ArticlePDF Available

Abstract and Figures

Conventional lattice Boltzmann models for the simulation of fluid dynamics are restricted by an error in the stress tensor that is negligible only for small flow velocity and at a singular value of the temperature. To that end, we propose a unified formulation that restores Galilean invariance and the isotropy of the stress tensor by introducing an extended equilibrium. This modification extends lattice Boltzmann models to simulations with higher values of the flow velocity and can be used at temperatures that are higher than the lattice reference temperature, which enhances computational efficiency by decreasing the number of required time steps. Furthermore, the extended model also remains valid for stretched lattices, which are useful when flow gradients are predominant in one direction. The model is validated by simulations of two- and three-dimensional benchmark problems, including the double shear layer flow, the decay of homogeneous isotropic turbulence, the laminar boundary layer over a flat plate and the turbulent channel flow.
Content may be subject to copyright.
entropy
Article
Extended Lattice Boltzmann Model
Mohammad Hossein Saadat, Benedikt Dorschner and Ilya Karlin *


Citation: Saadat, M.H.; Dorschner,
B.; Karlin, I.V. Extended Lattice
Boltzmann Model. Entropy 2021,23,
475. https://doi.org/10.3390/
e23040475
Academic Editor: Carlos
Mejía-Monasterio
Received: 21 February 2021
Accepted: 13 April 2021
Published: 17 April 2021
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2021 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
Department of Mechanical and Process Engineering, ETH Zurich, 8092 Zurich, Switzerland;
msaadat@ethz.ch (M.H.S.); bdorschn@ethz.ch (B.D.)
*Correspondence: ikarlin@ethz.ch
Abstract:
Conventional lattice Boltzmann models for the simulation of fluid dynamics are restricted
by an error in the stress tensor that is negligible only for small flow velocity and at a singular value of
the temperature. To that end, we propose a unified formulation that restores Galilean invariance and
the isotropy of the stress tensor by introducing an extended equilibrium. This modification extends
lattice Boltzmann models to simulations with higher values of the flow velocity and can be used at
temperatures that are higher than the lattice reference temperature, which enhances computational
efficiency by decreasing the number of required time steps. Furthermore, the extended model also
remains valid for stretched lattices, which are useful when flow gradients are predominant in one
direction. The model is validated by simulations of two- and three-dimensional benchmark problems,
including the double shear layer flow, the decay of homogeneous isotropic turbulence, the laminar
boundary layer over a flat plate and the turbulent channel flow.
Keywords: lattice Boltzmann method; Galilean invariance; extended equilibrium
1. Introduction
The lattice Boltzmann method (LBM) solves a Boltzmann-type kinetic equation on
a discrete velocity set, forming the links of a space-filling lattice. Efficiency of the LBM
makes it attractive for the simulation of a wide range of problems in fluid dynamics, see,
for example, [1,2].
In this paper, we revisit the restrictions of LBM due to the geometry of the discrete
velocities. It is well known that standard LBM velocities yield a persistent error in the
fluid stress tensor, which breaks Galilean invariance and limits the operation range of
LBM to small flow velocities and a singular value of the lattice reference temperature; only
under these conditions can the error be ignored. While one can cope with this error in
most incompressible flow applications [
1
,
2
], it certainly affects high-speed compressible
flows [312]
and sometimes even low-speed isothermal cases [
13
]. Moreover, the same
error is amplified when using stretched (rectangular) lattices instead of the conventional
(cubic) lattice, where in addition to the corrupted Galilean invariance, the stress tensor
becomes anisotropic [14,15].
The extension of LBM beyond its classical operation domain has so far been addressed
with different techniques, depending on the desired outcome. Introducing lattices with
more velocities (multi-speed lattices) [
16
] is one technique, which comes with a significant
increase of computational cost and has been used mostly for high-speed compressible
flow applications [
16
]. In the standard LB setting, on the other hand, one approach is to
alter the relaxation rates and use a multi-relaxation time collision operator [
9
]. Another
approach to extend the flow velocity and temperature range of the standard cubic lattices
is to add correction terms to the original LBM [
4
8
,
10
12
,
17
21
]. The realization varies
among different authors but none address the general case of rectangular lattices. On the
other hand, rectangular lattices may improve the computational efficiency of the LBM
by using a coarser mesh in the direction of smaller gradients in the flow. Unlike other
approaches of handling non-uniform grids (e.g., Eulerian [
22
,
23
] and semi-Lagrangian off-
lattice
LBM [2427]
or grid refinement techniques [
28
,
29
]), stretched lattices do not require
Entropy 2021,23, 475. https://doi.org/10.3390/e23040475 https://www.mdpi.com/journal/entropy
Entropy 2021,23, 475 2 of 23
a substantial change in the standard LBM algorithm. Recent work on the stretched LBM
restores the isotropy of the stress tensor by using multi-relaxation time LBM models [
30
32
].
However, these approaches do not address the flow velocity and
temperature restrictions
.
In this paper, we propose a unified view on the three aspects of the problem, the
velocity range, the temperature range and grid stretching, which all stem from the same
error, induced by constraints of the discrete velocity set. In particular, we propose to use
an extended equilibrium, which restores the Galilean invariance and isotropy of the stress
tensor, enabling simulations at higher flow velocities, higher temperatures using both cubic
and stretched lattices, yielding increased accuracy and efficiency.
The paper is organized as follows—in Section 2, we start by presenting the discrete
kinetic equations, following the standard single-relaxation time lattice Bhatnagar–Gross–
Krook (LBGK) setting, as well as the equilibrium and extended equilibrium formulation,
followed by the derivation of the model’s hydrodynamic limit. In addition, the locally
corrected LBM of Reference [9] is compared to the extended LBGK in Appendix A.
Subsequently, in Section 3, we assess the validity, accuracy and performance of our
model using both two- and three-dimensional benchmark problems. As a first step, we
verify Galilean invariance, temperature independence and isotropy of the model on the
example of an advected decaying shear wave. It is shown that the theoretical viscosity
is recovered for both cubic as well as stretched lattices in a large range of temperatures
and advection velocities. This also indicates that the model can readily be extended to
high-speed compressible flows, provided that it is augmented with a suitable solver for the
total energy. Next, for the example of homogeneous isotropic turbulence, we demonstrate
that a speed-up can be achieved by using an operating temperature, which is larger than
the lattice reference temperature. The present model can also be viewed as an alternative to
preconditioned LBM [33] for accelerating the convergence rate but without the restriction
to steady flows. Finally, accuracy and performance are assessed for rectangular lattices
using the doubly periodic shear flow, laminar flow over a flat plate and turbulent channel
flow as examples. Conclusions are drawn in Section 4.
2. Discrete Kinetic Equations
2.1. LBGK
We consider the LBGK equation for the populations
fi
, associated to the discrete
velocities vifor i=0, . . . , Q1,
fi(x+viδt,t+δt)fi(x,t) = ω(fex
ifi), (1)
where
x
denotes the location in space and
δt
is the time step. The extended equilibrium
fex
i
, which will be specified below, satisfies the local conservation laws for the density
ρ
and momentum ρu,
ρ=
Q1
i=0
fex
i=
Q1
i=0
fi, (2)
ρu=
Q1
i=0
vifex
i=
Q1
i=0
vifi. (3)
As we will show below, the relaxation parameter ωis related to the kinematic viscosity ν,
ν=1
ω1
2RTδt, (4)
where
T
is the temperature and
R
is the gas constant. We now proceed with identifying the
extended equilibrium.
Entropy 2021,23, 475 3 of 23
2.2. Discrete Velocities and Factorization
We use the
D
3
Q
27 lattice, where
D=
3 denotes the spatial dimension and
Q=
27 is
the number of discrete speeds, which are given by
ci= (cix ,ciy,ciz),ciα∈ {1, 0, 1}. (5)
With (5), we define the particles’ velocities viin a stretched cell as
vi= (λxcix ,λyciy,λzciz), (6)
where
λα
is the stretching factor in the direction
α
. To maintain on-lattice propagation, the
cell size is changed accordingly to δxα=λαδt.
The (normalized, M000 =1) moments Mlmn are defined using the convention
lx,my,nz;l,m,n=0, 1, 2, . . . , (7)
and thus
ρMlmn =
Q1
i=0
vl
ix vm
iy vn
iz fi. (8)
For convenience, we use a more specific notation for the first-order and the diagonal
second-order moments,
M100 =ux,M010 =uy,M001 =uz, (9)
M200 =Pxx ,M020 =Pyy,M002 =Pzz. (10)
We essentially follow [
34
] and consider a class of factorized populations. To that end, we
define a triplet of functions in the three variables, uα,Pαα and λα,
Ψ0(uα,Pαα,λα) = 1Pαα
λ2
α
, (11)
Ψ1(uα,Pαα,λα) = 1
2uα
λα
+Pαα
λ2
α, (12)
Ψ1(uα,Pαα,λα) = 1
2uα
λα
+Pαα
λ2
α. (13)
For the vectors u,P, and λ,
u= (ux,uy,uz), (14)
P= (Pxx ,Pyy,Pzz), (15)
λ= (λx,λy,λz), (16)
we consider a product-form, associated with the discrete velocities vi(6),
Ψi(u,P,λ) =
α=x,y,z
Ψciα(uα,Pαα,λα). (17)
The normalized moments of the product-form (17),
Mlmn =
Q1
i=0
vl
ix vm
iy vn
iz Ψi, (18)
are readily computed thanks to the factorization,
Mlmn =Ml00 M0m0M00n, (19)
Entropy 2021,23, 475 4 of 23
where
M000 =1, (20)
Ml00 =(λl1
xux,lodd
λl2
xPxx ,leven, (21)
M0m0=(λm1
yuy,modd
λm2
yPyy,meven, (22)
M00n=(λn1
zuz,nodd
λn2
zPzz,neven. (23)
For any stretching (16), the six-parametric family of normalized populations (17) is identi-
fied by the flow velocity (14) and the diagonal of the pressure tensor at unit
density (15)
,
and was termed the unidirectional quasi-equilibrium in Ref. [
34
]. We make use of the
product-form (17) to construct all pertinent populations, the equilibrium and the extended
equilibrium.
