence of the LBM discretization model is investigated in a D3Q27 framework as outlined in [19],
comparing D3Q-15/19/27 models with the result, that the accuracy but results of former studies revealed a minor related influence
and robustness of a D3Q15 model are insufficient and the com- for simpler collision models [8,12] which is briefly verified in this
putational surplus of the D3Q27 approach is hard to justify. Fre- study.
itas et al. [8] performed a more comprehensive comparative study The remainder of the paper is organized as follows:
of different discretizations and collision models when applied to Section 2 outlines the governing equations of the Lattice Boltzmann
a turbulent channel flow at Reτ = 200. To limit the computational framework including a concise description of the investigated col-
effort, the study was performed on a fairly compact domain. The lision models. Section 3 presents the numerical method in brief,
domain size was derived from recommendations of Jiménez and outlines the domain as well as the grids and the employed
Moin [23] on the minimal extension of the domain and covers scaling. Section 4 summarizes the data processing approach for
approximately 0.3π H in spanwise and π H in streamwise direc- the displayed results. Section 5 assesses the predictive quality
tion for a channel of height 2H. The resolution was assigned to by the fine grid studies and aims to provide an insight into the
x +i
= 3, which is slightly coarse to resolve all relevant scales. To coarse grid behavior of the LBM. Final conclusions are drawn in
avoid compressible influences, the Mach number was assigned to Section 6. Throughout the paper, vectors and tensors are symbol-
Ma = 0.05. Good predictive agreement was achieved for the aver- ically denoted in bold face and defined by reference to Cartesian
aged streamwise velocity and RMS velocity-components with the coordinates. The notation uses standard lower case Latin subscripts
BGK model for both, the D3Q19 and the D3Q27 model. The D3Q27 to mark Cartesian tensor coordinates. The velocities in streamwise
model was able to return slightly better RMS-peak values on the (x, x1 ), wall-normal (y, x2 ) and spanwise (z, x3 ) directions are
employed lattice spacing. More important, the study reveals that denoted by u (u1 ), v (u2 ) and w (u3 ). The superscript “ 0 ” refers to
the elaborated moment-based MRT and Cascaded LBE [24] offer no a specific reference location and the “ + ” superscript denotes a
benefits over the BGK model as regards the accuracy and the sta- normalization  with respect to inner parameters, i.e. the friction
bility of the LBM simulations. An interesting study was reported velocity Uτ = |τw |/ρ and the kinematic viscosity ν . Particle
by Bespalko et al. [13] for the turbulent channel flow at Reτ = 180 distribution functions are denoted by fα and carry a Greek sub-
using the BGK collision model for a D3Q19 discretization. The au- script that marks the respective direction of the discrete velocity
thors outline the necessity for a reasonable spanwise extent of space.
the computational domain and a small Mach number to suppress
the frequently observed over-prediction of the pressure fluctua-
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 249

Fig. 1. Three-dimensional discrete velocity spaces for the D3Q19 and D3Q27 model.

2. Governing equations lation one arrives at



m
1 
m
Results of the present study were obtained by solving the lattice δρ (t, x ) = fα (t , x ) resp. u(t , x ) = eα fα (t, x ), (4)
Boltzmann equation (LBE) α =1
ρ0 α =1

fα (t + t, x + t eα ) − fα (t, x ) = α (1) with m indicating the total number of collocation points of the par-
ticular lattice model. In Eq. (4) the density fluctuation δρ is re-
for a set of particle distribution functions (PDF) fα , which describe lated to a reference density ρ 0 and the fluid density ρ is given by
the probability to encounter a particle with velocity eα at time t ρ = ρ0 + δρ . Fluid pressures are obtained from a simple equation
located in point x = (x, y, z )T . The algorithm is based upon a set of state, i.e. p(t, x ) = δρ (t, x ) c√
2
s . The speed of sound is linked to
of discrete particle velocities eα . It employs an equidistant Carte- the lattice velocity c via cs = c/ 3.
sian grid with an isotropic spacing xi and a constant time step
t. The left-hand side of the LBE contains the transient change
and the advection of the PDF, while the discrete collision oper- 2.1. Collision models
ator α on the right-hand side models the particle interactions.
Eq. (1) is based on a discretization of the continuous Boltzmann To assess the predictive accuracy of LBM for direct numerical
equation (BE) in velocity space, space and time. The velocity dis- simulations of turbulent flows, three representative state-of-the-art
cretization converts the BE, which is a complex integro-differential collision models of varying complexity and level of sophistication
equation in space, time and particle velocity space, into a set of were selected.
discrete Boltzmann equations, which merely have to be discretized
in space and time. Several discretizations of the velocity space ex- 2.1.1. Single-relaxation-time model
ist for solving Navier–Stokes equations. In the scope of this paper, The SRT model is the most standard LBM collision operator,
the three-dimensional discrete velocity spaces of the D3Q19 and where the particle distribution functions are driven to an equilib-
D3Q27 model are used (cf. Fig. 1). They introduce 19 resp. 27 direc- rium state with a single (and constant) relaxation time τ . The SRT
tions or particle velocity vectors, which have shown to yield a fair operator reads
balance between computational accuracy and cost. For the D3Q19
model, the matrix eα = eα i composed from the 19 discrete velocity
vectors is defined as
 
0 1 −1 0 0 0 0 1 −1 1 −1 1 −1 1 −1 0 0 0 0
eα ={1...19} i ∈ {x,y,z} = 0 0 0 1 −1 0 0 1 −1 −1 1 0 0 0 0 1 −1 1 −1 . (2)
0 0 0 0 0 1 −1 0 0 0 0 1 −1 −1 1 1 −1 −1 1

The D3Q27 model is an enriched version of the D3Q19 model that t   t neq
supplementary includes the vertices of the unit cuboid as addi- α = − f − fαeq = − f = −t ω fαneq . (5)
τ α τ α
tional collocation points. The additional coordinates of the matrix
composed from the particle velocity vectors read The approach is, to some extent, based on a continuous relaxation
  operator first published by Bhatnagar et al. [16], which is why this
1 1 1 1 −1 −1 −1 −1 model is often called the BGK or lattice BGK (LBGK) model. The
eα ={20...27} i ∈ {x,y,z} = 1 −1 1 −1 1 −1 1 −1 . (3) eq
discrete equilibrium distribution functions fα are approximated by
1 1 −1 −1 1 1 −1 −1 a series expansion of second order related to the velocity of the
continuous Maxwellian distribution function. This yields
If the lattice constant c is chosen consistent with grid spacing and 

time step size, i.e. c = x/t, the discretized Boltzmann equation e ·u 9 (eα · u )2 3u·u
reduces to the simplified form presented in Eq. (1). This implies fαeq = wα δρ + ρ0 3 α 2 + 4
− 2
, (6)
c 2 c 2 c
that a particle (or a PDF, respectively) is advected exactly onto one
of its next neighbors during a discrete time step. in the employed weakly compressible formulation with ρ = ρ0 +
The macroscopic flow parameters are obtained from the mo- δρ and a density fluctuation δρ that is assumed much smaller
ments of the particle distribution functions. The density is given than a reference density ρ 0 . A Chapman–Enskog analysis reveals
by the zeroth order moment and the first order moment yields the that the relaxation time τ is related to the shear viscosity ν via
macroscopic velocity u = (u, v, w )T . For an incompressible formu- τ = 3ν /c2 + t/2 = 1/ω. The weighting factors wi for the selected
250 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

D3Q19 and D3Q27 discretizations read as follows that are related to momentum, i.e. jx , jy and jz , by explicitly adding
⎧1 the body force contribution per time step, e.g. jx = jx + t Fx .
⎪ α=1
⎨3 Even though MRT models show an increase in stability for most
D3Q19 : wα = 1
α ∈ [2, 7] and practical applications and, consequently, are used widely, several

⎩1
18
drawbacks were found: First, the choice of moments is rather ar-
α ∈ [8, 19] bitrary. Thus, no unique moment space and no unique MRT model
⎧ 36
⎪ 27 α = 1
8
⎪ exists. Second, the choice of relaxation rates is ambiguous and sev-

⎪ eral optimization strategies exist. Third, recent analyzes revealed
⎨ 2 α ∈ [2, 7]
27 severe stability issues, at least for some important benchmark
D3Q27 : wα = . (7)

⎪ 1
α ∈ [8, 19] problems, see e.g. [8,19].

