Lectures On The Lattice-Boltzmann Method
Lectures On The Lattice-Boltzmann Method
C Philippi
Lectures on the
Lattice-Boltzmann
method
Lectures on the Lattice Boltzmann method
Paulo C Philippi
Mechanical Engineering Graduate Program.
Pontical Catholic University of Paraná.
Curitiba, PR, Brazil
Preface
This set of lectures is intended to be a rst course in the Lattice-Boltzmann
Method (LBM) for undergraduate and graduate students, who, in addition to
learn the method, want to understand the theoretical fundamentals behind it.
LBM is a kinetic numerical method and this means that the behavior of a
physical system, usually a uid, is numerically predicted by solving a kinetic
model of this system.
In the LBM framework, uids are considered as a system of particles and
we have, mainly, three hierarchic scales in studying such systems: a) the mi-
croscopic or molecular scale; b) the mesoscopic or kinetic scale and c) the
macroscopic scale.
The microscopic or molecular scale is the subject of the Molecular Dynam-
ics, in which the motion of each individual molecule is followed by solving its
equation of motion, usually, the Newton's second law of mechanics and con-
sidering all the interaction forces with the other molecules and the boundaries.
Molecular dynamics simulations are usually performed in very large clusters of
computers and the main physical parameters of the whole system is obtained
through statistical ensemble averages.
The mesoscopic or kinetic framework is a less detailed level of description,
where the interest is not to follow each individual particle along its trajectory,
but to nd the expected number of particles that will be found in a given state,
or the distribution f of particles over the several possible states of the system.
The state of a particle is usually given by its molecular translational velocity, ξ,
but it can be also given by the angular velocity, ω , when the particle shape and
rotational degrees of freedom are considered and by the vibrational quantum
states when the molecular modes of internal vibration are considered. In this
level of description, the distribution f is found by solving a kinetic equation.
Historically, the rst and more well-known kinetic equation is the Boltzmann
equation, derived by Ludwig Boltzmann, in the late 19th century.
Finally, the macroscopic level of description is the less detailed one, where
the interest is to follow the time and space evolution of a given set of moments
of the distribution function f, the density, ρ, the macroscopic velocity u and
the temperature T, usually given in terms of a set of balance equations, the
Fourier-Navier-Stokes equations.
The purpose of LBM is not to nd the true distribution f of particles, but to
consistently retrieve the macroscopic behavior of a physical system. Therefore,
numerically solving a kinetic equation, the framework of LBM, must, rstly,
be thought as a method for solving a physical problem, with, in certain cases,
some advantages with respect to classical CFD due to its intrinsic Lagrangian
nature and hyperbolic basis.
In the reverse sense, the LBM framework enables to reveal, or to put in
evidence, the inuence of a number of molecular processes on the macroscopic
behavior of a physical system. In fact, the kinetic level of description, may be
thought as a two-way bridge linking the molecular and the macroscopic levels
of description. Therefore, although derived from the molecular domain and,
4
Paulo C Philippi
1 Introduction 15
1.1 Early history . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.1.1 Molecular dynamics (MD) . . . . . . . . . . . . . . . . . 15
1.1.2 MD simulation of phase transition . . . . . . . . . . . . 17
1.1.3 Lattice Gas Automata (LGA) . . . . . . . . . . . . . . . 17
1.2 Towards the lattice-Boltzmann method . . . . . . . . . . . . . . 21
1.3 Connection with the Boltzmann equation . . . . . . . . . . . . 23
1.3.1 High-order lattice Boltzmann . . . . . . . . . . . . . . . 24
4 Forces in LBM 63