2.3. Equilibrium and Extended Equilibrium
The equilibrium
feq
i
is defined by setting
Pαα
(10) equal to the equilibrium diagonal
element of the pressure tensor at unit density,
Peq
αα =RT +u2
α. (24)
Substituting (24) into (17), we get
feq
i=ρΨi(u,Peq,λ). (25)
With (18), we find the pressure tensor and the third-order moment tensor at the equilibrium
(25) as follows,
Peq =
Q1
i=0
vivifeq
i=PMB, (26)
Qeq =
Q1
i=0
vivivifeq
i=QMB +˜
Q. (27)
The isotropic parts,
PMB
and
QMB
, are the Maxwell–Boltzmann (MB) pressure tensor and
the third-order moment tensor, respectively,
PMB =pI+ρuu, (28)
QMB =sym(pIu) + ρuuu, (29)
where
p=ρRT
is the pressure,
sym(. . . )
denotes symmetrization and
I
is the unit tensor.
The anisotropy of the equilibrium (25) manifests with the deviation
˜
Q=Qeq QMB
in
(29), where only the diagonal elements are non-vanishing,
˜
Qαβγ =(ρuα(λ2
α3RT)ρu3
α, if α=β=γ,
0, otherwise. (30)
The origin of the diagonal anomaly (30) is the geometric constraint,
v3
iα=λ2
αviα
, which is
imposed by the choice of the discrete speeds (5), and is well known in the case of the
standard (unstretched) lattice with
λα=
1. A remedy in the latter case is to minimize
spurious effects of anisotropy by fixing the temperature T=TL,
Entropy 2021,23, 475 5 of 23
RTL=1
3, (31)
In order to eliminate the linear term
uα
in (30). Thus, using the equilibrium (25) in the
LBGK Equation (1) imposes a two-fold restriction on the operation domain: the temper-
ature cannot be chosen differently from (31) and the flow velocity has to be maintained
asymptotically vanishing. Moreover, for stretched lattices, the anisotropy becomes even
more pronounced since it is not possible to eliminate the linear deviation in all directions
simultaneously by fixing any temperature.
Alternatively, the spurious anisotropy effects can be canceled out by extending the
equilibrium such that the third-order moment anomaly is compensated in the hydrody-
namic limit. Because the anomaly only concerns the diagonal (unidirectional) elements
of the third-order moments, the cancellation can be achieved by redefining the diagonal
elements of the second-order moments. As demonstrated below, in order to cancel the
errors, the diagonal elements Pex
αα for the extended equilibrium must be chosen as
Pex
αα =Peq
αα +δt2ω
2ρω α˜
Qααα, (32)
where spatial derivative is evaluated using a second-order central difference scheme.
Hence, the extended equilibrium fex
iis specified by using the product-form (17),
fex
i=ρΨi(u,Pex,λ). (33)
We shall now proceed with the derivation of the Navier–Stokes equations in the hydrody-
namic limit of the proposed extended LBGK model.
2.4. Hydrodynamic Limit with Extended Equilibrium
Taylor expansion of the shift operator in (1) to second order gives,
δtDi+δt2
2DiDifi=ω(fex
ifi), (34)
where Diis the derivative along the characteristics,
Di=t+vi·. (35)
Introducing the multi-scale expansion,
fi=f(0)
i+δt f (1)
i+δt2f(2)
i+O(δt3), (36)
fex
i=fex(0)
i+δt f ex(1)
i+δt2fex(2)
i+O(δt3), (37)
t=(1)
t+δt(2)
t+O(δt2), (38)
substituting into (34) and using the notation,
D(1)
i=(1)
t+vi·, (39)
we obtain, from the zeroth to second order in the time step δt,
f(0)
i=fex(0)
i=feq
i, (40)
D(1)
if(0)
i=ωf(1)
i+ωfex(1)
i, (41)
(2)
tf(0)
i+vi· f(1)
iω
2D(1)
if(1)
i+ω
2D(1)
ifex(1)
i=ωf(2)
i+ωfex(2)
i. (42)
Entropy 2021,23, 475 6 of 23
With (40), the mass and the momentum conservation (2) and (3) imply the solvability
conditions,
Q1
i=0
fex(k)
i=
Q1
i=0
f(k)
i=0, k=1, 2 . . . ; (43)
Q1
i=0
vifex(k)
i=
Q1
i=0
vif(k)
i=0, k=1, 2, . . . . (44)
With the equilibrium (25), taking into account the solvability conditions (43) and
(44)
, and also making use of the equilibrium pressure tensor (26) and (28), the first-order
Equation (41) implies the following relations for the density and the momentum,
(1)
tρ=−∇ · (ρu), (45)
(1)
t(ρu) = −∇ · (pI+ρuu). (46)
Moreover, the first-order constitutive relation for the nonequilibrium pressure tensor
P(1)
is found from (41) as follows,
ωP(1)+ωPex(1)=(1)
tPeq +∇ · Qeq, (47)
where
P(1)=
Q1
i=0
vivif(1)
i, (48)
Pex(1)=
Q1
i=0
vivifex(1)
i. (49)
With the help of Equations (28) and (29), the first-order constitutive relation (47) is trans-
formed to make explicit the contribution of the anomalous term (30),
ωP(1)+ωPex(1)= · ˜
Q+(1)
tPMB +∇ · QMB. (50)
The last term is evaluated using (45) and (46) to give,
(1)
tPMB +∇ · QMB =ρRTu+u, (51)
where
(·)
denotes transposition. Combining (51) and (50), the first-order constitutive
relation becomes,
ωP(1)=· ˜
QωPex(1)+ρRTu+u. (52)
Note that, if we had used the equilibrium
feq
i
instead of the extended equilibrium
fex
i
in
(1), at this stage of the derivation, we get instead of (52),
ωP(1)=· ˜
Q+ρRTu+u. (53)
The anomalous term
· ˜
Q
cannot be canceled in the latter expression, rather, by choosing
T=TL
(31), its effect can be ignored but only under the assumption of an asymptotically
small flow velocity. Moreover, for a quasi-incompressible flow (
Ma
0, density variation
ρMa2
, where
Ma
is a characteristic Mach number), it is possible to further reduce
the effect of the anomaly by rescaling the relaxation parameter [
9
], see a discussion in
Appendix A.
Entropy 2021,23, 475 7 of 23
In a contrast, using the present formulation, the cancellation is possible by finding
the corresponding expression for the correction term Pex(1), to which end we need to con-
sider the second-order contribution to the momentum equation. Applying the solvability
condition (43) and (44) to the second-order Equation (42), we obtain,
(2)
tρ=0, (54)
(2)
t(ρu) = −∇ · h1ω
2P(1)+ω
2Pex(1)i. (55)
The latter is transformed by virtue of (52),
(2)
t(ρu) = −∇·1
ω1
2ρRT(u+u)+ · 1
ω1
2· ˜
QPex(1). (56)
The last (anomalous) term is canceled out by choosing,
Pex(1)=2ω
2ω· ˜
Q. (57)
Combining the result (57) with the zeroth-order (equilibrium) value, we arrive at the
extended pressure tensor
Pex =Peq +δtPex(1)
=pI+ρuu+δt2ω
2ω· ˜
Q. (58)
Since the anomalous contribution is a diagonal tensor, cf. Equation (30), the result (58) is im-
plemented with the extended equilibrium in the product-form by choosing the normalized
(at unit density) diagonal elements of the pressure tensor as follows,
Pex
αα =RT +u2
α+δt2ω
2ρω αρuαλ2
α3RT u2
α, (59)
which is equivalent to (32). Finally, combining the first- and second-order contributions
to the density and the momentum equation, (45), (46), (54) and (56), using a notation,
t=(1)
t+δt(2)
t
, and also taking into account the cancellation of the anomalous term in
(56), we arrive at the continuity and the flow equations as follows,
tρ+·(ρu) = 0, (60)
tu+u·u+1
ρp+1
ρ·Π=0, (61)
where pis the pressure of ideal gas at constant temperature T,
p=ρRT, (62)
Πis the viscous pressure tensor,
Π=µS, (63)
with Sthe rate of strain,
S=u+u, (64)
and µthe dynamic viscosity,
Entropy 2021,23, 475 8 of 23
µ=1
ω1
2pδt. (65)
The above considerations can be summarized as follows—because of the third-order
moment anomaly (30), the LBGK Equation (1) with the product-form equilibrium (25) is
restricted in several ways, namely:
(i)
The temperature is restricted to a single value, the lattice reference temperature
TL
(31);
(ii)
The flow velocity has to be asymptotically vanishing;
(iii)
Stretched velocities amplify these restrictions by making it impossible to cancel even
the linear (in velocity) anomaly in all the directions simultaneously.
Note that, in addition to all of the restrictions above, when using the conventional
second-order equilibrium obtained by retaining the terms up to the order of
uαuβ
in (25),
the anomaly becomes not only confined to the diagonal elements
Qeq
ααα
but also contaminates
the off-diagonal elements
Qeq
αββ
. While the diagonal anomaly (30) is genuine, that is, it is
caused by the geometry of the discrete velocities, this additional off-diagonal deviation is due
to an unsolicited second-order truncation of the product-form equilibrium (25).
The proposed revision of the LBGK model is based on extending the product-form
equilibrium such that the anomaly of the diagonal third-order moment is compensated
in the hydrodynamic limit by counter terms, which are added to the diagonal of the
equilibrium pressure tensor. With this, all three restrictions mentioned above are addressed
at once, without making a special distinction between the temperature, flow velocity or
stretching as separate causes for the anisotropy.