⎪ 54
⎩ 1
216
α ∈ [20, 27] 2.1.3. Cumulant model
Note that the relaxation time is not changed locally in the present Recently, more advanced, high-fidelity collision operators were
study, neither by additional viscosity contributions nor by an en- suggested. They are motivated by the lack of a unique procedure to
tropic limiter. Body forces F = (Fx , Fy , Fz )T are applied in a conven- identify optimal collision parameters, the almost arbitrary choice
tional way by adding a contribution to the PDFs in every time step of the moment space and formal deficiencies, e.g. a lack of Galilean
[25,26], i.e. invariance. The Cumulant model is an example of a more elaborate
alternative [19]. First benchmarks of the model were carried out
fα = fα + 3wα t (eα · F ) (8)
in [34–37] and showed promising results, in terms of stability and
The BGK model is perhaps the most widely used approach. As accuracy, compared to existing alternatives. The model is based on
the model features only one free parameter, the shear viscosity a transformation from phase space referred to as ξ = (ξ , υ , ζ )T to
is linked to the bulk viscosity of the fluid. While its implementa- frequency space denoted  = (, ϒ , Z )T by a two-sided Laplace
tion is rather trivial, the model is deemed to offer an unsatisfactory transform
amount of stability for practical applications. However, recent the-  +∞
oretical and numerical work [8,19] shows that the SRT might also L { f (ξ , υ , ζ )} = F (, ϒ , Z ) = e − ξ · f ( ξ , υ , ζ ) d ξ . (10)
−∞
perform more stable and accurate than more sophisticated models.
Cumulants are defined as
2.1.2. Multiple-Relaxation-Time model 
∂α∂β∂γ 
MRT models have been suggested in response to the stability c α β γ = c −α −β −γ ln ( F ( , ϒ , Z ))  . (11)
issues of the BGK model. The MRT model is based on a transfor- ∂ α ∂ ϒ β ∂ Zγ =ϒ =Z=0
mation from particle distribution space to moment space, where Herein α , β , γ ∈ (0, 1, 2) represent the order of each cumulant.
the actual relaxation takes place. The moment space is spanned by The formalism leads to a total of 27 cumulants with (α , β , γ ) =
linear combinations of PDFs. It offers the advantage of assigning (2, 2, 2 ) as the highest occurring order of 6. Due to the complex-
single moments to physical equivalents. This allows to distinguish ity of the expression, the determination of cumulants directly from
between moments related to hydrodynamics and other quantities. (11) is expensive. A simpler approach is to initially compute central
The general form of the D3Q19 MRT collision operator reads moments and obtain the cumulants from combinations of those
   

19 
19 
19 central moments. A simplified equation to generate central mo-
α = − −1
Mαβ Sβγ Mγ δ f δ − meq
γ . (9) ments reads
β =1 γ =1 δ =1 
kα β γ = (m − u/c ) α (n − v/c ) β (o − w/c ) γ fm n o, (12)
The matrices Sαβ and Mαβ are diagonal matrices, which in- m,n,o
here the particular relaxation rates, and a linear transformation
with (α , β , γ ) ∈ (0, 1, 2) indicating the order of the central mo-
matrix, respectively. Within the MRT model, the relaxation of hy-
ment, (m, n, o) ∈ (−1, 0, 1 ) denoting the collocation point of the
drodynamic moments is physically motivated and similar to the
underlying D3Q27 lattice and the macroscopic velocity vector u =
BGK model. The non-hydrodynamic moments can be relaxed in-
(u, v, w )T . The collision in cumulant space
dividually. Their relaxation does not affect the convergence to
the Navier–Stokes equations on the leading order, but influences c̄α β γ = cαeqβ γ + cα β γ (1 − ωα β γ ) (13)
higher-order contributions that either yield diffusive or dispersive
eq
errors and aims to increase the methods accuracy and stability. The describes the relaxation towards an equilibrium state cα β γ of the
transformation into moment space is invariable and linear, hence cumulants with relaxation frequency ωα β γ . The bar expresses the
no complex matrix-vector operations are required. The computa- post collisional state. Only one relaxation frequency is linked to the
tional effort of MRT and SRT models is therefore comparable. shear viscosity of the underlying fluid. All other relaxation param-
Several different MRT formulations exist, which are distin- eters can be freely chosen in the range of [0, 2] to have a positive
guished by the definition of the transformation matrix Mαβ and impact on the accuracy and the stability of the operator.
the choice of collision rates. The present model refers to Toelke The model is solely derived for the D3Q27 velocity space in the
et al. [18]. The employed transformation matrix Mαβ , equilibrium current version. Different from the employed BGK and MRT mod-
eq
moments mα and diagonal matrix Sαβ can be found in Appendix A. els, the Cumulant model is a fully compressible model, yielding a
Opposite to the original MRT of d’Humières et al. [17], the ba- zeroth order moment that directly corresponds to the fluid density
sis vectors adhere to a weighted rather than an unweighted or- ρ . Body forces are again added when switching from raw moment
thogonality [27,28]. The model has been successfully applied to space to central moments, as the relaxation is done in a moving
high-Reynolds number and free surface flows in an LES framework frame of reference, whereby e.g. u = u + t/2Fx /ρ .
[18,29–33]. For the choice of collision parameters related to non- Similar to MRT, the relaxation rates related to the higher-order
hydrodynamic moments, different values are reported in the lit- Cumulants are set to unity, as recommended by the original work
erature. In the present work, all resulting five independent relax- of Geier et al. [19]. As this basic model is deemed superior to the
ation rates are assigned to one to limit the parameter space. Body MRT, the influence of further optimization is not in the scope of
forces Fi are added in moment space by modifying the moments this work.
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 251