4.1 Lattice Boltzmann equation . . . . . . . . . . . . . . . . . . . . 63
4.1.1 Discretization of the stream term . . . . . . . . . . . . . 65
4.2 A smart procedure for overcoming the impliciteness of the nu-
merical scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.3 Another alternative: changing variables . . . . . . . . . . . . . 68
5
6 CONTENTS
6 Chapman-Enskog analysis 99
6.1 The continuous case . . . . . . . . . . . . . . . . . . . . . . . . 99
6.1.1 Macroscopic equations . . . . . . . . . . . . . . . . . . . 102
6.1.2 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
6.2 Scale analysis of a LB equation . . . . . . . . . . . . . . . . . . 107
6.3 Zero-th order transport equations . . . . . . . . . . . . . . . . . 109
6.3.1 Mass conservation . . . . . . . . . . . . . . . . . . . . . 109
6.3.2 Momentum balance . . . . . . . . . . . . . . . . . . . . 110
6.3.3 Energy conservation . . . . . . . . . . . . . . . . . . . . 111
6.3.4 Internal energy balance equation . . . . . . . . . . . . . 112
6.4 First-order transport equations . . . . . . . . . . . . . . . . . . 113
6.4.1 Mass conservation equation . . . . . . . . . . . . . . . . 114
6.4.2 Momentum balance equation . . . . . . . . . . . . . . . 114
6.4.3 Energy conservation . . . . . . . . . . . . . . . . . . . . 116
6.4.4 Internal energy balance equation . . . . . . . . . . . . . 118
6.5 Macroscopic equations . . . . . . . . . . . . . . . . . . . . . . . 119
6.5.1 Mass conservation equation . . . . . . . . . . . . . . . . 119
6.5.2 Momentum balance equation . . . . . . . . . . . . . . . 119
6.5.3 Internal energy balance equation . . . . . . . . . . . . . 119
6.6 Summary of the necessary conditions for retrieving the macro-
scopic transport equations . . . . . . . . . . . . . . . . . . . . . 120
6.7 The viscous stress tensor . . . . . . . . . . . . . . . . . . . . . . 121
6.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
6.9 LBE with a forcing term . . . . . . . . . . . . . . . . . . . . . . 123
CONTENTS 7
9
10 LIST OF FIGURES
7.3 Hermitian basis of the Hilbert subspaces Hq=6 in (a) and Hq=9
in (b), related to the D2Q9 LBE. . . . . . . . . . . . . . . . . . 138
7.8 Initial velocity proles used for the simulation of the double pe-
riodic shear ow. . . . . . . . . . . . . . . . . . . . . . . . . . . 147
7.10 Vorticity eld for t∗ = 1 showing the simulation results for the
LBGK model (left) and the moments-based method (right). It
is clearly seen that the LBGK model is plagued by spurious
secondary vortex even for Re = 32000 whereas the moments-
based scheme produces accurate and stable results up to Re =
80000 [66] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
7.11 Third-order Hermitian basis of the subspace Hq=10 and its re-
lated D2V17 set of lattice vectors. . . . . . . . . . . . . . . . . 155
8.4 Numerical results for Re=400, 1000 and 5000 showing the rel-
ative velocity u∗x /u∗0 along the mid-line y ∗ = L/2. Results are
compared with Ghia et al. [30], Botella and Peyret [8] and Er-
turk, Corke and Gökçöl [25]. Continuous lines are from Bazarin
et al. [ ? ]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
12 LIST OF FIGURES
8.5 Numerical results for Re=400, 1000 and 5000 showing the rel-
ative velocity u∗y /u∗0 along the mid-line x∗ = L/2. Results are
compared with Ghia et al. [30], Botella and Peyret [8] and Er-
turk, Corke and Gökçöl [25]. Continuous lines are from Bazarin
et al. [ ? ]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.6 Stream function and vorticity isolines for (a) Re = 100, (b) Re
= 400, (c) Re = 1000, (d) Re = 3200 and (e) Re = 5000. The
top row illustrate the stream function and the bottom row the
vorticity, Bazarin et al. [ ? ]. . . . . . . . . . . . . . . . . . . . . 171
8.7 Average density in terms of the number of time steps. We com-
pare the results of Latt et al. (2008), Malaspinas et al. (2011)
and Mohammed and Reis (2017) with the solution methods pre-
sented by Bazarin et al. (2021). Both Malaspinas et al. and Latt
et al. models fail in preserving the total mass of the system.