3. Numerical Results
In this section, we shall access the accuracy and performance of the proposed LB
model in a variety of scenarios of activating spurious anisotropy. First, we test Galilean
invariance, isotropy and temperature independence of the model with both regular and
rectangular lattices in the simulation of a decaying shear wave. Second, we validate the
model for the more complex case of decaying homogeneous isotropic turbulence and
show the effectiveness of using higher temperatures in saving compute time. Third, we
investigate the applicability of the proposed model with stretched lattices in a periodic
double shear layer flow, in a laminar flow over a flat plate, and finally in the case of the
turbulent channel flow. In the simulations below, the gas constant was set to
R=
1, the
time step is
δt=
1 and Grad’s approximation, as proposed in [
35
], was used for wall
boundary conditions.
3.1. Galilean Invariance, Isotropy and Temperature Independence Test
To probe the Galilean invariance and temperature independence of the model, the
kinematic viscosity
ν=µ/ρ
(4) is measured for the decay of a plane shear wave. First, we
consider the axis-aligned setup, with the initial condition,
ρ=ρ0,ux=a0sin(2πy/Ly),uy=MaT, (66)
where
Ma =u0/T
is the advection Mach number,
a0=
0.001 is the amplitude,
Ly=
200
is number of grid nodes in the
y
direction,
ρ0=
1. The nominal kinematic viscosity is
set to
ν=
0.01. Periodic boundary conditions are imposed in both
x
- and
y
- directions.
The numerical viscosity (
νnum
) is measured by fitting an exponential to the time decay of
maximum flow velocity
ux,max eνt(2π/Ly)2
. In this special case, the diagonal anomaly
(30)
is dormant and does not trigger any spurious effects because the derivatives
x˜
Qxxx
and
y˜
Qyyy
both vanish. Consequently, the extended equilibrium (33) becomes equivalent
to the product-form equilibrium (25) in this case.
In order to compare with the standard LBGK, the standard velocities
λα=
1 were
used in this simulation. Figures 1and 2show the importance of using the product-form
equilibrium (25) as opposed to the conventional LBGK model with the second-order equi-
Entropy 2021,23, 475 9 of 23
librium. A strong dependence of the viscosity on the reference frame for the second-order
equilibrium can be seen in Figure 1, where the viscosity drops with increasing advection
Mach number. This well-known artifact of the second-order equilibrium is due to the non-
vanishing anomaly in the off-diagonal moments
Qeq
αββ
and, unlike the diagonal anomaly, is
caused only by the approximate treatment of the product-form equilibrium. Moreover, as
shown in Figure 2, even at a small enough velocity this spurious feature improves only
at the lattice reference temperature
TL
. In contrast, as is shown in
Figures 1and 2
, the
product-form equilibrium of the present model is able to accurately predict the viscosity in
this setup for a wide range of temperatures and reference frame velocities.
Ma
νnum / ν
0.1 0.2 0.3 0.4 0.5 0.6
0.65
0.7
0.75
0.8
0.85
0.9
0.95
1
Second-order equilibrium
Product-form equilibrium
T = 1/3
Figure 1.
Numerical measurement of viscosity for axis-aligned setup at temperature
T=
1
/
3 for
different velocities. The exact solution corresponds to νnum/ν=1.
Next, in order to trigger the anisotropy of the deviation terms (30) and to show
the necessity of using the extended equilibrium, the shear wave is rotated by
π/
4. The
anisotropy is further increased by also conducting simulations on a stretched grid with
λx=
2. The temperature is kept at
T=
1
/
3. The viscosity measurement is shown in
Figure 3for different advection Mach numbers and stretching factors. It can be observed
that the model lacks Galilean invariance for larger velocities when using the product-form
equilibrium without correction (25). Furthermore, the stretching factor
λx=
2 results in a
significant hyper-viscosity since the deviation (30) in this case amounts to a large positive
number. However, once the correction term is included and the extended equilibrium
(33)
is used, the present model recovers the imposed viscosity, independent of the frame
velocity and stretching factor.
Entropy 2021,23, 475 10 of 23
T / TL
νnum / ν
0.4 0.8 1.2 1.6 2
1
2
3
4
5Second-order equilibrium
Product-form equilibrium
Ma = 0.1
Figure 2.
Numerical measurement of viscosity for axis-aligned setup at Mach number
Ma =
0.1 for
different temperatures. The exact solution corresponds to νnum/ν=1.
Ma
νnum / ν
0.1 0.2 0.3 0.4 0.5 0.6 0.7
0.5
1
1.5
2
2.5
3
3.5
Product-form equilibrium (λx=1)
Product-form equilibrium (λx=2)
Extended equilibrium (λx=1, 2)
T = 1/3
Figure 3.
Numerical measurement of viscosity for rotated setup at temperature
T=
1
/
3 for different
velocities and stretching ratios. The exact solution corresponds to νnum/ν=1.
3.2. Decaying Homogeneous Isotropic Turbulence
In order to further validate the model as a reliable method for the simulation of
complex flows and to show the application of using higher temperatures, decaying homo-
geneous isotropic turbulence was considered. The initial condition, in a box of the size
L×L×L
, was set at unit density and constant temperature along with a divergence-free
velocity field, which follows the specified energy spectrum,
E(κ) = Aκ4e2(κ/κ0)2, (67)
where
κ
is the wave number,
κ0
is the wave number at which the spectrum peaks and
A
is the parameter that controls the initial kinetic energy [
36
]. The initial velocity field is
Entropy 2021,23, 475 11 of 23
generated using a kinematic simulation as proposed in [
37
]. The turbulent Mach number is
defined as
Mat=u·u
cs, (68)
where
cs=T
is the speed of sound. The Reynolds number is based on the Taylor
microscale,
Λ2=u2
rms
(xux)2, (69)
and is given by
ReΛ=ρurms Λ
µ, (70)
where
urms =u·u/3
is the root mean square (rms) of the velocity and overbar denotes
the volume average over the entire computational domain.
Simulations were performed at
Mat=
0.1,
ReΛ=
72,
κ0=
16
π/L
, at two different
temperatures,
T=
1
/
3 and
T=
0.55, and with
L=
256 grid points. Figure 4shows a
snapshot of the velocity magnitude
u·u
at time
t=t/τ=
1.0, where
τ=LI/urms,0
is
the eddy turnover time, which is defined based on the initial rms of the velocity and the
integral length scale LI=2π/κ0.
Figure 4.
Velocity magnitude in lattice units for the decaying homogeneous isotropic turbulence at
Mat=0.1, ReΛ=72 and t=1.0 with temperature T=0.55.
To quantitatively assess the accuracy of the model at different temperatures, the time
evolution of the turbulent kinetic energy,
K=3
2u2
rms, (71)
normalized with its initial value (
K0
), and of the Taylor microscale Reynolds number are
compared in Figures 5and 6with results from direct numerical simulations (DNS) [
36
]. It
is apparent that the two working temperatures yields almost identical results that agree
well with the DNS simulation. This indicates that the correction terms do not degrade the
accuracy of the model at higher temperatures, even though the magnitude of error term
(30) is higher due to amplification of the linear term.
Entropy 2021,23, 475 12 of 23
t*
K / K0
12345
0.2
0.4
0.6
0.8
1
T = 1/3
T = 0.55
DNS
Figure 5. Time evolution of the turbulent kinetic energy for decaying isotropic turbulence at Mat=
0.1, ReΛ=72. Lines: present model; symbol: DNS [36].
t*
ReΛ
10-1 100
40
60
80
100
120
T = 1/3
T = 0.55
DNS
Figure 6.
Time evolution of the Taylor microscale Reynolds number for decaying isotropic turbulence
at Mat=0.1, ReΛ=72. Lines: present model; symbol: direct numerical simulations (DNS) [36].
The immediate advantage of using the present model at a temperature higher than the
lattice temperature
TL=
1
/
3 is that it effectively increases the characteristic velocity (here
urms,0
) and therefore the time step by a factor of
T/TL
. A larger time step is equivalent
to fewer number of time steps. The present model, therefore, speeds up the simulation
by a factor of
T/TL
compared to the conventional LBM, which can operate only at the
lattice temperature
TL
. Furthermore, this speedup strategy can be used for both steady
and unsteady flows. This is in contrast to the preconditioned LBM [
33
], which works
by altering the effective Mach number and therefore reduces the disparity between the
speeds of the acoustic wave propagation and the waves propagating with the fluid velocity,
cf. [33]
. This makes preconditioned LBM restricted to steady state applications. In contrast,
Entropy 2021,23, 475 13 of 23
the present model enables us to increase the speed of sound without changing the Mach
number. This increases the effective time step of the solver. Therefore, the present model
increases the computational efficiency by decreasing the number of required time steps.
Note that the theoretical temperature range of the model (like any other models based
on the D1Q3 lattice) is 0
T
1, beyond that the populations become negative and the
model is unstable. Therefore, while small temperature is possible but not beneficial, large
temperature greater than 1 is out of the stability domain
3.3. Periodic Double Shear Layer
The next validation case to test the accuracy of the proposed model with the stretched
lattice is the periodic double shear layer flow with the initial condition,
ux=u0tanh(α(y/L0.25)),yL/2,
u0tanh(α(0.75 y/L)),y>L/2, (72)
uy=δu0sin(2π(x/L+0.25)), (73)
where
L
is the domain length in both
x
and
y
directions,
u0=
0.1 is characteristic velocity,
δ=
0.05 is a perturbation of the
y
-velocity and
α=
80 controls the width of the shear layer.
The Reynolds number is set to Re =u0L/ν=104and the temperature is T=1/3.