2.1.4. Model coefficients


The employed MRT model aims at minimizing the non-
equilibrium influence and the employed Cumulant model uses the
recommended baseline relaxation rate parameters [19]. A particu-
lar advantage of the LBM is the possibility to tune the accuracy
and/or the stability of the algorithm from the choice of the phys-
ically not stipulated relaxation rates. Typically the convergence of
the LBM model towards target solutions, i.e. Stokes, Navier–Stokes
flow or steady convection-diffusion models, following a modified
equation analysis is related to the relaxation rate values.
Examples for alternative optimized coefficients of MRT models Fig. 2. Illustration of the simulated box domain.
aiming at improved accuracy (and stability) as compared to lin-
earized Navier–Stokes equations are reported by Xu and Sagaut Table 1
[38] for a two-dimensional environment, using a matrix pertur- Physical resolution, LBM viscosity, amount of grid points for the investigated
bation theory in conjunction with a modified equation approach, grids and collision models employed (Note that ν ≡ 1.5 × 10−5 , λv ≡ 33.52 and
Ma ≡ 0.05).
or by Dubois et al. [39,40] using Taylor expansions. Other recent
optimization attempts were concerned with the analysis of steady No. x+i xi [m] t [s] ν LBM NP Models
flows [41] often in the context of a two-relaxation-time model 1 2.0 5.61 × 10−4 1.67 × 10−5 7.97 × 10−4 115 816 960 all
(TRT) [42–44]. 2 3.2 8.98 × 10−4 2.68 × 10−5 4.98 × 10−4 26 500 140 all
The influence of the relaxation rate on the predictive perfor- 3 6 16.84 × 10−4 5.02 × 10−5 2.66 × 10−4 4 547 328 all
4a 7 19.64 × 10−4 5.86 × 10−5 2.28 × 10−4 2 903 340 BGK
mance is evident. It is however noted, that so far no alternative
4b 8 22.45 × 10−4 6.70 × 10−5 1.99 × 10−4 1 989 980 MRT,CUM
relaxation rate parameters were published for the Cumulant model 5 10 28.06 × 10−4 8.37 × 10−5 1.59 × 10−4 1 040 288 CUM
and none of the mentioned MRT-suggestions precisely addresses
the flow situation of the present study. Moreover, literature re-
ported results, e.g. [45] indicate that parameter choices derived
from stability and accuracy aspects do not necessarily agree and a U τ denotes a spatially and temporally averaged target friction ve-
unique best choice is thus not available. We therefore restrict our- locity that should be met by means of an appropriate forcing.
selves to three baseline approaches. The resolution of Kolmogorov scales requires a lattice spacing of
roughly two wall units [20,21], thus an adequately fine grid can
be constructed from a lattice spacing of H/xi > 90. The employed
3. Computational model homogeneous isotropic grid consists of NY = 184 nodes in normal,
NX = 1120 nodes in streamwise and NZ = 562 nodes in spanwise
The numerical framework is based upon the efficient Lattice direction. As opposed to the reference studies, the present mesh is
Boltzmann environment elbe [31]. The versatile toolkit can address significantly finer in the streamwise and the spanwise direction,
complex flow problems with and without a free surface including where homogeneous lattice distances of x+ = 10 and z+ = 5
effects of turbulence and fluid-structure interaction. Either homo- are often deemed sufficient. The wall-normal resolution of alterna-
geneous or inhomogeneous (overlapping) grids are employed. The tive methods is usually realized with inhomogeneous meshes that
code offers the use of various collision models and is dedicated to stretch from near-wall spacings around y+ = 0.1 to y+ = 5 at
parallel GPGPU-machines, both in shared and distributed memory the horizontal midplane.
mode. Following the fine-grid studies, successively coarser meshes us-
The present study utilized a shared memory desktop system of ing NY = 111 (x+ i
≈ 3.2), NY = 62 (x+i
≈ 6), NY = 47 (x+i
≈ 8)
+
four NVIDIA-K40 GPUs, each featuring 12GB of memory and almost and NY = 38 (xi ≈ 10) were investigated to assess the sensitivity
2500 CUDA processor cores, in combination with simple homoge- of LBM results to the lattice spacing (cf. Table 1). As some of the
neous isotropic grids. The computation performance depends on collision models became unstable when the grid was coarsened,
the employed flow model, grid type and communication model. intermediate grids were used to assess their limits. Mind that the
In the present study, a performance of approximately 1400 mil- near-wall distance of the first fluid node always refers to half the
lion node updates per second (MNUPS) was achieved, which yields respective lattice spacing.
about 90 h of wall-clock time on a desktop system to perform the
4 million discrete time steps of the fine grid computation. 3.2. Physical units and scaling

3.1. Test case, computational domain and grid layout Spatial and temporal unity spacing, i.e. xLBM
i
= 1 and t LBM =
1, is required to comply with the governing Eq. (1) of the LBM.
The flow geometry and the coordinate system are illustrated in Using a channel height of 2H = 0.1021m, NY = 184 points in wall-
Fig. 2. The box-type computational domain is chosen in line with normal direction, and a wall location at half a lattice distance,
the study of KMM-87. The height of the domain is 2H. It spans the physical spacing reads xi = 2H/(NY − 2 ) = 5.61 × 10−4 m. Due
approximately 4π H in streamwise direction and 2π H in span. The to the unit spacing convention, the length scaling refers to l =
study is restricted to fully developed flow conditions and assumes 5.61 × 10−4 m.
periodic boundaries in streamwise and spanwise directions. The A small Mach number is a prerequisite for a fair convergence
box has been designed to compute the largest eddies without any of the weakly compressible nature of LBM to the Navier–Stokes
influence of the periodic boundaries. The employed wall bound- equations. A target mean centerline velocity U c is selected as the
ary conditions refer to simple bounce-back [46] conditions which reference (upper) velocity of the flow field. Baseline results were
– with the exception of a slight error – assume the wall location obtained from a Mach number choice of Ma = √ U c /cs = 0.05. The
at half the lattice spacing. speed of sound is usually assigned to csLBM = c/ 3 in LBM units,
LBM
Fine-grid computations were performed on a mesh with ap- hence the centerline velocity in LBM units follows from U c =
proximately 116 million lattice nodes. The targeted Reynolds num- Ma csLBM = 0.0289. The target mean friction velocity in LBM units is
ber based on the friction velocity reads Reτ = U τ · H/ν = 180. Here obtained from correlations reported by MKM-99, i.e. U c /U τ = 18.3,
252 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

LBM with xi denoting the axes of coordinates x, y and z provide an in-


which provides U τ = 1.58 × 10−3 and thus a viscosity of ν LBM =
7.97 × 104 to meet with Reτ = 0.5 (NY − 2 ) U τ /ν LBM = 180. To- sight into the computed structures and the sensitivity of the colli-
gether with an (arbitrary) kinematic viscosity of the fluid – e.g. air sion model to resolution aspects. To avoid cancellation of positive
at ν = 1.5 × 10−5 m2 /s – the physical velocities are recovered from and negative values, the integral lengths are evaluated from their
LBM LBM (positive) origin at δ xi = 0 to the first zero passage of their argu-
U τ = vU τ and U c = vU c with v = ν /(ν LBM l ) = 33.52m/s.
ment.
The density is assigned to unity in physical and LBM units. The
physical time step follows from t = l /v t LBM = 1.67 × 10−5 s.
5. Results
Table 1 summarizes the resolution dependent parameters of the
present study for the baseline Mach number of Ma = 0.05. Sup-
5.1. Fine grid results
plementary, Mach number as well as discrete velocity space influ-
ences were studied by using half and double the reference center-
5.1.1. Adequacy of domain and resolution
line velocities in LBM units for grid No. 2.
The adequacy of the employed fine-grid resolution and the con-
sidered domain is indicated by two-point correlations and en-
4. Computational strategy and data processing ergy spectra displayed in Figs. 3–5. The figures refer to a compar-
ison of MKM-99 data with exemplary results obtained from the
The channel flow is driven by body forces [25,26]. The applied BGK model in conjunction with the two finest grids mentioned in
forcing is resolution dependent and refers to a simple balance Table 1. They depict the predicted normalized two-point correla-
LBM
of friction and pressure forces referring to F x = (−∇x pLBM ) = tions Rx/z x/z
uu and Rvv on the left side and related 1D energy spectra
2 LBM LBM
ρ LBMU τ /(Reτ xi ). For each computation, the initial forcing F x on the right side which were extracted in streamwise (top) and
is iteratively adjusted until stable compliance with the target spanwise (bottom) direction for a fixed wall distance. The two wall
Reynolds number Reτ = 180 has approximately been achieved, i.e. distances shown refer to y+ = 5 and y+ = 79, respectively. The cor-
LBM LBM relations drop down to small values well away from the domain
(UτLBM − U τ ≤ 5%. )/U τ
boundaries and the energy density drops by approximately 4 or-
Subsequent to the adjustment of the forcing, an initial tran-
ders of magnitude between the low and the high wave numbers.
sient period of Ti = 2 × 106 t was computed before time averaged
No accumulation of energy in the high wave number region is indi-
data and statistics have been compiled from a data extraction pe-
cated by the energy spectra. Moreover the results do not differ sig-
riod of Te = 2 × 106 t for the baseline Mach number. The nondi-
nificantly when the resolution is slightly coarsened to x+ = 3.2
mensional extraction and initial transient periods correspond to i
and display a fair level of agreement with the reference data. Mind
Te U τ /H = Ti U τ /H ≈ 35 on the fine grid. Thus, the structures are at
that displayed spectra were filtered with a 3rd-order Savitzky–
least advected 3 times through the entire domain during both pe-
Golay filter using 11 base points for the sake of clarity. The filter-
riods, even based upon a slow sublayer advection at U + = 1. This
ing procedure does not change the result in streamwise direction
cycle was not altered when the grid was coarsened, which implies
but removes minor high wave number oscillations occurring in the
a longer averaging and extraction period at the root of the coarse
spanwise direction, as indicated by a comparison between the raw
grid results and was probably lavish in terms of computational ef-
and the filtered data in Fig. 5.
forts. Statistical convergence has been verified by comparing the
respective time averaged and statistical quantities for Te /5, 2Te /5,
5.1.2. Mean velocity
3Te /5 etc. In this regard, no substantial changes were observed be-
Table 2 outlines the predicted mean flow and log-law param-
yond 3Te /5 for any grid. For consistency reasons, identical initial
eters for the fine-grid results. The companion Fig. 6 displays the
transient and averaging periods have been used for the other two
mean velocity profile in comparison with reference data reported
Mach numbers, which therefore refer to double (Ma = 0.025) and
by Moser et al. [20]. The predictive discrepancy between the differ-
half (Ma = 0.1) the amount of time steps, respectively.
ent collision models is small and the BGK model provides a sur-
Spatial averages, e.g. mean velocity profiles or the friction ve-
prisingly fair agreement with the reference data.
locity, were derived from spatial averages of the time averaged
The parameter Uc /Um shows an almost perfect agreement be-
field. Note that inner parameters indicated by the superscripted
tween the considered data sets. All collision models display a fairly
“ + ” are computed using the individual friction velocity Uτ of the
moderate variation ( < 3%) of the von-Kármán parameter κ , where
respective simulation. Two sequences of 1D two-point correlations
the biggest deviation from the reference data is around 4.2% for the
were obtained in streamwise (x) and spanwise (z) direction for a
Cumulant model. The onset of the turbulent regime displays larger
range of fixed y+ -values by means of time averaged line probes,
variations with respect to each other and the reference data. The
viz.
   maximum difference between the LBM-predicted B-parameters is
Rxui ui x0 , x0 + δ x , y0 , z01...50 = u
i (x0 ) u
i (x0 + δ x ) and (14) slightly more than 16% and the value returned by the Cumulant
model exceeds the reference data by approximately 18%. This can
   be attributed to an enhanced ratio between the centerline velocity
Rzui ui x01...124 , y0 , z0 , z0 + δ z = u
i (z0 ) u
i (z0 + δ z ). (15) and the friction velocity Uc /Uτ for the MRT and particularly the
The analyzed correlations were computed from the (spatial) av-
erages of the 50 streamwise and 124 spanwise line probes ex- Table 2
tracted for each y+ -value and finally normalized with its partic- Computed mean bulk velocity Um , mean centerline ve-
x/z x/z x/z locity Uc , mean friction velocity Uτ and parameters of
ular maximum. Displayed 1D energy spectra Euu , Evv and Eww
+ the log-law U + = (ln y+ )/κ + B for the fine grid in
at fixed values of y were computed from the averaged normal- comparison with data of Moser et al. [20] (MKM-99).
ized streamwise and spanwise two-point correlations. Related plots
were mostly prepared to support the comparison with the MKM- Model BGK MRT Cumulant MKM-99