Method IV based on the concept of density is neither a mass
preserving method, but mass leaking is much smaller, Bazarin
et al. [ ? ]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
8.8 Average density measured with a more accurate weigth balance
in terms of the number of time steps. method IV based on den-
sity, presents the same linear behavior as observed in Malaspinas
et al. (2011) and Latt et al. (2008) models, although with a
much less mass variation, Bazarin et al. [ ? ]. . . . . . . . . . . 173
8.9 Local mass leakage for Mohammed and Reis (2017) model after
50000 time steps: (a) north and south boundaries; (b) east and
west boundaries; . . . . . . . . . . . . . . . . . . . . . . . . . . 174
List of Tables
13
14 LIST OF TABLES
Chapter 1
Introduction
d2 xi X
m = Fij , i = 1, ..., n (1.1)
dt2
j6=i
the position xi and velocity ξi = dxi /dt of each particle may be followed
along the time by solving the above system of equations. Two conditions are,
nevertheless, required to be satised for this solution:
2. the initial position and velocity must be known for each particle.
Intermolecular forces
Molecules attract themselves with an electrostatic force that is dependent
on their polar moment. Asymmetric molecules such as H2 0 and CO2 have
a permanent dipole, the geometric center of the electronic clouds does not
coinciding with the geometric center of the positive charges.
The dipole length and the dipolar moment of a polar molecule are subjected
to uctuations produced by electrostatic elds originated from other molecules.
15
16 CHAPTER 1. INTRODUCTION
In the same way, a non-polar molecule will acquire a dipolar moment, µind , un-
der the induction of an electrical eld, E. We dene the molecular polarizibility,
α, by
µind = αE,
giving a measure of the easiness a molecule becomes polarized, or changes its
dipolar moment, when subjected to an electrical eld.
Symmetrical molecules do not have any permanent electrical dipole and the
attractive forces among these molecules was rst described by London in 1930
as due to very rapid uctuations of the geometric center of their molecular
electronic clouds.
All these attractive forces between neutral molecules can be writen in terms
of a two-particles r−6 potential energy given by
B
Φ12 = − ,
r6
where B is a molecular property dependent on the electrostatic properties of
the interacting molecules 1 and 2 and r = |r2 − r1 |.
When molecules are close enough they experience a repulsion force due to
the overlapping of their electronic orbitals. The combined attraction-repulsion
eect can be described by the 12-6 Lennard-Jones potential [61] shown in Figure
1.1.
A B
Φ12 = − 6
r12 r
In molecular dynamics simulations it is usual to replace parameters A and
B by
Initial conditions
Molecular dynamics is a deterministic science, meaning that the initial
state
xi , ξi , i = 1, ..., n
is considered to be known at time t = 0.
Simulation of uid ow is performed by taking the ensemble average of
several realizations, starting from dierent initial conditions.
site is zero. So, if we want to simulate a collision process, we must preserve the
mass
X
mni = ρ, (1.2)
i
and the momentum
X
mni ei = ρu. (1.3)
i
Thus, we must choose, between the possible post-collisional congurations,
the ones that satisfy Eqs. (1.2-1.3). In this case, there are only two post-
collisional congurations that satisfy these restrictions. Figure 1.4.b show one
of such congurations.
Boolean models or `Lattice Gas Automatas (LGA)' as they were named
represent a drastic simplication of the molecular dynamics where:
2. the collision process is only required to preserve the mass and momentum
of the particles.
Nevertheless, they are able to produce a sketch of uid ow and represented a
new paradigm in CFD methods for hydrodynamics (Figure 1.5).
They were invented by Hardy, Pomeau and de Pazzis in 1973 [37], [36] who
adopted a square lattice with 4 neighbors for each site, the HPP lattice.
The HPP Boolean model was criticized by Frish, Hasslacher and Pomeau
[26]. In the words of these authors: When density and momentum are varied
slowly in space and time, "macrodynamical" equations dier from the nonlinear
Navier-Stokes equations in three respects. The discrepancies may be classied
as (1) lack of Galilean invariance, (2) lack of isotropy, and (3) a crossover
dimension problem .
20 CHAPTER 1. INTRODUCTION
In fact, the HPP does not have enough symmetry properties to insure the
isotropy of fourth rank lattice tensors and, so, fails to relate the viscous stress
tensor to the velocity gradients.