Figure 7shows the vorticity field at non-dimensional time
t=tu0/L=
1 using the
conventional square lattice
λx=λy=
1 and the rectangular lattice with
λx=
2,
λy=
1.
Both lattice models perform qualitatively the same.
-0
0.5
l
Figure 7.
Vorticity field for double shear layer flow at
t=
1 with regular lattice (
left
) and stretched
lattice (right). Vorticity magnitude is normalized by its maximum value.
To quantify the effect of stretching on the accuracy, the time evolution of the mean
kinetic energy and of the mean enstrophy
=ω2/u2
0
L2
, with
ω
the vorticity magnitude,
are compared in Figure 8. The results show only minor discrepancies, which indicates the
validity of the model also on stretched meshes.
Entropy 2021,23, 475 14 of 23
t*
K / 0.5 u0
2
012345
0.86
0.88
0.9
0.92
0.94 256×512 (λx = 2)
512×512 (λx = 1)
t*
012345
100
150
200 256×512 (λx = 2)
512×512 (λx = 1)
Figure 8.
Evolution of kinetic energy (
left
) and enstrophy (
right
) for double shear layer flow at
Re =104.
3.4. Laminar Boundary Layer over a Flat Plate
The next test case validates our model for wall-bounded flows. We consider the
laminar flow over a flat plate with an incoming Mach number
Ma=u/T=
0.1,
temperature
T=
1
/
3 and Reynolds number
Re =ρuL/µ=
4000, where
L
is the
length of flat plate. Since the flow gradients in the transverse
y
-direction are much larger
compared to the gradients in the streamwise
x
-direction, the mesh can be stretched in
x
-direction without significantly affecting the accuracy of the results. The computational
domain was set to
[Lx×Ly]=[
200
×
200
]
and a rectangular lattice with
λx=
2 was used.
The flat plate starts at a distance of
Lx/
4 from the inlet and symmetry boundary conditions
were imposed at 0
xLx/
4. In Figure 9, the horizontal velocity profile at the end of
the plate is compared with the results of a regular lattice and with the Blasius similarity
solution, where ηis the dimensionless coordinate [38],
η=yru
νx. (74)
It can be seen that results for the regular and the rectangular lattice nearly coincide
and agree well with the Blasius solution. Thus, the model achieves accurate results with
half of grid points compared to the regular lattice. Furthermore, the distribution of skin
friction coefficient over the plate,
Cf=τwall
1
2ρu2
, (75)
with the wall shear stress
τwall =µ(u
y)y=0
, is shown in Figure 10 in comparison with the
analytical solution
Cf=
0.664
/Rex
, where
Rex=ux/ν
[
38
]. Also here, the results of
the model with the regular and the stretched velocities are almost identical and in good
agreement with the analytical solution.
Entropy 2021,23, 475 15 of 23
u / U
η
0 0.2 0.4 0.6 0.8 1
0
2
4
6
8
(400×200) (λx=1)
(200×200) (λx=2)
Figure 9.
Comparison of the velocity profile at
x=Lx
for flow over a flat plate at different stretching
ratios. Lines: present model; symbols: Blasius solution.
x / L
Cf
0.2 0.4 0.6 0.8 1
0.02
0.04
0.06
0.08
0.1
0.12
0.14
(400×200) (λx=1)
(200×200) (λx=2)
Figure 10.
Comparison of the skin friction coefficient for flow over a flat plate at different stretching
ratio. Lines: present model; symbols: analytical solution.
3.5. Turbulent Channel Flow
In the final test case, we assess the accuracy and performance of the extended LBM
for the turbulent flow in a rectangular channel, for which many numerical [
39
41
] and
experimental [
42
,
43
] results are available. The channel geometry was chosen as
[
5.6
H×
2H×2H], where His the channel half-width. The friction Reynolds number,
Reτ=uτH
ν, (76)
Entropy 2021,23, 475 16 of 23
based on the friction velocity
uτ=pτw/ρ
, was set to
Reτ=
180. The initial friction
velocity was estimated by
uτ=u0
1
KlnReτ+5.5 , (77)
where
K=
0.41 is the von Kármán constant and
u0=
0.1 is the mean center-line velocity.
Periodic boundary conditions were imposed in the streamwise
x
-direction and the spanwise
z-direction. The flow was driven by a constant body force in the x-direction,
g=Re2
τν2/H3. (78)
In order to accelerate the transition to turbulence, a non-uniform divergence-free forcing field
as proposed in [44] was added to the flow for some period of time, until t=tH/uτ=5.
Similar to the previous test case, grid stretching in
x
-direction with
λx=
1.4 was used
in order to reduce the number of grid points in that direction while the temperature was
set to T=0.55, same as in Section 3.2
A snapshot of the velocity magnitude u·uis shown in Figure 11.
Figure 11.
Snapshot of the velocity magnitude in lattice units for turbulent channel flow at
Reτ=
180
with λx=1.4.
Quantitatively, we compare the mean velocity profile with the DNS results of [
40
] in
Figure 12. In wall units, the mean velocity is given by
u+=¯
u/uτ
and the spatial coordinate
is
y+=yuτ/ν
. The statistics are collected after 30 eddy turnover times, i.e., after
t=
30.
It is apparent that the viscous sublayer (
y+<
5), the buffer layer (5
<y+<
30) and the
log-law region (
y+>
30) are captured well with our model and the mean velocity profile
agrees well with that of the DNS.
Entropy 2021,23, 475 17 of 23
y+
u+
100101102
0
5
10
15
20
Present
DNS
Figure 12.
Comparison of the mean velocity profile in a turbulent channel flow at
Reτ=
180 with
λx=1.4.
For a more thorough analysis, we compare the root mean square of the velocity
fluctuations with the DNS data in Figure 13. Here,
ux,rms =qu0
xu0
x
and
uy,rms
and
uz,rms
are defined in a similar way. It can be seen that the results are in excellent agreement
with the DNS results [
40
]. This demonstrates that the LBGK model, also in the presence
of a severe anisotropy triggered by stretched velocities, can be used for the simulation of
high Reynolds number wall-bounded flows once the corrections are incorporated with the
extended equilibrium.
y+
ux, rms, uy, rms, uz, rms
0 50 100 150
0
0.5
1
1.5
2
2.5
3
ux, rms
uy, rms
uz, rms
Figure 13.
Comparison of the rms of the velocity fluctuations in a turbulent channel flow at
Reτ=
180
with λx=1.4. Symbols: present model; lines: DNS [40].
4. Conclusions
While even with the standard discrete speeds (5) it is possible to develop an error-free,
fully Galilean invariant kinetic model in the co-moving reference frame, it does require
Entropy 2021,23, 475 18 of 23
off-lattice particles’ velocities [
45
,
46
]. Sticking with the fixed, lattice-conform velocities
(6)
, one is faced with an inevitable and persistent error, which spoils the hydrodynamic
equations whenever the flow velocity is increased or the temperature deviates from the
lattice reference value, or the discrete speeds are stretched differently in different directions.
We proposed an upgrade of the LBGK model to enlarge its operation domain in terms
of velocity, temperature and grid stretching by suggesting an extended equilibrium. The
extended equilibrium is realized through a product-form, which allows us to compensate
the diagonal third-order moment anomaly in the hydrodynamic limit by adding consistent
correction terms to the diagonal elements of the second-order moment. As a result, the
extended LBGK model restores Galilean invariance and temperature independence in a
sufficiently wide range, and can also be used with rectangular lattices. Similar to previous
proposals [
4
6
,
10
,
12
], the relaxation term of the present model remains almost local as
it uses only nearest-neighbor information for computation of the first-order derivatives
in the extended equilibrium populations. The extended LBGK model was validated in a
range of benchmark problems, probing different aspects of anomaly triggered either by
increased velocity or temperature deviation from the lattice reference temperature, or by
grid stretching. In all cases, the extended LBGK model featured excellent performance and
accuracy in both two and three dimensions. In particular, the simulation of homogeneous
isotropic turbulence demonstrated the expected speed-up when a higher temperature was
used, while simulations of the laminar boundary layer and of the turbulent channel flow
using stretched grids demonstrated good accuracy with a reduced number of grid points.
Furthermore, the present model can be extended to other applications including but
not limited to high-speed compressible flows, which can be achieved by incorporating
another solver for the total energy (see, e.g., the models proposed in [
10
,
12
]). Advanced
collision models, such as multiple relaxation times (MRT) schemes, can also readily be
employed in the present approach, which can be beneficial when running under-resolved
simulations. These two avenues shall be subject of further development and application of
the extended LBM.
Note added in proof: After the paper was submitted to peer-review, a later preprint
“Central Moment MRT Lattice Boltzmann Method on a Rectangular Lattice” by E.
Yahia and K. N. Premnath, arXiv:2103.02119 [physics.flu-dyn], became known to us
which adresses a correction for the two-dimensional rectangular lattice in a multiple
relaxation time setting.
Author Contributions:
Conceptualization, I.K.; methodology, M.H.S., B.D. and I.K.; analysis, M.H.S.
and I.K.; implementation, M.H.S.; simulation, M.H.S.; validation, M.H.S. and B.D.; visualization,
M.H.S.; simulation—supervision, B.D.; writing—original-draft preparation, M.H.S.; writing—review
and editing, I.K., M.H.S. and B.D.; supervision, I.K. All authors contributed to writing the paper. All
authors have read and agreed to the published version of the manuscript.
Funding:
This work was supported by the European Research Council (ERC) Advanced Grant
834763-PonD and the ETH Research Grant ETH-13 17-1. Computational resources at the Swiss
National Super Computing Center CSCS were provided under the Grant s897.
Data Availability Statement: Not applicable.
Acknowledgments:
We thank anonymous referee who suggested comparison with the local correc-
tion method of Ref. [
9
]. Authors are grateful to Seyed Ali Hosseini at ETHZ for computing spectral
dissipation presented in Figure A1.