99 reference data, which host significantly less wave numbers due Um /U τ 15.72 16.07 16.23 15.73
to the much coarser resolution. Moreover, integral lengths defined Uc /Uτ 18.28 18.68 18.89 18.30
Uc /Um 1.16 1.16 1.16 1.16
by
 κ 0.390 0.385 0.396 0.380
  uτ B 5.13 5.43 5.94 5.03
λ +
Rxuii ui = Rxuii ui dxi (16)
xi
ν
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 253

uu and Rvv (left) returned by the BKG predictions for two different resolutions – i.e. xi = 2 (solid), xi = 3.2 (dashed) – and related 1D
+ +
Fig. 3. Two-point correlations Rx/z x/z
x/z x/z
energy spectra Euu and Evv (right) in streamwise (top) and spanwise (bottom) direction for a near-wall position of y+ = 5 in comparison to reference data of Moser et al.
[20] (MKM-99).

uu and Rvv (left) returned by the BKG predictions for two different resolutions – i.e. xi = 2 (solid), xi = 3.2 (dashed) – and related 1D
+ +
Fig. 4. Two-point correlations Rx/z x/z
x/z x/z
energy spectra Euu and Evv (right) in streamwise (top) and spanwise (bottom) direction for y+ = 79 in comparison to reference data of Moser et al. [20] (MKM-99).
254 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

z/x
Fig. 5. Comparison of raw data (dashed line) and filtered data (solid line) for spanwise (left) and streamwise (right) 1D energy spectra Euu computed on the fine grid
(x+i
= 2) for y+ = 5.

the results of other two models, usually closer to the Cumulant


model.
Whilst an increase of peak and log-layer intensities is of mi-
nor importance for the mean flow, a closer look at the near-wall
situation depicted by Fig. 8 reveals much larger differences which
comply with the findings for the mean velocities. The Cumulant
model predictions of near-wall intensities for the streamwise RMS-
values are around 10 − −15% lower as the BGK-predictions and the
respective difference for wall-normal component is more than 30%.
In accord with the large difference in wall-normal direction, the
predicted turbulent shear stress is significantly lower for the Cu-
mulant model in comparison with the BGK result and also reflects
a much smaller shear-exerting gradient near the wall. Mind that
the near-wall agreement of the intensities and the shear stress
with the MKM-99 data is not necessarily the best for the BGK
model, but the LBM results are consistent to each other.

Fig. 6. Predicted nondimensional mean velocity for the fine grid in comparison to
data of Moser et al. [20] (MKM-99). 5.1.4. Two-point correlations & turbulent structures
An interesting point refers to the behavior of the predicted two-
Cumulant model. It appears that the predicted shear stress falls point correlations. Emphasis is given to normalized 1D streamwise
x
below the MKM-99 [20] reference value by more than 3.5% [2%] and spanwise correlations Ruii ui for a sequence of fixed nondimen-
for the Cumulant [MRT] model. For the MRT model, an increase +
sional wall distances y as described in Section 4.
of 2% for Uc+ matches the observed increase of almost 8% of the The occurrence and location of the minima of Rzui ui support the
B-parameter fairly accurate. The data also compares favorably with identification of structures and gives an estimation of their size.
an experimental study of Zaounu et al. [47] which suggests κ = 1/e Similar to the reference data, both spanwise correlations Rzuu and
and B = 10/e in conjunction with a parameter value of e ≈ 2.7 for Rzvv outlined in Fig. 9 display local minima below wall distance val-
large enough Reynolds numbers Reτ ≥ 20 0 0 (i.e. κ = 0.37 and B = ues of y+ < 80. All spanwise correlation minima reach down to
3.7). Similar to the findings of Lammers et al. [12] and Moser et al. negative values between −0.2 and −0.3, in line with findings of
[20] a slightly different relation is found due to the low Reynolds previous researchers [13,20]. The detected Rzvv -minima reveal the
number (e ≈ 2.6, B ≈ 13.2/e). occurrence of streamwise vortices. Their location between z+ ≈ 25
The reason for a lower wall shear stress displayed by the MRT near the wall and z+ ≈ 50 at y+ = 40 gives an impression about
and the Cumulant model can be analyzed when looking at the their diameter. In accordance with formerly reported results, the
single-point and two-point turbulence statistics, where the Cu- spanwise Rzvv -minimum occurs closer to the wall than the Rzuu -
mulant model appears to return a higher correlation length in minimum. The latter marks the separation between the high-speed
streamwise direction and a more pronounced near-wall damping and the low-speed fluid at approximately half the mean streak
as outlined in the next subsections. On the contrary, the single- spacing. Similar to former results, the spanwise Rzww -minimum is
relaxation-time BGK model displays a smaller wall damping and restricted to regions y+ < 30 which does not support the existence
an earlier increase of turbulence motion when leaving the near- of counter-rotating vortex pairs [48]. The shape of the streamwise
wall regime. correlations is much simpler and shows a fair agreement, as indi-
cated by the exemplary results outlined by Fig. 10.
5.1.3. Turbulence intensities and Reynolds stresses In general, the identification of predictive differences from the
The different near-wall behavior of the three investigated colli- data displayed in Figs. 9 and 10 is quite difficult, hence integral val-
sion models is depicted by the turbulence intensities. All intensi- ues might support a distinction of the collision model behavior.
ties shown in Fig. 7 reveal very similar general characteristics and Former analysis, e.g. [49], reveals that the normalized streamwise
closely follow the MKM-99 data. The BGK model provides peak and correlation length of the streamwise velocity peaks at the end of
log-layer intensities which are slightly higher than the Cumulant the buffer layer, where the respective values significantly exceed
results. Values returned by the MRT model are located in between the y+ -values, and subsequently drops in the log-layer. A similar
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 255

Fig. 7. Predicted nondimensional turbulence intensities and turbulent shear stresses in comparison with the data of Moser et al. [20] (MKM-99).