In this same paper, Ref. [26], Frish, Hasslacher and Pomeau proposed a
new LGAthe FHP Boolean model based on the triangular lattice shown
in Figure 1.3, avoiding the HPP drawbacks.
The algorithm
In the collision step the Boolean variables ni (x, t) related to site x are
modied in accordance with
α : S × S → {0, 1} .
Figure 1.6 shows the transition matrix for the HPP model. There are 16
possible states. The only transition that is allowed in the HPP model is the
one that changes the state (0, 1, 0, 1) to state (1, 0, 1, 0) and vice-versa.
ωi (n1 , ..., nb )
b
s
XX Y 1−sj
= α(s, s0) (s0i − si ) nj j (1 − nj ) . (1.5)
s s0 j=1
b
s
Y 1−sj
nj j (1 − nj ) 6= 0,
j=1
is the state s = (0, 1, 0, 1) and since α(s, s0) is diferent from zero only when
s0 = (1, 0, 1, 0), we will get
ω1 = 1, ω2 = −1, ω3 = 1, ω4 = −1.
and these values changes the state s = (0, 1, 0, 1) into the state s0 = (1, 0, 1, 0).
In the propagation step the Boolean variables are propagated to the neighbor
sites.
In FHP model there are 64 possible states and more than one state s0 for
some states s with two or three particles. In this case, we use a random variable
to decide which post-collisional Boolean state is to be chosen.
Finally, in some later versions of the FHP model a rest particle was added
and the Boolean state was written as
(n0 , n1 , ..., nb )
increasing the number of possible states to 128 and with benecial eects for
the simulations.
description is obtained with a set of kinetic equations for the expected values
of Ni = hni (x, t)i, obtained over a set of realizations. However, even this
description is redundant for many purposes. Thus, on a third level we look for
a closed system of equations involving the moments of Ni leading, e.g., under
certain hypotheses, to the Navier-Stokes equations.
So, consider several realizations of a Boolean model and the ensemble aver-
age of Eq. (1.4)
where
Ωi (N1 , ..., Nb )
b
s
XX Y 1−sj
= A(s, s0) (s0i − si ) Nj j (1 − Nj ) . (1.7)
s s0 j=1
X
A(s, s0) = 1.
s0
b
s
Y 1−sj
Nj j (1 − Nj ) ,
j=1
is now diferent from 0 or 1 and can be seen as the probability P (s) of Boolean
state s = (s1 , ..., sb ) considering all the dierent realizations. In the same way
b
Y s0 1−s0j
P (s0 ) = Nj j (1 − Nj ) ,
j=1
In equilibrium, when state s can change to state s0 , i.e., when they have the
same mass and momentum, their probability must be identical
P (s) = P (s0 ) .
So
b b
Y s 1−sj
Y s0 1−s0j
Nj j (1 − Nj ) = Nj j (1 − Nj ) .
j=1 j=1
b b b
Y s 1−sj
X Nj X
ln Nj j (1 − Nj ) = sj ln + ln (1 − Nj ) ,
j=1 j=1
1 − Nj j=1
b b b
Y s0 1−s0j
X Nj X
ln Nj j (1 − Nj ) = s0j ln + ln (1 − Nj ) ,
j=1 j=1
1 − Nj j=1
and so
b b
X Nj X Nj
sj ln = s0j ln ,
j=1
1 − Nj j=1
1 − Nj
We know that s and s0 must satisfy the mass and momentum preservation
conditions
b
X b
X b
X b
X
sj = s0j and sj ej = s0j ej ,
j=1 j=1 j=1 j=1
Nevertheless, sj and s0j can have dierent values (0 or 1) for the same j. So
this identity is only satised if
Nj
ln ,
1 − Nj
is a collisional invariant.
Following this line of reasoning, Chen et al. [16] suggested replacing the
collision term by a single relaxation-time term, followed by Qian et al. [79] and
Chen et al. [13], who introduced a model based on the celebrated kinetic-theory
idea of Bhatnagar, Gross, and Krook (BGK) [4], but adding rest particles and
retrieving the correct incompressible Navier-Stokes equations, with third-order
non-physical terms in the local Mach number.