Conflicts of Interest: The authors declare no conflict of interest.
Abbreviations
The following abbreviations are used in this manuscript:
LBM Lattice Boltzmann method
LBGK Lattice Bhatnagar–Gross–Krook
Entropy 2021,23, 475 19 of 23
Appendix A. Comparison of Extended LBGK to Locally Corrected LBM
Below, we compare the locally corrected lattice Boltzmann model (LC LBM) [
9
] with
both the standard and the present extended LBGK. To that end, it suffices to consider
the one-dimensional
D
1
Q
3 lattice. In order to introduce the LC LBM, we begin with the
standard LBGK (δt=1, R=1),
fi(x+vi,t+1)fi(x,t) = ω(feq
ifi). (A1)
The equilibrium populations in (A1) are given by (24) and (25),
feq
i=ρΨiux,Peq
xx , 1,Peq
xx =T+u2
x,i∈ {−1, 0, 1}. (A2)
Thanks to the diagonal anomaly, the second-order asymptotic analysis of Section 2.4
results in the following viscous stress in the one-dimensional version of the Navier–Stokes
Equation (61),
Πxx =21
ω1
2ρTxux+˜
Πxx . (A3)
For the LBGK model, the anomalous (second) term in (A3) reads,
˜
ΠLBGK
xx =1
ω1
213T
2T3u2
x
2T2ρTxux+ux(13T)u3
xxρ. (A4)
Upon realizing that the first term of the anomalous contribution (A4) is similar in its
structure to the relevant (first) term in the LBGK stress (A3), the locally corrected (LC)
LBGK groups these two terms together and replaces the relaxation parameter
ω
in (A1)
with a new relaxation
ωLC
, which depends on the flow velocity. While the original
work [9]
addressed the case of the lattice temperature,
T=TL=
1
/
3 (31), we first consider a slightly
more general formulation for a flexible temperature parameter. Consequently, the locally
corrected relaxation parameter ωLC in the LBGK Equation (A1) is defined as,
1
ωLC 1
2=1
ω1
2X, (A5)
where the renormalization factor Xreads,
X=1+13T
2T3u2
x
2T1
. (A6)
The LBGK model with the locally corrected relaxation parameter
ωLC
(A5) results in the
viscous stress of the form (A3), with the remaining error term,
˜
ΠLC
xx =1
ω1
2Xux(13T)u3
xxρ. (A7)
For the sake of a discussion, let us introduce the local Mach number,
Max=ux/T
. For
a quasi-incompressible (slow) flow, the density variation scales as
xρMa2
x
. Thus, for
T6=TL
, the error (A7) can be estimated as,
˜
ΠLC
xx Ma3
x
. This is two orders of magnitude
lower than the error of the original LBGK at
T6=TL
, cf. Equation (A4), at small Mach
number. Moreover, by setting the temperature
T=TL=
1
/
3, it was first realized in
Ref. [9]
that the error (A7) reduces to,
˜
ΠLC
xx =1
ω1
2 T3/2
LMa3
x
1(3/2)Ma2
x!xρ. (A8)
Entropy 2021,23, 475 20 of 23
In this case, the scaling at
Max
0 becomes,
˜
ΠLC
xx Ma5
x
. In other words, the local
correction at
T=TL
provides a gain of two orders of magnitude in accuracy with respect to
the standard LBGK under the quasi-incompressible flow conditions [
9
]. This consideration
extends straightforwardly to the
D
2
Q
9 and
D
3
Q
27 lattices by constructing a multiple
relaxation time LBM that corrects the relaxation of each diagonal component of the pressure
tensor [9].
However, for a generic isothermal flow, the error (A7) becomes amplified through
the renormalization factor (A6) as the velocity increases and eventually diverges when
u2
x(1T)/3
. This error persists also for the special case
T=TL
(A8). On the other hand,
the second-order analysis of Section 2.4 reveals that the present LBGK with the extended
equilibrium (33) removes the entire anomalous term,
˜
Πex
xx =
0. Thus, the difference between
the extended LBGK and the LC LBM [9] is expected beyond the asymptotic Max0.
In order to demonstrate this point, a spectral analysis was performed for the two-
dimensional
D
2
Q
9 lattice (see [
21
] for details of the spectral analysis in the LBM context).
The normalized spectral dissipation of acoustic modes
=(ωκ)/νκ2
x
, is shown in Figure A1,
for
T=TL
and the background flow velocity
(ux
,
uy) = (
0.3, 0
)
, for the three models, the
standard LBGK, the present extended LBGK and the LC LBM of Ref. [9].
It can be seen that, the extended LBGK recovers the correct dissipation rate in the
continuum limit (vanishing wave number
κx
), confirming its Galilean invariance. However,
both the standard LBGK and the LC LBM show deviations in the form of under-dissipation
at low wave numbers, while the deviation for the LC LBM is indeed smaller. This non-
vanishing deviation is amplified in cases with different working temperature and/or
non-unit stretching factor for both the standard LBGK and the LC LBM, which makes their
applications limited to the quasi-incompressible flow regime at the lattice temperature.
κx
(ωκ) / ν κ x
2
2
1.5
1
0.5
0
0π / 2 π
Figure A1.
Spectral dissipation of acoustic modes for different models. Red symbols: LBGK; black
symbols: extended LBM (33); blue symbols: LC LBM [
9
]; dashed line: Navier–Stokes. The velocity
and temperature are set to ux=0.3 and T=TL.
Entropy 2021,23, 475 21 of 23
x
ρ
0 200 400 600 800
0.4
0.6
0.8
1
1.2
1.4
1.6
LBGK
Extended LBM
LC LBM
Figure A2.
Comparison of density profile for shock tube problem at density ratio
ρl/ρr=
3, after
500 iterations. Solid line: LBGK; dashed line: extended LBM (33); symbols: LC LBM [9].
Finally, it is interesting to note that, in the case of shock capturing, all the three models
are expected to behave similarly, given that their respective dissipation rates at the wave
number
κx=π
are close in value, see Figure A1. This observation is confirmed by the
simulation of a shock tube with the following initial condition,
(ρ,ux,T) = (ρl, 0, 1/3),xL/2,
(ρr, 0, 1/3),x>L/2, (A9)
with
L=
800 grid points and viscosity
ν=
0.04. Results are presented in
Figures A2 and A3
for
ρl=
1.5,
ρr=
0.5, corresponding to the initial density ratio
ρl/ρr=
3.
Figures A2 and A3
demonstrate that all models produce almost indistinguishable results, with a similar oscil-
lation pattern at the shock front.
x
Ma
0 200 400 600 800
0
0.1
0.2
0.3
0.4
0.5
0.6
Figure A3.
Mach number profile,
Ma =u/TL
, for the shock tube problem at density ratio
ρl/ρr=
3,
after 500 iterations. Solid line: LBGK; dashed line: extended LBM (33); symbols: LC LBM [9].
Entropy 2021,23, 475 22 of 23
References
1. Succi, S. The Lattice Boltzmann Equation: For Complex States of Flowing Matter; Oxford University Press: Oxford, UK, 2018.
2.
Krüger, T.; Kusumaatmaja, H.; Kuzmin, A.; Shardt, O.; Silva, G.; Viggen, E.M. The Lattice Boltzmann Method; Springer International
Publishing: Switzerland, 2017.
3.
Qian, Y.; Orszag, S. Lattice BGK models for the Navier-Stokes equation: Nonlinear deviation in compressible regimes. EPL
Europhys. Lett. 1993,21, 255.
4.
Prasianakis, N.I.; Karlin, I.V. Lattice Boltzmann method for thermal flow simulation on standard lattices. Phys. Rev. E
2007
,
76, 016702.
5.
Prasianakis, N.I.; Karlin, I.V. Lattice Boltzmann method for simulation of compressible flows on standard lattices. Phys. Rev. E
2008,78, 016704.
6.
Prasianakis, N.I.; Karlin, I.V.; Mantzaras, J.; Boulouchos, K. Lattice Boltzmann method with restored Galilean invariance. Phys.
Rev. E 2009,79, 066702.
7.
Kang, J.; Prasianakis, N.I.; Mantzaras, J. Lattice Boltzmann model for thermal binary-mixture gas flows. Phys. Rev. E
2013
,
87, 053304.
8.
Li, Q.; Luo, K.; He, Y.; Gao, Y.; Tao, W. Coupling lattice Boltzmann model for simulation of thermal flows on standard lattices.
Phys. Rev. E 2012,85, 016710.
9.
Dellar, P.J. Lattice Boltzmann algorithms without cubic defects in Galilean invariance on standard lattices. J. Comput. Phys.
2014
,
259, 270–283.
10.
Saadat, M.H.; Bösch, F.; Karlin, I.V. Lattice Boltzmann model for compressible flows on standard lattices: Variable Prandtl number
and adiabatic exponent. Phys. Rev. E 2019,99, 013306.
11.
Guo, S.; Feng, Y.; Jacob, J.; Renard, F.; Sagaut, P. An efficient lattice Boltzmann method for compressible aerodynamics on D3Q19
lattice. J. Comput. Phys. 2020,418, 109570.
12.
Sawant, N.; Dorschner, B.; Karlin, I. Consistent lattice Boltzmann model for multicomponent mixtures. J. Fluid Mech.
2021
,909,
doi:10.1017/jfm.2020.853.
13.
Clausen, J.R.; Aidun, C.K. Galilean invariance in the lattice-Boltzmann method and its effect on the calculation of rheological
properties in suspensions. Int. J. Multiph. Flow 2009,35, 307–311.
14.
Bouzidi, M.; d’Humières, D.; Lallemand, P.; Luo, L.S. Lattice Boltzmann equation on a two-dimensional rectangular grid. J.