Fig. 8. Near-wall behavior of the predicted nondimensional intensities and turbulent shear stresses in comparison with the data of Moser et al. [20] (MKM-99) for the fine
grid computations.

behavior is found for all other correlations in streamwise direction. other two collision models in the near-wall regime, where the MRT
Fig. 11 compares the integral length scales for the Rxuu - and the Rxvv - and particularly the Cumulant model seem to indicate some hyper-
components obtained from the present study with the reference viscous contributions, but the differences vanish in the core flow.
data reported by Moser et al. [20]. The comparison indicates that Mind that Geier et al. [19] have found a negative hyperviscosity
the BGK model delivers length scales which follow the MKM-99 re- contribution during their analysis of another MRT model variant,
sults very closely for the primary component (Rxuu ) and also display which would explain potential instabilities occurring in the core
a lower peak for the wall-normal component (Rxvv ) that agrees bet- flow regime (cf. Section 5.3).
ter with the MKM-99 results. On the contrary, the MRT model and,
in particular, the Cumulant model display significantly larger peak 5.2. Mach number and discretization influences
values, which are approximately 15% higher than the BGK-result for
the Cumulant model. The generally lower correlation-length val- To verify the reliability of the results, Mach number influences
ues predicted by the BGK model increase the turbulence energy have been studied for the BGK model on grid No. 2 (x+ i
= 3.2).
levels and also support smaller scales. The latter is visible from In general the respective differences obtained from halving or dou-
the smaller decline of the spectral energy at higher wave numbers bling the baseline Mach number of Ma = 0.05 virtually do not alter
displayed in Fig. 12. In line with the more pronounced wall damp- the predictive response. Fig. 13 outlines the predictive differences
z spec-
ing reflected by the MRT and the Cumulant results, the Evv for the normalized mean velocity U + and the RMS-value for the
tra of the BGK model shown in Fig. 12 are above the data of the primary fluctuation u+ RMS
. Disregarding the specific Mach number,
256 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

Fig. 9. Predicted spanwise normalized two-point correlations Rzuu (left), Rzvv (center) and Rzww (right) at different wall-normal distances (top: y+ = 5; middle: y+ = 19; bottom:
y+ = 40) in comparison to reference data of Moser et al. [20] (MKM-99).

Fig. 10. Predicted streamwise normalized two-point correlations Rxuu (left), Rxvv (center) and Rxww (right) at different wall-normal distances (top: y+ = 5; bottom: y+ = 40) in
comparison to reference data of Moser et al. [20] (MKM-99).

Fig. 11. Predicted streamwise turbulent length scales for the normalized two-point correlations Rxuu (left) and Rxvv (right) in comparison to results of Moser et al. [20] (MKM-
99).
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 257

z
Fig. 12. Spanwise 1D energy spectra Evv at wall distances y+ = 5 (left) and y+ = 178 (right) obtained by the fine grid computations in comparison to results of Moser et al.
[20] (MKM-99).

Fig. 13. Predicted differences displayed by the BGK model for the mean velocity U + (left) and the RMS-value of u+
RMS
(right) for the larger (Ma = 0.1) and the smaller
(Ma = 0.025) Mach number when compared with the baseline results at Ma = 0.05 and the reference data (MKM-99) [20] on grid No. 2.

x
Fig. 14. Predicted differences displayed by the BGK model for the streamwise 1D energy spectrum Euu at y+ = 40 (left) and y+ = 178 (right) for the larger (Ma = 0.1) and
the smaller (Ma = 0.025) Mach number when compared with the baseline results at Ma = 0.05 on grid No. 2.

cent regime and increase with the wave number. The differences
for a reduction of the Mach number are generally above the dif-
ference experienced from a Mach number increase, which might
reveal minor room for improvement.
In conjunction with the BGK model, the discretization influence
has been studied in brief on grid No. 2. The benefits of a D3Q27
discretization over the baseline D3Q19 approach are small. As in-
dicated by the mean velocity illustrated in Fig. 15, the results of
the BGK model are shifted a small step towards the Cumulant re-
sults by means of an increased von Kármán parameter κ = 0.394
and a slightly augmented log-law parameter B = 5.4. The differ-
ences originate from a reduction of near-wall RMS and shear stress
values which return a slightly better related agreement with the
MKM-99 reference data (cf. Fig. 16).

5.3. Coarse grid results


Fig. 15. Predicted nondimensional mean velocity obtained for the BGK model on
xi = 3.2 (grid No. 2) for the baseline D3Q19 and the D3Q27 discretization in com- Figs. 17–19 display illustrative exemplary snapshots of the nor-
parison to the MKM-99 data [20]. malized instantaneous ux −velocity component in the xy-symmetry
plane for three different grids. As illustrated by Fig. 17, the BGK
the difference between the LBM and the Navier–Stokes solution is model catches unphysical peaks in the near-wall region. The MRT
in the per mil regime or below, with the largest differences in the model shows oscillatory structures in the core flow regime for grid
buffer layer and the viscous sublayer. Fig. 14 compares the respec- No. 3 (cf. Fig. 18). The Cumulant model displays a successive in-
x extracted
tive differences of the streamwise 1D energy spectra Euu crease of structures with an increase of the lattice spacing, but no
+ +
for y = 40 and y = 178. The respective differences are in the per- unphysical features are found in Fig. 19 at all.
258 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

Fig. 16. Predicted mean velocity and RMS-value differences obtained for the BGK model on xi = 3.2 (grid No. 2) for the baseline D3Q19 and the D3Q27 discretization in
comparison to the MKM-99 data [20].

Fig. 17. Instantaneous normalized streamwise velocity ux snapshots in the xy-


symmetry plane predicted by the BGK model for 3 different grids. Image detail for
x+i ≈ 6 refers to pressure contours that display instabilities more clearly.

Fig. 20. Predicted nondimensional mean velocity for xi = 6 (grid No. 3) compared
with MKM-99 data [20].

Table 3
Computed mean bulk velocity Um , mean centerline velocity Uc ,
mean friction velocity Uτ and parameters of the log-law U + =
(ln y+ )/κ + B for grid No. 3 (x+i = 6) in comparison with data
of Moser et al. [20].

Model MKM-99 BGK MRT Cumulant

Um /U τ 15.73 16.83 20.92 20.59


Uc /U τ 18.30 19.75 23.98 23.74
U c /U m 1.16 1.17 1.15 1.15
κ 0.380 0.385 0.550 0.426
B 5.03 6.70 14.98 11.87

Fig. 18. Instantaneous normalized streamwise velocity ux snapshots in the xy-


symmetry plane predicted by the MRT model for 3 different grids. Image detail for
x+i ≈ 6 refers to pressure contours that display instabilities more clearly.
Table 3 outlines the predicted mean flow results computed on
the third grid featuring x+ i
= 6. Fig. 20 illustrates the related
mean streamwise velocity. It is seen that the BGK results seem to
deteriorate least pronounced when the grid is coarsened. The Cu-
mulant approach delivers an acceptable inclination of the log-layer
and the MRT approach displays major deficiencies. Mind that the
MRT and the BGK simulations both fail when the lattice spacing is
further increased, i.e. xi ≥ 7 (BGK) and xi ≥ 8 (MRT).
With attention directed to the near-wall behavior of the nor-
malized turbulence intensities illustrated in Fig. 21, it is observed
that the BGK model converges much faster to the final result than
the other two collision models for the three finest grids. The inten-
sities decrease for the MRT and the Cumulant model when the grid
is coarsened. Erroneously, the BGK model displays a slightly en-
hanced near-wall intensity when changing from the second finest
Fig. 19. Instantaneous normalized streamwise velocity ux snapshots in the xy- grid (No. 2) to the next coarser grid (No. 3), which reveals first
symmetry plane predicted by the Cumulant model for 3 different grids.
signs for an unphysical behavior. An inspection of the MRT re-
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 259

Fig. 21. Resolution sensitivity of the near-wall normalized turbulence intensities for the BGK (left), MRT (middle) and the Cumulant (right) model computed on the three
finest grids in comparison to MKM-99 [20] data.