The single relaxation time BGK collision term describes the relaxation of the
distribution function to an equilibrium distribution. This discrete equilibrium
distribution was settled by writing it as a second-order polynomial expansion
in the particle velocity, with parameters adjusted to retrieve the mass density,
the local velocity, and the momentum ux equilibrium moments, which are
necessary conditions for satisfying the Navier-Stokes equations.
Nowadays, LBM subjects range from the fundamentals of the kinetic the-
ory and quantum transport to applied subjects, including some ones such as
such as: magnetohydrodynamics (MHD); quantum lattice Boltzmann mod-
els; unstructured lattice Boltzmann equation (ULBE); non-Newtonian ows;
molecular dynamics. They also include lattice Boltzmann approaches for the
analysis of precursor lms; droplet spreading on solid surfaces; biopolymer
translocation; nano and microuidics; bioengineering and biophysics; granular
ows; combustion; hybrid methods and high performance computing.
26 CHAPTER 1. INTRODUCTION
Chapter 2
2.1 Introduction
The concept of `heat' or `internal energy' of a thermodynamic system as
the result of the kinetic energy of its individual molecules was long advocated
by philosophers and scientists such as Francis Bacon (1561-1626), John Locke
(1632-1704) and Thomas Hooke (1635-1703). Heat is motion in the words
of Francis Bacon (around 1600). Heat is nothing else but the motion of the
particles that form a body (Hooke, around 1650).
27
28 CHAPTER 2. THE BOLTZMANN EQUATION
δQ = dU
where U is the internal energy of a thermodynamic system, i.e., the energy of
its molecules.
On the mechanics side, people were worried with the movement of the plan-
ets and moons, i.e., with systems for which the mechanical energy is preserved.
Nowadays, the conservation of mechanical energy would be written as
δW = dEc + dEp
The rst half of the 19th century saw the birth of thermodynamics as a uni-
fying theory and the law of energy conservation was postulated by Helmholtz,
in 1847 [43],
δW + δQ = dU + dEc + dEp
2.1. INTRODUCTION 29
The Joule's principle of the mechanical equivalent of the caloric [49] can be
derived by integrating this equation, when a thermodynamic system performs
a cycle, returning to its initial state
(δW + δQ) = 0
James Joule equivalence principle and the XIX Century mechanical theory
of heat were highly inuenced by the atomistic philosophy.
Nevertheless, the main question that imposed severe diculties for the ac-
ceptance of the heat and work as equivalent forms of energy was only solved
by Clausius, a German scientist [18], [20].
Indeed, in the words of William Thompson Nature has an arrow, and the
natural tendency is to transform vis-viva (or energy) into heat and not the
inverse .
So how to advocate the Joule's equivalence principle?
Clausius postulated this as the second law of thermodynamics, showing that
if this was possible, heat would ow from a colder to a hotter body .
These two principles are stated in his 1867 book [20]: " Die Energie der
Welt ist constant. Die Entropie der Welt strebt einem Maximum zu."
But how to conciliate the dissipative second law of thermodynamics with
the conservative laws of mechanics? Indeed, in an isolated universe of particles
satisfying the Newton's laws, the mechanical energy is preserved and this sys-
tem is reversible: its previous states can always be attained by reversing the
time in the equations of motion.
So, if `heat ' has a `mechanical foundation ' how to explain irreversibility
based on the mechanical laws of motion?
1
Joule was tutored as a young man by the famous scientist John Dalton and
2
was strongly inuenced by chemist William Henry . His equivalence principle
was based on the atomistic philosophy and he also explained the pressure of
gases by the impact of their molecules, and has calculated the velocity which
they must have in order to produce the pressure observed in particular gases.
Maxwell [67].
In accordance with Maxwell [67], Clausius paper On the kind of motion
3
we call heat is a complete exposition of the molecular theory adopted in this
paper. After reading his investigation of the distance described by each molecule
1 The most important of all Dalton's investigations are those concerned with the beginning
atomic theory in chemistry. The main points of Dalton's atomic theory were: i) elements are
made of extremely small particles called atoms; ii) atoms of a given element are identical in
size, mass, and other properties; atoms of dierent elements dier in size, mass, and other
properties; iii) atoms cannot be subdivided, created, or destroyed; iv) atoms of dierent
elements combine in simple whole-number ratios to form chemical compounds; v) in chemical
reactions, atoms are combined, separated, or rearranged.