Comput. Phys. 2001,172, 704–717.
15. Chikatamarla, S.; Karlin, I. Comment on “Rectangular lattice Boltzmann method”. Phys. Rev. E 2011,83, 048701.
16.
Frapolli, N.; Chikatamarla, S.S.; Karlin, I.V. Theory, analysis, and applications of the entropic lattice Boltzmann model for
compressible flows. Entropy 2020,22, 370.
17. Wagner, A.; Yeomans, J. Phase separation under shear in two-dimensional binary fluids. Phys. Rev. E 1999,59, 4366.
18. Dellar, P.J. Bulk and shear viscosities in lattice Boltzmann equations. Phys. Rev. E 2001,64, 031203.
19. Geier, M.; Schönherr, M.; Pasquali, A.; Krafczyk, M. The cumulant lattice Boltzmann equation in three dimensions: Theory and
validation. Comput. Math. Appl. 2015,70, 507–547.
20.
Feng, Y.; Sagaut, P.; Tao, W.Q. A compressible lattice Boltzmann finite volume model for high subsonic and transonic flows on
regular lattices. Comput. Fluids 2016,131, 45–55.
21.
Hosseini, S.A.; Darabiha, N.; Thévenin, D. Compressibility in lattice Boltzmann on standard stencils: Effects of deviation from
reference temperature. Philos. Trans. R. Soc. A 2020,378, 20190399.
22.
Patil, D.V.; Lakshmisha, K. Finite volume TVD formulation of lattice Boltzmann simulation on unstructured mesh. J. Comput.
Phys. 2009,228, 5262–5279.
23.
Hejranfar, K.; Ghaffarian, A. A high-order accurate unstructured spectral difference lattice Boltzmann method for computing
inviscid and viscous compressible flows. Aerosp. Sci. Technol. 2020,98, 105661.
24.
Krämer, A.; Küllmer, K.; Reith, D.; Joppich, W.; Foysi, H. Semi-Lagrangian off-lattice Boltzmann method for weakly compressible
flows. Phys. Rev. E 2017,95, 023305.
25.
Di Ilio, G.; Dorschner, B.; Bella, G.; Succi, S.; Karlin, I.V. Simulation of turbulent flows with the entropic multirelaxation time
lattice Boltzmann method on body-fitted meshes. J. Fluid Mech. 2018,849, 35–56.
26.
Saadat, M.H.; Bösch, F.; Karlin, I.V. Semi-Lagrangian lattice Boltzmann model for compressible flows on unstructured meshes.
Phys. Rev. E 2020,101, 023311.
27.
Wilde, D.; Krämer, A.; Reith, D.; Foysi, H. Semi-Lagrangian lattice Boltzmann method for compressible flows. Phys. Rev. E
2020
,
101, 053306.
28.
Lagrava, D.; Malaspinas, O.; Latt, J.; Chopard, B. Advances in multi-domain lattice Boltzmann grid refinement. J. Comput. Phys.
2012,231, 4808–4822.
29.
Dorschner, B.; Frapolli, N.; Chikatamarla, S.S.; Karlin, I.V. Grid refinement for entropic lattice Boltzmann models. Phys. Rev. E
2016,94, 053311.
30.
Zong, Y.; Peng, C.; Guo, Z.; Wang, L.P. Designing correct fluid hydrodynamics on a rectangular grid using MRT lattice Boltzmann
approach. Comput. Math. Appl. 2016,72, 288–310.
31.
Peng, C.; Min, H.; Guo, Z.; Wang, L.P. A hydrodynamically-consistent MRT lattice Boltzmann model on a 2D rectangular grid. J.
Comput. Phys. 2016,326, 893–912.
Entropy 2021,23, 475 23 of 23
32.
Zecevic, V.; Kirkpatrick, M.P.; Armfield, S.W. Rectangular lattice Boltzmann method using multiple relaxation time collision
operator in two and three dimensions. Comput. Fluids 2020,202, 104492.
33. Guo, Z.; Zhao, T.S.; Shi, Y. Preconditioned lattice-Boltzmann method for steady flows. Phys. Rev. E 2004,70, 066706.
34.
Karlin, I.; Asinari, P. Factorization symmetry in the lattice Boltzmann method. Phys. A Stat. Mech. Its Appl.
2010
,389, 1530–1548.
35.
Dorschner, B.; Chikatamarla, S.S.; Bösch, F.; Karlin, I.V. Grad’s approximation for moving and stationary walls in entropic lattice
Boltzmann simulations. J. Comput. Phys. 2015,295, 340–354.
36.
Samtaney, R.; Pullin, D.I.; Kosovi´c, B. Direct numerical simulation of decaying compressible turbulence and shocklet statistics.
Phys. Fluids 2001,13, 1415–1430.
37.
Ducasse, L.; Pumir, A. Inertial particle collisions in turbulent synthetic flows: quantifying the sling effect. Phys. Rev. E
2009
,
80, 066312.
38. White, F.M. Fluid Mechanics; Tata McGraw-Hill Education: New York, USA, 1979.
39.
Kim, J.; Moin, P.; Moser, R. Turbulence statistics in fully developed channel flow at low Reynolds number. J. Fluid Mech.
1987
,
177, 133–166.
40.
Moser, R.D.; Kim, J.; Mansour, N.N. Direct numerical simulation of turbulent channel flow up to
Reτ=
590. Phys. Fluids
1999
,
11, 943–945.
41. Lee, M.; Moser, R.D. Direct numerical simulation of turbulent channel flow up to Re 5200. J. Fluid Mech. 2015,774, 395–415.
42.
Eckelmann, H. The structure of the viscous sublayer and the adjacent wall region in a turbulent channel flow. J. Fluid Mech.
1974
,
65, 439–459.
43.
Kreplin, H.P.; Eckelmann, H. Behavior of the three fluctuating velocity components in the wall region of a turbulent channel flow.
Phys. Fluids 1979,22, 1233–1239.
44.
Wang, L.P.; Peng, C.; Guo, Z.; Yu, Z. Lattice Boltzmann simulation of particle-laden turbulent channel flow. Comput. Fluids
2016
,
124, 226–236.
45. Dorschner, B.; Bösch, F.; Karlin, I.V. Particles on Demand for Kinetic Theory. Phys. Rev. Lett. 2018,121, 130602(5).
46.
Reyhanian, E.; Dorschner, B.; Karlin, I.V. Thermokinetic lattice Boltzmann model of nonideal fluids. Phys. Rev. E
2020
,102, 020103.
... For example, Hegeler et al. [16] developed a rectangular SRT-LB model that can derive the correct macroscopic equations by adding discrete velocity directions to increase degrees of freedom. Peng et al. [18] and Saadat et al. [19] respectively proposed another SRT-LB model on a rectangular grid that can reproduce the correct NSEs by introducing an extended equilibrium distribution function. Besides, Wang et al. [20] constructed a SRT-LB model through including some artificial counteracting forcing terms, which are used to remove the anisotropy caused by the rectangular lattice. ...
... It can be found from Eqs. (19) and (20) that when c s1 = c s2 = c s3 , one can get c 1 = c 2 = c 3 , which implies that there are no rD3Q15 and rD3Q13 lattice models. ...
Preprint
In this paper, we develop a general rectangular multiple-relaxation-time lattice Boltzmann (RMRT-LB) method for the Navier-Stokes equations (NSEs) and nonlinear convection-diffusion equation (NCDE) by extending our recent unified framework of MRT-LB method [Z.H. Chai and B.C. Shi, Phys. Rev. E 102, 023306 (2020)], where a rectangular equilibrium distribution function (REDF) [J.H. Lu et al, Phil. Trans. R. Soc. A 369, 2311-2319 (2011)] on a rectangular lattice is utilized. Due to the anisotropy of discrete velocities on a rectangular lattice, the third-order moment of REDF is inconsistent with that of popular LB method, and thus the previous unified framework of MRT-LB method cannot be directly applied to the NSEs using the REDF on the rectangular rDdQq lattice. The macroscopic NSEs can be recovered from the RMRT-LB method through the direct Taylor expansion method by properly selecting the relaxation sub-matrix $\mathbf{S}_2$ which is related to kinetic viscosity and bulk viscosity. While the rectangular lattice does not lead to the change of the zero-th, first and second-order moments of REDF, thus the unified framework of MRT-LB method can be directly applied to the NCDE. It should be noted that the RMRT-LB model for NSEs can be derived on the rDdQq lattice, including rD2Q9, rD3Q19, and rD3Q27 lattices, and the RMRT-LB versions (if existed) of the previous MRT-LB models can be obtained, including those based on raw (natural)-moment, central-moment, Hermite-moment, and central Hermite-moment, respectively.
... Martin et al. [37] proposed a methodology for simulating multiphase flow in heterogeneous porous media using the lattice Boltzmann method, combining grayscale with a multi-component single equation approach. Saadat et al. [38] restored the Galilean invariance and isotropy of the stress tensor by introducing extended equilibria, expanding the lattice Boltzmann model to handle simulations with higher flow rate values, and enhancing computational efficiency by reducing the required number of time steps. Zhao et al. [39] employed micron CT scanning imaging experiments and the lattice Boltzmann simulation method to qualitatively and quantitatively characterize of the microscopic pore structure of tight sandstone and conduct flow simulations, consequently determining the permeability of tight sandstone samples. ...