Fig. 22. RMS-values v+


RMS
and w+
RMS
(left) predicted by the MRT model on grid No. 4b supplemented by streamwise 1D energy spectra Euxi ui (right) at y+ = 178.

Fig. 23. RMS-values v+


RMS
and w+
RMS
(left) predicted by the BGK model on grid No. 4a supplemented by streamwise 1D energy spectra Euxi ui (right) at y+ = 5.

sults discloses indications of instabilities for grid No. 4b (x+


i
= 8), the vicinity of the outer buffer layer as indicated in Fig. 23. Fig. 24
which are clearly visible from the wall-normal and the lateral in- compares the streamwise 1D energy spectra Euu x for all three colli-
+
tensities in the outer regime and the respective spectra depicted by sion models on grid 4a (BGK, xi = 7) and grid 4b (MRT / CUM,
Fig. 22. The figure exposes, that the instabilities displayed by the x +
i
= 8) for a near-wall (left) and a core flow position (right). The
MRT model occur first in the outer part of the log-layer. In con- Cumulant model uniquely remains stable even for a coarse resolu-
trast to the MRT model, the BGK model shows first instabilities in tion of x+i
= 8.
260 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

x
Fig. 24. Predicted streamwise 1D energy spectra Euu at y+ = 5 (left) and y+ = 178 (right) in comparison to data of Moser et al. [20] (MKM-99).

0.5 MRT
-(u'v') / (u'RMS v'RMS) [-]

MKM-99
0.4 Δx+
i ≈ 2.0
Δx+
i ≈ 3.2 CUMULANT
BGK
0.3 Δx+
i ≈ 6.0 MKM-99
MKM-99 +
Δxi ≈ 8.0 Δx+
i ≈ 2.0
Δx+
i ≈ 2.0
0.2 Δx+
i ≈ 8.6 Δx+
i ≈ 3.2
Δx+
i ≈ 3.2 +
Δxi ≈ 6.0
Δx+
i ≈ 6.0
0.1 + Δx+
i ≈ 8.0
Δxi ≈ 7.0
Δx+
i ≈ 10.0
0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
y / δ [-] y / δ [-] y / δ [-]

Fig. 25. Grid sensitivity of the predicted anisotropy coefficient for the BGK (left), MRT (center) and Cumulant (right) model in comparison to data of Moser et al. [20] (MKM-
99).

Fig. 26. Spanwise (top) and streamwise (bottom) normalized two-point correlations Rz/x z/x +
uu (left) and Rvv (right) at y = 5 obtained for the different resolutions for the BGK
and the Cumulant model in comparison to results of Moser et al. [20] (MKM-99).

The findings are confirmed by Fig. 25, which describes the evo- With attention given to the evolution of the two-point correla-
lution of the anisotropy coefficient −u
v
/(u
RMS v
RMS ) with an in- tions for the initial coarsening from x+i
= 2 to x+i
= 6, the BGK
crease of the lattice spacing. The shear stress is not that signifi- model shows only small changes of the structure indicating min-
cantly affected by initial instabilities and its ratio to the square of ima for the near-wall spanwise correlations Rzuu and Rzvv , whilst the
the friction velocity might seem acceptable, but the anisotropy is other two models exhibit a clear shift of these minima (cf. Fig. 26).
reduced. The figure also affirms that the BGK model starts to fail The expected increase of the near-wall streamwise correlation Rxuu
in the inner part of the boundary layer, whereas the MRT model and Rxvv is observed with an increase of the grid spacing for the Cu-
suffers from a similar phenomena originating from the core flow mulant (and the MRT) model. The results of the BGK model, how-
region. On the contrary, the Cumulant model delivers a fairly sta- ever, display a decrease of these correlations as shown in Fig. 26.
ble level of anisotropy for all grids. Moreover, the BGK result erroneously shows negative Rxvv sections
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 261

coarse resolution. The size of the vortices increases moderately


with the resolution and the shift of Rzvv indicates an increase of
the associated length scale (cf. Fig. 26). The streamwise two-point
correlations – particularly for the primary velocity Rxui ui – returned
by the Cumulant model are, however, significantly enhanced when
the grid spacing is increased. In fact, the values found at the end
of the domain drop down to continuously increasing finite values,
e.g. Rxui ui → 0.15 [0.5] close to the wall or Rxui ui → 0.08 [0.18] in the
core flow for xi = 6 [10]. The phenomena is not seen in the re-
sults of the other two collision models. Fig. 27 displays isosurfaces
of the streamwise vorticity component ωx = ∂ uz /∂ y − ∂ uy /∂ z for
the various resolutions and collision models. Different from the
other two collision models, the vortices displayed by the Cumulant
results significantly stretch in longitudinal direction when the grid
is coarsened. This greatly supports the stability of the simulations
and explains, why the model is deemed to have some implicit sub-
grid modeling features. A typical example is given in Fig. 28 which
reveals that the energy density is progressively redistributed into
larger scales in line with an increased (slightly hyperviscous) de-
cline of small-scale energy when the grid is coarsened.
In conclusion, the Cumulant model also reflects a resolution
dependent damping of small-scale structures similar to a sub-
grid scale model. Sub-grid scale effects are traditionally mod-
eled by the use of a gradient-based isotropic eddy viscosity ν t
in large eddy simulation (LES). The most prominent example is
the Smagorinsky–Lilly functional  eddy viscosity closure [50], i.e.
τ im ∼ 2ν t Sim and νt = (Cs xi )2 2S jk S jk , where Sij refers to the
strain-rate tensor and Cs is the Smagorinsky constant. When at-
tention is directed to LBM, the expression of turbulent stresses has
a link to baseline rationale of the approach. Chen et al. [15] derived
turbulent stresses from a second-order Chapman–Enskog expan-
sion which results in a nonlinear (essentially nonisotropic) eddy
viscosity framework that is supplemented by memory effects. Dif-
ferent from traditional turbulence modeling practices, this non-
linear closure is not controlled by additional coefficients related
to empirical or heuristic approaches, but subjected to the specific
choice of the relaxation time (and the level of turbulent fluctu-
ations). Further LES contributions from empirical closures can, of
course, be added to mimic the influence of unresolved physics that
Fig. 27. Evolution of streamwise vortices for successively coarsened grids. Lateral is misrepresented by the inherent features.
view on isosurfaces of the streamwise vorticity component ωx = ∂ uz /∂ y − ∂ uy /∂ z
As regards the Cumulant model, we suspect, that the inherent
for the BGK (upper subfigure), MRT (center subfigure) and Cumulant (lower sub-
figure) model for x+ = 3.2 (top group), x+ = 6 (center group) and x+ = 7, 8 sub-grid scale treatment features strong anisotropic contributions,
i i i
(bottom group). because of the rather strong anisotropy in the response to a coarser
lattice spacing displayed by the two-point correlations. Moreover,
it is interesting to note that the directionality displayed by the Cu-
mulant model seems to correlate with the velocity and not the
between x+ = 100 and x+ = 400, which marks the onset of insta- velocity gradient, since the streamwise velocity is the only rele-
bilities. vant mean velocity component, but mean streamwise gradients are
The spanwise Rzvv results of the Cumulant model always display zero. This, in turn, suggests a directional mixing length τ ui embed-
clear indications of streamwise vortices even for an inadequate ded in the sub-grid response of the Cumulant model, which might

x
Fig. 28. Resolution dependent evolution of streamwise 1D energy spectra Euu x
(left) and Evv (right) at y+ = 178 returned by the Cumulant model in comparison to data of
Moser et al. [20] (MKM-99).
262 M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263