2 Henry W (1803) Experiments on the Quantity of Gases Absorbed by Water, at Dierent
Temperatures, and under Dierent Pressures. Philos Trans R Soc London 93:29274. doi:
10.1098/rstl.1803.0004 [44].
3 Originally published under the title Ueber die Art der Bewegung, welche wir Winne
nennen , Annalen der Physik, Vol. 100, pp. 353-80 (1857); English translation in Philosoph-
ical Magazine, Vol. 14, pp. 108-27 (1857).
30 CHAPTER 2. THE BOLTZMANN EQUATION
Figure 2.3: Considering all the molecules inside a parallelepiped of sides ∆x,
∆y and ∆z, only f (x, ξ, t)∆x∆y∆z∆ξx ∆ξy ∆ξz have velocities between ξ and
ξ + ∆ξ .
assume on the basis of the kinetic theory of gases that the perfectly elastic
balls representing the molecules always tended to change their actual posi-
tions. Consequently, any conguration, however improbable, could conceivably
occur as time went on: "The calculus of probabilities teaches us precisely this:
any non-uniform distribution, unlikely as it may be, is not strictly speaking
impossible."
It now remained to give to this conclusion a rigorous derivation and quan-
titative applicability, which Boltzmann did in his memoir submitted to the
academy on October, 1877, "On the Relation between the Second Law of the
Mechanical Theory of Heat and the Probability Calculus with respect to the
Propositions about Heat-Equivalence ". These memoirs were published in Ger-
man in 1896 and 1898 [7]. He concluded that the entropy of a state is propor-
tional to the probability of the conguration of its component particles. In this
way, entropy became a purely statistical attribute of the overall system. As the
entropy increases, the conguration of the system becomes more probable.
+
Nevertheless, there are (∂t f ) ∆x∆ξ∆t particles in the volume ∆x that
acquired the velocity ξ in the course of the time interval ∆t because they have
−
collided with other particles. On the other hand, (∂t f ) ∆x∆ξ∆t particles in
∆x, loss, in this time interval, the velocity ξ due also to the collisions they
suered with other particles.
So
+ −
∂t f + ξ · (∇x f ) + g · (∇ξ f ) = (∂t f ) − (∂t f ) . (2.4)
| {z } | {z } | {z }
stream term external forces colision term
In Boltzmann's model, the particles behave like billard balls, each collision
producing a sharp change in their velocity. A molecule may be thought as a
material point with a force eld around it. This force eld has a strong repulsion
kernel. When two molecules have a frontal collision, they will experience this
strong repulsion when their electronic orbits begin to intercept with themselves.
Around this repulsion kernel, molecules have a soft attraction eld produced
by electrostatic forces.
Considering σ to be a length related to a parameter possible to be identied
with the molecular diameter, the two particles will experience a repulsion force
when their centers are at a distance equal to σ. At this distance the attraction
and repulsion elds canceal themselves but the kinetic energy of the bullet with
respect to the target particle is not null and the center of the bullet particle will
penetrate into the σ -sphere, where it will be frained along the radial coordinate.
This radial frainage will produce a deviation in the bullet trajectory de-
picted in Figure (2.4). Using the label 1 for the bullet particle, the relative
velocity of the bullet with respect to the target before the collision is γ = ξ1 − ξ
when the target and bullet have the velocities ξ and ξ1 , respectively.
The collision term Ω takes account of these losses and gains in the particle
populations due to radical changes of the velocity ξ happening in a very small
time interval.
Considering that our reference is on a single target particle with velocity ξ,
this particle will change its velocity to ξ0 if it is hit by a bullet particle with
velocity ξ1 during the time interval ∆t.
2.2. THE BOLTZMANN EQUATION 33
Figure 2.4: Trajectory deviation of the bullet particle produced by the radial
frainage due to the repulsion forces. Vector
→
−
α is a unitary vector along γ −γ
0
.
Parameter b is the impact factor and angle ε is the azimuth angle of the collision
plane.
bullet particles.