Article
Full-text available
Steel cord materials were found to have internal porous microstructures and complex fluid flow properties. However, current studies have rarely reported the transport behavior of steel cord materials from a microscopic viewpoint. The computed tomography (CT) scanning technology and lattice Boltzmann method (LBM) were used in this study to reconstruct and compare the real three-dimensional (3D) pore structures and fluid flow in the original and tensile (by loading 800 N force) steel cord samples. The pore-scale LBM results showed that fluid velocities increased as displacement differential pressure increased in both the original and tensile steel cord samples, but with two different critical values of 3.3273 Pa and 2.6122 Pa, respectively. The original steel cord sample had higher maximal and average seepage velocities at the 1/2 sections of 3D construction images than the tensile steel cord sample. These phenomena should be attributed to the fact that when the original steel cord sample was stretched, its porosity decreased, pore radius increased, flow channel connectivity improved, and thus flow velocity increased. Moreover, when the internal porosity of tensile steel cord sample was increased by 1 time, lead the maximum velocity to increase by 1.52 times, and the average velocity was increased by 1.66 times. Furthermore, when the density range was determined to be 0–38, the pore phase showed the best consistency with the segmentation area. Depending on the Zou-He Boundary and Regularized Boundary, the relative error of simulated average velocities was only 0.2602 percent.
... However, despite the high efficiency and low numerical dissipation of LBM for nearly incompressible flows, the domain of high-speed compressible flows presents a number of severe challenges [18,19,[30][31][32]. The main directions to extend conventional LBM towards the compressible realm includes standard LBM augmented with correction terms [33][34][35][36], multispeed lattices [37][38][39][40][41], and hybrid approaches [42][43][44][45]. ...
Article
Full-text available
The particles on demand method [Phys. Rev. Lett. 121, 130602 (2018)] was recently formulated with a conservative finite-volume discretization and validated against challenging benchmarks. In this work, we focus on the properties of the reference frame transformation and its implications on the accuracy of the model. Based on these considerations, we propose strategies that simplify the scheme and generalize it to include a tunable Prandtl number via quasi-equilibrium relaxation. Finally, we adapt concepts from the multiscale semi-Lagrangian lattice Boltzmann formulation to the proposed framework, further improving the potential and the operating range of the kinetic model. Numerical simulations of high Mach compressible flows demonstrate excellent accuracy and stability of the model over a wide range of conditions.
... Methods exist in which LBM adapts to a stretched [30] or nonrectangular [31] mesh. The fact that a composition of different stencils was not previously used for adaptive grids is due to the following issue: A lattice with varied geometry requires different discrete velocity sets in some nodes, and a discrete particle distribution function corresponds to each velocity. ...
Article
Full-text available
Grid refinement is used to reduce computing costs while maintaining the precision of fluid simulation. In the lattice Boltzmann method (LBM), grid refinement often uses interpolated values. Here, we developed a method in which interpolation in space and time is not required. For this purpose, we used the moment matching condition and rescaled the nonequilibrium part of the populations, thereby developing a recalibration procedure that allows for the transfer of information between different LBM stencils in the simulation domain. Then, we built a nonuniform lattice that uses stencils with different shapes on the transition. The resulting procedure was verified by performing benchmarks with the 2D Poisselle flow and the advected vortex. It is suggested that grids with adaptive geometry can be built with the proposed method.
... Methods, where LBM adapts to a stretched [30] or non-rectangular [31] mesh exist. 75 The fact that a composition of different stencil was not used for adaptive grids before is due 76 to the following issue. ...
Preprint
Full-text available
Grid refinement is used to save computing costs while preserving high precision of fluid simulation. In the Lattice Boltzmann Method, grid refinement often uses interpolated values. We found a method, where interpolation in space and time is not required. For this purpose, we used the moment matching condition and rescaling the non-equilibrium part of populations, and developed the recalibration procedure that allows to transfer information between different LBM stencils in the simulation domain. Then, we build non-uniform lattice which uses stencils with different shapes on the transition. The resulting procedure is verified on the 2D Poisselle flow and the advected vortex benchmark. In prospect, grids with adaptive geometry can be built with the use of the proposed method.
... high efficiency and low numerical dissipation of LBM for nearly incompressible flows, the domain of high-speed compressible flows presents a number of severe challenges [18,[30][31][32][33]. The main directions to extend conventional LBM towards the compressible realm includes standard lattices LBM augmented with correction terms [34][35][36][37], multi-speed lattices [38][39][40][41][42] and hybrid approaches [43][44][45][46]. ...
Preprint
Full-text available
The Particles on Demand method [B. Dorschner, F. B\"{o}sch and I. V. Karlin, {\it Phys. Rev. Lett.} {\bf 121}, 130602 (2018)] was recently formulated with a conservative finite volume discretization and validated against challenging benchmarks. In this work, we rigorously analyze the properties of the reference frame transformation and its implications on the accuracy of the model. Based on these considerations, we propose strategies to boost the efficiency of the scheme and to reduce the computational cost. Additionally, we generalize the model such that it includes a tunable Prandlt number via quasi-equilibrium relaxation. Finally, we adapt concepts from the multi-scale semi-Lagrangian lattice Boltzmann formulation to the proposed framework, further improving the potential and the operating range of the kinetic model. Numerical simulations of high Mach compressible flows demonstrate excellent accuracy and stability of the model over a wide range of conditions.
... The LBM models the macroscopic velocity and temperature as deviations from these reference values. Large deviations from the reference values can lead to instability or errors, although flows with Mach numbers as high as 2 36,44 and temperature ratios as high as 10 42,45 are feasible using more sophisticated LB models. In order to circumvent the aforementioned errors due to large deviations, PonD describes the evolution of the population in a local reference frame λ(u(x, t), T(x, t)), where u(x, t) is the local macroscopic fluid velocity and T(x, t) is the local fluid temperature at a given location in space and time. ...
Article
Full-text available
A kinetic model based on the particles on demand method is introduced for gas phase detonation hydrodynamics in conjunction with the Lee–Tarver reaction model. The proposed model is realized on two- and three-dimensional lattices and is validated with a set of benchmarks. Quantitative validation is performed with the Chapman–Jouguet theory up to a detonation wave speed of Mach 20 in one dimension. Two-dimensional outward expanding circular detonation is tested for isotropy of the model as well as for the asymptotic detonation wave speed. Mach reflection angles are verified in setups consisting of interacting strong bow shocks emanating from detonation. Spherical detonation is computed to show the viability of the proposed model for three-dimensional simulations.
... However, the gain in both the temperature and Mach number is perceived as moderate when compared to the increased complexity of the higher-order lattices. A second approach uses standard lattices, while LBM is augmented with corrections terms tailored to eliminate error terms in momentum and energy equations [22][23][24][25]. While trans-and supersonic flows can be captured efficiently with this approach, it, too, remained limited to moderate Mach number and discontinuities. ...
Article
Full-text available
Particles on Demand formulation of kinetic theory [B. Dorschner, F. B\"{o}sch and I. V. Karlin, Phys. Rev. Lett. 121, 130602 (2018)] is used to simulate a variety of compressible flows with strong discontinuities in density, pressure and velocity. Two modifications are applied to the original formulation of the Particles on Demand method. First, a regularization by Grad's projection of particles populations is combined with the reference frame transformations in order to enhance stability and accuracy. Second, a finite-volume scheme is implemented which allows tight control of mass, momentum and energy conservation. The proposed model is validated with an array of challenging one- and two-dimensional benchmarks of compressible flows, including hypersonic and near-vacuum situations, Richtmyer-Meshkov instability, double Mach reflection and astrophysical jet. Excellent performance of the modified Particles on Demand method is demonstrated beyond the limitations of other lattice Boltzmann-like approaches to compressible flows.
Article
In this paper, we develop a general rectangular multiple-relaxation-time lattice Boltzmann (RMRT-LB) method for the Navier-Stokes equations (NSEs) and nonlinear convection-diffusion equation (NCDE) by extending our recent unified framework of the multiple-relaxation-time lattice Boltzmann (MRT-LB) method [Chai and Shi, Phys. Rev. E 102, 023306 (2020)], where an equilibrium distribution function (EDF) [Lu et al., Philos. Trans. R. Soc. A 369, 2311 (2011)] on a rectangular lattice is utilized. The anisotropy of the lattice tensor on a rectangular lattice leads to anisotropy of the third-order moment of the EDF, which is inconsistent with the isotropy of the viscous stress tensor of the NSEs. To eliminate this inconsistency, we extend the relaxation matrix related to the dynamic and bulk viscosities. As a result, the macroscopic NSEs can be recovered from the RMRT-LB method through the direct Taylor expansion method. Whereas the rectangular lattice does not lead to the change of the zero-, first- and second-order moments of the EDF, the unified framework of the MRT-LB method can be directly applied to the NCDE. It should be noted that the RMRT-LB model for NSEs can be derived on the rDdQq (q discrete velocities in d-dimensional space, d≥1) lattice, including rD2Q9, rD3Q19, and rD3Q27 lattices, while there are no rectangular D3Q13 and D3Q15 lattices within this framework of the RMRT-LB method. Thanks to the block-lower triangular relaxation matrix introduced in the unified framework, the RMRT-LB versions (if existing) of the previous MRT-LB models can be obtained, including those based on raw (natural) moment, central moment, Hermite moment, and central Hermite moment. It is also found that when the parameter cs is an adjustable parameter in the standard or rectangular lattice, the present RMRT-LB method becomes a kind of MRT-LB method for the NSEs and NCDE, and the commonly used MRT-LB models on the DdQq lattice are only its special cases. We also perform some numerical simulations, and the results show that the present RMRT-LB method can give accurate results and also have a good numerical stability.
Article
Full-text available
A new lattice Boltzmann model (LBM) for chemically reactive mixtures is presented. The approach capitalizes on the recently introduced thermodynamically consistent LBM for multicomponent mixtures of ideal gases. Similar to the non-reactive case, the present LBM features Stefan–Maxwell diffusion of chemical species and a fully on-lattice mean-field realization of the momentum and energy of the flow. Besides introducing the reaction mechanism into the kinetic equations for the species, the proposed LBM also features a new realization of the compressible flow by using a concept of extended equilibrium on a standard lattice in three dimensions. The full thermodynamic consistency of the original non-reactive multicomponent LBM enables us to extend the temperature dynamics to the reactive mixtures by merely including the enthalpy of formation in addition to the sensible energy considered previously. Furthermore, we describe in detail the boundary conditions to be used for reactive flows of practical interest. The model is validated against a direct numerical simulation of various burning regimes of a hydrogen/air mixture in a microchannel, in two and three dimensions. Excellent comparison in these demanding benchmarks indicates that the proposed LBM can be a valuable and universal model for complex reactive flows.