ing a sub-grid scale model. Although a formal and quantitative link


to an inherent subgrid-scale model is subject to future research
and the detailed response of the Cumulant model might be de-
batable, results of the present study clearly reveal inherent mecha-
nisms to mimic unresolved aspects. A rigorous analysis based upon
higher-order Chapman–Enskog expansion along the line outlined
by Chen et al. [15] for the single-relaxation-time framework might
reveal the turbulent stress tensor of the Cumulant model.
Different from the Cumulant model, the BGK and MRT models
Fig. 29. Comparison of lateral view snapshots on isosurfaces of the streamwise vor-
require stabilization when the lattice distance exceeds twice the
ticity component ωx = ∂ uz /∂ y − ∂ uy /∂ z for the Cumulant-DNS (top) and MRT–LES
(bottom) approach for x+ = 8.
Kolmogorov scale. Instabilities of BGK model origin in the outer
i
buffer layer, whereas the MRT model starts to become unstable in
the core flow.
be part of a memory term that works on higher wave numbers and
would explain the apparent directionality of length scales when Funding
the grid is coarsened. The assumption is confirmed by a compari-
son of the DNS-Cumulant results with results returned by an LES- This research was funded by a Hamburg University of Technol-
stabilized MRT approach on grid No. 4b (x+ = 8). The LES sta- ogy (TUHH) internal grant. The research did not receive any spe-
i
bilization refers to a standard Smagorinsky sub-grid scale model. cific grant from funding agencies, neither in the public, commercial
Further details of the MRT–LES are beyond the scope of the present or non-profit sectors.
paper. Fig. 29 compares a side view of the axial vortices for these
two simulations. An additional comparison of the spanwise two- Appendix A. Details of the employed D3Q19 MRT model
point correlations reveals that the streamwise vortices of the LES
approach are somewhat larger. Their streamwise length, however, The transformation matrix [18] from particle distribution space
to moment space reads
⎡ ⎤
1 · (1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1)
⎢ c2 · (−1 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 1 1 1) ⎥
⎢ c4 · (1 −2 −2 −2 −2 −2 −2 1 1 1 1 1 1)

⎢ 1 1 1 1 1 1 ⎥
⎢ c· (0 1 −1 0 0 0 0 1 −1 1 −1 1 −1 1 −1 0 0 0 0 ) ⎥
⎢ ⎥
⎢ c3 · (0 −2 2 0 0 0 0 1 −1 1 −1 1 −1 1 −1 0 0 0 0 ) ⎥
⎢ ⎥
⎢ c· (0 0 0 1 −1 0 0 1 −1 −1 1 0 0 0 0 1 −1 1 −1 ) ⎥
⎢ c3 · (0 0 0 −2 −1 −1 1 0 0 0 1 −1 1 −1 ) ⎥
⎢ 2 0 0 1 0 ⎥
⎢ c· (0 0 0 0 0 1 −1 0 0 0 0 1 −1 −1 1 1 −1 −1 1 ) ⎥
⎢ ⎥
⎢ c3 · (0 0 0 0 0 −2 2 0 0 0 0 1 −1 −1 1 1 −1 −1 1 ) ⎥
⎢ c2 · (0 2 2 −1 1 −2 −2 −2 −2 )

Mαβ =⎢ −1 −1 −1 1 1 1 1 1 1 1 ⎥ (A.1)
⎢ c4 · (0 −2 −2 1 1 −2 −2 −2 −2 ) ⎥
⎢ 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ c2 · (0 0 0 1 1 −1 −1 1 1 1 1 −1 −1 −1 −1 0 0 0 0 ) ⎥
⎢ ⎥
⎢ c4 · (0 0 0 −1 −1 1 1 1 1 1 1 −1 −1 −1 −1 0 0 0 0 ) ⎥
⎢ c2 · (0 0 0 0 0 0 0 0 0)

⎢ 0 0 0 1 1 −1 −1 0 0 0 ⎥
⎢ c2 · (0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 −1 −1 ) ⎥
⎢ ⎥
⎢ c2 · (0 0 0 0 0 0 0 0 0 0 0 1 1 −1 −1 0 0 0 0 ) ⎥
⎢ ⎥
⎢ c3 · (0 0 0 0 0 0 0 1 −1 1 −1 −1 1 −1 1 0 0 0 0) ⎥
⎣ c3 · (0 0 0 0 0 0 0 −1 1 1 −1 0 0 0 0 1 −1 1 −1 ) ⎦
c3 · (0 0 0 0 0 0 0 0 0 0 0 1 −1 −1 1 −1 1 1 −1 )

The resulting moments are termed


seems fairly similar, perhaps slightly shorter [longer] in the near-
mα = (ρ , e,  , jx , qx , jy , qy , jz , qz , 3 pxx , 3πxx ,
wall [core flow] region for the MRT–LES.
pww , πww , pxy , pyz , pxz , mx , my , mz )T . (A.2)
6. Conclusions The equilibrium moments depending on macroscopic flow param-
eters are defined as
A fair prediction of incompressible turbulence physics with LBM  
requires Mach numbers Ma ≤ 0.1 and a spatial resolution in the meq
1
= ρ , meq
2
= eeq = ρ0 u2x + u2y + u2z ,
range of the Kolmogorov scale. An adequate domain size is also im- meq =ρ0 ux , meq = ρ0 uy , meq = ρ0 u z ,
portant to capture the details, as indicated by previous researchers.
4

6

8
 2 
Under such circumstances, the D3Q19 LBM-BGK model offers a sat-
m10 = 3 pxx = ρ0 2ux − uy − uz , meq
eq eq 2 2 2
12
= peqzz = ρ0 uy − uz ,
2

meq xy = ρ0 ux uy , m15 = pyz = ρ0 uy uz , m16 = pxz = ρ0 ux uz ,


= peq eq eq eq eq
isfactory predictive accuracy, which seems slightly superior to the 14
more advanced MRT and Cumulant approaches. All relevant physi-
meq
3
= meq
5
= meq
7
= meq
9
= meq
11
= meq
13
= meq
17
= meq
18
= meq
19
= 0. (A.3)
cal details are, however, captured by any of the investigated colli-
sion models under ideal conditions. The diagonal matrix containing the collision rates reads
The Cumulant approach uniquely returns streamwise correla- Sαβ = diag(0, sa , sb , 0, sc , 0, sc , 0, sc , sω , sd , sω , sd , sω , sω ,
tion lengths that are rigidly aligned to the lattice spacing. While
sω , se , se , se ), (A.4)
this delays the convergence to formerly reported DNS Navier–
Stokes data in the limit of fine grids, it also provides stable mean- wherein sω = t/τ = ω is linked to shear viscosity and sa , sb , sc ,
ingful results when the resolution is insufficient without consider- sd , se ∈ ]0,2[, each to be chosen freely.
M. Gehrke et al. / Computers and Fluids 156 (2017) 247–263 263