Therefore, since there are f (x, ξ1 , t) ∆ξ∆x target particles inside ∆x, dur-
ing the time interval ∆t, the number of target particles that loss the velocity
ξ, during ∆t will be,
−
(∂t f ) ∆tdξ∆x
2π σ
= f (x, ξ, t) f (x, ξ1 , t) bdεdbγdξ1 ∆t∆ξ∆x. (2.7)
ξ1 ε=0 b=0
34 CHAPTER 2. THE BOLTZMANN EQUATION
considering all the possible velocities ξ1 in the velocity space and all the impact
parameters b in each azimuthal plane ε,
So
2π σ
−
(∂t f ) = f (x, ξ, t) f (x, ξ1 , t) bdεdbγdξ1 (2.8)
ξ1 ε=0 b=0
In the same way,
2π σ
+
(∂t f ) = f (x, ξ 0 , t) f (x, ξ10 , t) b0 dεdb0 γ 0 dξ1 (2.9)
ξ1 ε=0 b=0
where ξ 0 , ξ10 must be such that ξ 0 , ξ10 → ξ, ξ1 after the collision. When the
potential energy is originated by a central eld, i.e., when this potential energy
is dependent only on the relative distance between two molecules Φ = Φ (krk),
0
it can be shown that γ 0 = γ and b = b. It is also possible ξ 0 , ξ10
to calculate
from known values of ξ, ξ1 using the mechanics of Newton.
In its nal form, the Boltzmann equation is written as,
∂t f + ξ · (∇x f ) + g · (∇ξ f )
2π σ
f (x, ξ 0 , t) f (x, ξ10 , t)
= bdεdbγdξ1 (2.10)
ξ1 ε=0 b=0 −f (x, ξ, t) f (x, ξ1 , t)
The acceleration g is given by
∂t f + ξ · (∇x f ) + g(e) + g(`d) · (∇ξ f ) = Ω, (2.12)
2π σ
f (x, ξ 0 , t) f (x, ξ10 , t)
Ω= bdεdbgdξ1 , (2.13)
ξ1 ε=0 b=0 −f (x, ξ, t) f (x, ξ1 , t)
gives the net balance between the molecules that acquire the velocity ξ and the
ones that loss this velocity. In equilibrum conditions this balance must be null,
otherwise the distribution f would vary in the course of the time and in the
space. So,
0
f x, ξ , t f (x, ξ10 , t) = f (x, ξ, t) f (x, ξ1 , t) , (2.14)
or
meaning that
m = mass, (2.17)
mξ = momentum, (2.18)
1 2
mξ = kinetic energy. (2.19)
2
Therefore, any linear combination of these invariants is also a collisional
invariant,
ln f (eq) = A + B · ξ + Cξ 2
2
= a + b (ξ − c) , (2.20)
or
2 2
f (eq) = ea eb(ξ−c) = deb(ξ−c) . (2.21)
n = f (eq) dξ = number density of molecules, (2.22)
1
u = f (eq) ξdξ = mean molecular velocity, (2.23)
n
1 1 2
ec,f = f (eq) m (ξ − u) dξ = mean peculiar kinetic energy.(2.24)
n 2
In fact, for a uid in equilibrium the quantities
n= f dξ, (2.25)
1
u = hξi = f ξdξ, (2.26)
n
must be constants and given by replacing the distribution f by the distribution
the system has at equilibrium, i.e., f (eq) .
The third integral deserves a more lengthy discussion. The mean kinetic
energy per molecule is given by
1
1 1
ec = m ξ 2 = f mξ 2 dξ. (2.27)
2 n 2
Writting,
2
ξ 2 = (ξ − u) + 2 (ξ − u) · u + u2 , (2.28)
it can be easily seen that this kinetic energy has two components
J 1 1 2 1
ec = f m (ξ − u) dξ + mu2 . (2.29)
molecule n 2 2
| {z }
| {z }
=ec,f advection energy
The thermodynamic internal energy, e, per molecule, is the sum of the the
peculiar kinetic energy ec,f , i.e., the energy due to the molecular random motion
of the molecules and the potential energy due to the intermolecular interaction
among these molecules,
Φ
e = ec,f + , (2.30)
n
where Φ is the intermolecular potential energy per unit volume.