Article
Full-text available
A new lattice Boltzmann model for multicomponent ideal gas mixtures is presented. The model development consists of two parts. First, a new kinetic model for Stefan-Maxwell diffusion amongst the species is proposed and realized as a lattice Boltzmann equation on the standard discrete velocity set. Second, a compressible lattice Boltzmann model for the momentum and energy of the mixture is established. Both parts are consistently coupled through mixture composition, momentum, pressure, energy and enthalpy whereby a passive scalar advection-diffusion coupling is obviated, unlike in previous approaches. The proposed model is realized on the standard three-dimensional lattices and is validated with a set of benchmarks highlighting various physical aspects of compressible mixtures. Stefan-Maxwell diffusion is tested against experiment and theory of uphill diffusion of argon and methane in a ternary mixture with hydrogen. The speed of sound is measured in various binary and ternary compositions. We further validate the Stefan-Maxwell diffusion coupling with hydrodynamics by simulating diffusion in opposed jets and the three-dimensional Kelvin-Helmholtz instability of shear layers in a two-component mixture. Apart from the multicomponent compressible mixture, the proposed lattice Boltzmann model also provides an extension of the lattice Boltzmann equation to the compressible flow regime on the standard three-dimensional lattice.
Article
Full-text available
We present a kinetic model for nonideal fluids, where the local thermodynamic pressure is imposed through appropriate rescaling of the particle's velocities, accounting for both long-and short-range effects and hence full thermodynamic consistency. The model features full Galilean invariance together with mass, momentum, and energy conservation and enables simulations ranging from subcritical to supercritical flows, which is illustrated on various benchmark flows such as anomalous shock waves or shock droplet interaction.
Article
Full-text available
With growing interest in the simulation of compressible flows using the lattice Boltzmann (LB) method, a number of different approaches have been developed. These methods can be classified as pertaining to one of two major categories: (i) solvers relying on high-order stencils recovering the Navier–Stokes–Fourier equations, and (ii) approaches relying on classical first-neighbour stencils for the compressible Navier–Stokes equations coupled to an additional (LB-based or classical) solver for the energy balance equation. In most cases, the latter relies on a thermal Hermite expansion of the continuous equilibrium distribution function (EDF) to allow for compressibility. Even though recovering the correct equation of state at the Euler level, it has been observed that deviations of local flow temperature from the reference can result in instabilities and/or over-dissipation. The aim of the present study is to evaluate the stability domain of different EDFs, different collision models, with and without the correction terms for the third-order moments. The study is first based on a linear von Neumann analysis. The correction term for the space- and time-discretized equations is derived via a Chapman–Enskog analysis and further corroborated through spectral dispersion–dissipation curves. Finally, a number of numerical simulations are performed to illustrate the proposed theoretical study. This article is part of the theme issue ‘Fluid dynamics, soft matter and complex systems: recent results and new methods’.
Article
Full-text available
This work thoroughly investigates a semi-Lagrangian lattice Boltzmann (SLLBM) solver for compressible flows. In contrast to other LBM for compressible flows, the vertices are organized in cells, and interpolation polynomials up to fourth order are used to attain the off-vertex distribution function values. Differing from the recently introduced Particles on Demand (PoD) method [Dorschner, Bösch, and Karlin, Phys. Rev. Lett. 121, 130602 (2018)], the method operates in a static, nonmoving reference frame. Yet the SLLBM in the present formulation grants supersonic flows and exhibits a high degree of Galilean invariance. The SLLBM solver allows for an independent time step size due to the integration along characteristics and for the use of unusual velocity sets, like the D2Q25, which is constructed by the roots of the fifth-order Hermite polynomial. The properties of the present model are shown in diverse example simulations of a two-dimensional Taylor-Green vortex, a Sod shock tube, a two-dimensional Riemann problem, and a shock-vortex interaction. It is shown that the cell-based interpolation and the use of Gauss-Lobatto-Chebyshev support points allow for spatially high-order solutions and minimize the mass loss caused by the interpolation. Transformed grids in the shock-vortex interaction show the general applicability to nonuniform grids.
Article
Full-text available
The entropic lattice Boltzmann method for the simulation of compressible flows is studied in detail and new opportunities for extending operating range are explored. We address limitations on the maximum Mach number and temperature range allowed for a given lattice. Solutions to both these problems are presented by modifying the original lattices without increasing the number of discrete velocities and without altering the numerical algorithm. In order to increase the Mach number, we employ shifted lattices while the magnitude of lattice speeds is increased in order to extend the temperature range. Accuracy and efficiency of the shifted lattices are demonstrated with simulations of the supersonic flow field around a diamond-shaped and NACA0012 airfoil, the subsonic, transonic, and supersonic flow field around the Busemann biplane, and the interaction of vortices with a planar shock wave. For the lattices with extended temperature range, the model is validated with the simulation of the Richtmyer–Meshkov instability. We also discuss some key ideas of how to reduce the number of discrete speeds in three-dimensional simulations by pruning of the higher-order lattices, and introduce a new construction of the corresponding guided equilibrium by entropy minimization.
Article
Full-text available
Compressible lattice Boltzmann model on standard lattices [M. H. Saadat, F.Bösch and I. V. Karlin, Phys. Rev. E 99, 013306 (2019)] is extended to deal with complex flows on unstructured grid. Semi-Lagrangian propagation [A. Kramer et al., Phys. Rev. E 95, 023305 (2017)] is performed on an unstructured second-order accurate finite element mesh and a consistent wall boundary condition is implemented which makes it possible to simulate compressible flows over complex geometries. The model is validated through simulation of Sod shock tube, subsonic/supersonic flow over NACA0012 airfoil and shock-vortex interaction in Schardin's problem. Numerical results demonstrate that the present model on standard lattices is able to simulate compressible flows involving shock waves on unstructured meshes with good accuracy and without using any artificial dissipation or limiter.
Article
Full-text available
A lattice Boltzmann model for compressible flows on standard lattices is developed and analyzed. A consistent two-population thermal lattice Boltzmann is used which allows a variable Prandtl number and a variable adiabatic exponent, and appropriate correction terms are introduced into the kinetic equations to compensate for deviations in the hydrodynamic limit. Using the concept of a shifted lattice, the model is extended to supersonic flows involving shock waves, and the shock-vortex interaction problem is simulated to show the accuracy of the proposed model. Numerical results demonstrate that the proposed model is a viable candidate for compressible flow simulations.
Article
An efficient lattice Boltzmann (LB) model relying on a hybrid recursive regularization (HRR) collision operator on D3Q19 stencil is proposed for the simulation of three-dimensional high-speed compressible flows in both subsonic and supersonic regimes. An improved thermal equilibrium distribution function on D3Q19 lattice is derived to reduce the complexity of correcting terms. A simple shock capturing scheme and an upwind biased discretization of correction terms are implemented for supersonic flows with shocks. Mass and momentum equations are recovered by an efficient streaming, collision and forcing process on D3Q19 lattice. Then a non-conservative formulation of the entropy evolution equation is used, that is solved using a finite volume method. The proposed method is assessed considering the simulation of i) 2D isentropic vortex convection, ii) 3D non-isothermal acoustic pulse, iii) 2D supersonic flow over a bump, iv) 3D shock explosion in a box, v) 2D vortex interaction with shock wave, vi) 2D laminar flows over a flat plate at Ma of 0.5, 1.0 and 1.5.
Article
We present a lattice Boltzmann (lb) method using a rectangular, non-isotropic lattice based on d2q9 and d3q27 velocity sets in two and three dimensions. A second order multi-scale expansion ensures that the scheme correctly reproduces hydrodynamic behaviour. A novel set of basis vectors is introduced in order to allow independent adjustment of eigenvalues corresponding to second order moments as required in order to ensure correct hydrodynamic behaviour using the non-isotropic lattice. Errors are reduced compared to other rectangular grid implementations. Linear perturbation analysis indicates that our scheme has similar stability properties to the isotropic lb method. We investigate the error behaviour of our scheme by performing Taylor-Green vortex flow simulations and comparing our results to simulations using a square grid and also to analytical results. We demonstrate that our scheme is well suited to direct numerical simulation of wall bounded turbulent flows and compare to well known benchmark results.
Article
In the present work, the spectral difference lattice Boltzmann method (SDLBM) is implemented on unstructured meshes for the solution methodology to be capable of accurately simulating the compressible flows over complex geometries. Both the inviscid and viscous compressible flows are computed by applying the unstructured SDLBM. The compressible form of the discrete Boltzmann–BGK equation with the Watari model is considered and the solution of the resulting system of equations is obtained by applying the spectral difference method on arbitrary quadrilateral meshes. The accuracy and robustness of the unstructured SDLBM for simulating the compressible flows are demonstrated by simulating four problems that are steady inviscid supersonic flow past a bump, steady inviscid subsonic flow over the two-element NACA 4412-4415 airfoil with and without the ground effect, steady viscous transonic flow around the NACA 0012 airfoil and unsteady viscous subsonic flow past two side-by-side cylinders. The results obtained by applying the unstructured SDLBM are in good agreement with those of the available high-order accurate Euler/Navier-Stokes solvers and also the experimental data. The present study introduces the unstructured SDLBM as an appropriate inviscid and viscous compressible LBM flow solver for accurately simulating fluid flows over practical problems.