References [26] Buick JM, Greated CA. Gravity in a lattice Boltzmann model. Phys Rev E
20 0 0;61:5307–20.
[1] Aidun CK, Clausen JR. Lattice-Boltzmann method for complex flows. Annu Rev [27] Dellar PJ. Nonhydrodynamic modes and a priori construction of shallow water
Fluid Mech 2010;42:439–72. lattice Boltzmann equations. Phys Rev E 2002;65(3):036309.
[2] Feichtinger C, Habich J, Köstlera H, Hager G, Rüde U, Wellein G. A flexi- [28] Asinari P. Asymptotic analysis of multiple-relaxation-time lattice Boltzmann
ble patch-Based lattice Boltzmann parallelization approach for heterogeneous schemes for mixture modeling. Comput Math Appl 2008;55(7):1392–407.
GPU-CPU clusters. Parallel Comput 2011;37:535–49. [29] Stiebler M, Krafczyk M, Freudiger S, Geier M. Lattice Boltzmann large eddy
[3] Tölke J, Krafczyk M. Teraflop computing on a desktop PC with GPUs for 3D simulation of subcritical flows around a sphere on non-uniform grids. Comput
CFD. Int J Comput Fluid Dyn 2008;22:443–56. Math Appl 2011;61(12):3475–84.
[4] Wang X, Aoki T. High performance computation by multi-node GPU cluster-T- [30] Janßen CF, Krafczyk M. Free surface flow simulations on GPGPUs using the
SUBAME 2.0 on the air flow in an urban city using lattice Boltzmann method. LBM. Comput Math Appl 2011;61(12):3549–63.
Int J Aerosp Lightweight Struct 2012;2:77–86. [31] Janßen CF, Mierke D, Überrück M, Gralher S, Rung T. Validation of the GPU-Ac-
[5] Yu H, Girimaji SS, Luo L-S. DNS And LES of decaying isotropic turbulence with celerated CFD solver ELBE for free surface flow problems in civil and environ-
and without frame rotation using lattice boltzmann method. J Comput Phys mental engineering. Computation 2015;3:354–85.
2005;209:599–616. [32] Mierke D, Janßen CF, Rung T. GPU-Accelerated Large-Eddy Simulation of
[6] Yu H, Luo L-S, Girimaji SS. LES Of turbulent square jet flow using an MRT lat- Ship-Ice Interactions. In: Proceedings of the VI international conference
tice Boltzmann model. Comput Fluids 2006;35:957–65. on computational methods in marine engineering (MARINE 2015); 2015.
[7] Girimaji SS. Boltzmann kinetic equation for filtered fluid turbulence. Phys Rev p. 850–62.
Lett 2007;99. 034501/1–034501/4. [33] Janßen C.F., Mierke D., Rung T. On the development of an efficient numerical
[8] Freitas RK, Henze A, Meinke M, Schröder W. Analysis of lattice-Boltzmann ice tank for the simulation of fluid-ship-rigid-ice interactions on graphics pro-
methods for internal flows. Comput Fluids 2011;47:115–21. cessing units. Comput. Fluids. doi:10.1016/j.compfluid.2017.05.006, currently in
[9] Wang X, Shangguan Y, Onodera N, Kobayashi H, Aoki T. Direct numer- print.
ical simulation and large eddy simulation on a turbulent wall-Bounded [34] Krafczyk M, Kucher K, Wang Y, Geier M. DNS/LES Studies of turbulent flows
flow using lattice Boltzmann method and multiple GPUs. Math Prob Eng based on the cumulant lattice boltzmann approach. In: High performance
2014;2014(742432):1–10. computing in science and engineering 14. Springer; 2015. p. 519–31.
[10] Spasov M, Rempfer D, Mokhasi P. Simulation of a turbulent channel flow [35] Pasquali A, Schönherr M, Geier M, Krafczyk M. Simulation of external aerody-
with an entropic lattice boltzmann method. Int J Numer Meth Fluids namics of the drivaer model with the LBM on GPGPUs, parallel computing: on
2009;60:1241–58. the road to exascale. Adv Parallel Comput 2015;27:391–400.
[11] Kang SK, Hassan YA. The effect of lattice models within the lattice Boltzmann [36] Pasquali A. Enabling the cumulant lattice Boltzmann method for complex CFD
method in the simulation of wall-bounded turbulent flows. J Comput Phys engineering problems. Ph.D. thesis. Technical University Braunschweig; 2016.
2013;232:100–17. [37] Far EK, Geier M, Kutscher K, Krafczyk M. Simulation of micro aggregate break-
[12] Lammers P, Beronov KN, Volkert R, Brenner G, Durst F. Lattice BGK direct age in turbulent flows by the cumulant lattice Boltzmann method. Comput
numerical simulation of fully developed turbulence in incompressible plane Fluids 2016;140:222–31.
channel flow. Comput Fluids 2006;35:1137–53. [38] Xu H, Sagaut P. Optimal low-dispersion low-dissipation LBM schemes for com-
[13] Bespalko D, Pollard A, Uddin M. Analysis of pressure fluctuations of an LBM putational aeroacoustics. J Comput Phys 2011;230(13):5353–82.
simulation of turbulent channel flow. Comput Fluids 2012;54:143–6. [39] Dubois F, Lallemand P. Towards higher order lattice Boltzmann schemes. J Stat
[14] Orszag SA, Chen H, Succi S, Latt J, Chopard B. Turbulence effects on kinetic Mech: Theory Exp 20 09;20 09(06):P06006.
equations. J Sci Comput 2004;28:459–66. [40] Dubois F, Lallemand P, Tekitek M. On a superconvergent lattice boltzmann
[15] Chen H, Orszag SA, Staroselsky I, Succi S. Expanded analogy between Boltz- boundary scheme. Comput Math Appl 2010;59(7):2141–9.
mann kinetic theory of fluids and turbulence. J Fluid Mech 2006;519:301–14. [41] d’Humières D, Ginzburg I. Viscosity independent numerical errors for lattice
[16] Bhatnagar P, Gross E, Krook M. A model for collision processes in gases. I. Boltzmann models: from recurrence equations to magic collision numbers.
Small amplitude processes in charged and neutral one-component systems. Comput Math Appl 2009;58(5):823–40.
Phys Rev 1954;94:511–25. [42] Ginzburg I. Variably saturated flow described with the anisotropic lattice
[17] d’Humières D, Ginzburg I, Krafczyk M, Lallemand P, Luo L-S. Multiple-relax- Boltzmann methods. Comput Fluids 2006;35(8):831–48.
ation-time lattice Boltzmann models in three dimensions. R Soc Lond Philos [43] Ginzburg I, Verhaeghe F, d’Humières D. Two-relaxation-time lattice Boltzmann
Trans Ser-A 2002;360:437–51. scheme: about parametrization, velocity, pressure and mixed boundary condi-
[18] Tölke J, Freudiger S, Krafczyk M. An adaptive scheme using hierarchical tions. Commun Comput Phys 2008a;3(2):427–78.
grids for lattice Boltzmann multi-phase flow simulations. Comput Fluids [44] Ginzburg I, Verhaeghe F, d’Humières D. Study of simple hydrodynamic solu-
2006;35(8):820–30. tions with the two-relaxation-times lattice Boltzmann scheme. Commun Com-
[19] Geier M, Schönherr M, Pasquali A, Krafczyk M. The cumulant lattice Boltz- put Phys 2008b;3(3):519–81.
mann equation in three dimensions: theory and validation. Comput Math Appl [45] Ginzburg I. Prediction of the moments in advection-diffusion lattice Boltz-
2015;70:507–47. mann method. I. Truncation dispersion, skewness, and kurtosis. Phys Rev E
[20] Moser RD, Kim J, Mansour NN. Direct numerical simulation of turbulent chan- 2017;95(1):013304.
nel flow up to reτ =590. Phys Fluids 1999;11:943–5. [46] Ginzburg I, d’Humières D. Multireflection boundary conditions for lattice
[21] Kim J, Moin P, Moser RD. Turbulence statistics in fully developed channel flow boltzmann models. Phys Rev E 2003;68(6):066614.
at low Reynolds number. J Fluid Mech 1987;177:133–66. [47] Zanoun E-S, Durst F, Nagib H. Evaluating the law of the wall in two-dimen-
[22] Lammers P. Direkte numerische Simulationen wandgebundener Strömungen sional fully developed turbulent channel flows. Phys Fluids 2003;15:3079–89.
kleiner Reynoldszahlen mit dem lattice Boltzmann Verfahren. Ph.D. thesis. [48] Moser RD, Moin P. Direct numerical simulation of curved turbulent channel
Technische Fakultät der Universität Nürnberg-Erlangen; 2004. flow. Technical Report TF-20 (also NASA TM 85974). Department of Mechanical
[23] Jiménez J, Moin P. The minimal flow unit in near wall turbulence. J Fluid Mech Engineering, Stanford University, Stanford, California, USA.; 1984.
1991;225:213–40. [49] Tritton DJ. Some new correlation measurements in a turbulent boundary layer.
[24] Geier M, Greiner A, Korvink JG. Cascaded digital lattice Boltzmann automata J Fluid Mech 1967;28:133–66.
for high Reynolds number flow. Phys Rev E 2006;73:066705. [50] Smagorinsky J. General circulation experiments with the primitive equations.
[25] Guo Z, Zheng C, Shi B. Discrete lattice effects on the forcing term in the lattice Mon Weather Rev 1963;91:99–164.
Boltzmann method. Phys Rev E 2002;65(4):046308.

You might also like