For thermodynamic systems in equilibrium, the thermodynamic internal
energy, satises,
∂P
de = cv dT + T − P dv . (2.31)
| {z } ∂T v
energy due to molecular motion | {z }
intermolecular potential energy Φ
2.3. THE EQUILIBRIM SOLUTION 37
kT
P = nkT = , (2.32)
v
where,
R
k= , (2.33)
N
9
N being the Avogadro number and k= Boltzmann constant. So, the second
term in Eq. (2.31) is null and,
de = cv dT, (2.34)
kT a
P = − , (2.35)
v − b v2
the parameter ”a” is related to the intermolecular forces and we get
a
de = cv dT + dv. (2.36)
v2
In both cases, the molal heat cv is a molecular parameter related to the
10
motion degrees of freedom of the molecules ,
9 The Avogadro constant is named after the early nineteenth century Italian scientist
Amedeo Avogadro, who, in 1811, rst proposed that the volume of a gas (at a given pressure
and temperature) is proportional to the number of atoms or molecules regardless of the
nature of the gas. The French physicist Jean Perrin in 1909 proposed naming the constant
in honor of Avogadro. Perrin won the 1926 Nobel Prize in Physics, in a large part for his
work in determining the Avogadro constant by several dierent methods.
The value of the Avogadro constant was rst indicated by Johann Josef Loschmidt who, in
1865, estimated the average diameter of the molecules in air by a method that is equivalent
to calculating the number of particles in a given volume of gas. This latter value, the number
density of particles in an ideal gas, is now called the Loschmidt constant in his honour, and
is approximately proportional to the Avogadro constant. The connection with Loschmidt is
the root of the symbol L sometimes used for the Avogadro constant, and German language
literature may refer to both constants by the same name, distinguished only by the units of
measurement.
Accurate determinations of Avogadro's number require the measurement of a single quan-
tity on both the atomic and macroscopic scales using the same unit of measurement. This
became possible for the rst time when American physicist Robert Millikan measured the
charge on an electron in 1910. The charge of a mole of electrons is the constant called the
Faraday and had been known since 1834 when Michael Faraday published his works on elec-
trolysis. By dividing the charge on a mole of electrons by the charge on a single electron the
value of Avogadro's number is obtained. Since 1910, newer calculations have more accurately
determined the values for Faraday's constant and the elementary charge. (from Wikipedia )
10 With monatomic gases, thermal energy comprises only translational motions. Trans-
lational motions are ordinary, whole-body movements in 3D space whereby particles move
38 CHAPTER 2. THE BOLTZMANN EQUATION
D
cv = cv,transl = k, (2.38)
2
where D= space dimension.
Therefore, by integrating Eq. (2.36), we get for a van der Waals equation
of state,
D 1
e (T, v) = kT + −a (in J/molecule), (2.39)
2 v
plus a constant C (T0 , v0 ) that depends only on the initial state.
This means that, in a molecular basis,
1 1 2 D
ec,f = f (eq) m (ξ − u) dξ = kT (in J/molecule). (2.40)
n 2 2
about and exchange energy in collisionslike rubber balls in a vigorously shaken container
(see animation here). These simple movements in the three X, Y, and Zaxis dimensions of
space means individual atoms have three translational degrees of freedom. A degree of free-
dom is any form of energy in which heat transferred into an object can be stored. This can
be in translational kinetic energy, rotational kinetic energy, or other forms such as potential
energy in vibrational modes. Only three translational degrees of freedom (corresponding to
the three independent directions in space) are available for any individual atom, whether it
is free, as a monatomic molecule, or bound into a polyatomic molecule.
As to rotation about an atom's axis (again, whether the atom is bound or free), its energy
of rotation is proportional to the moment of inertia for the atom, which is extremely small
compared to moments of inertia of collections of atoms. This is because almost all of the
mass of a single atom is concentrated in its nucleus, which has a radius too small to give a
signicant moment of inertia. In contrast, the spacing of quantum energy levels for a rotating
object is inversely proportional to its moment of inertia, and so this spacing becomes very
large for objects with very small moments of inertia. For these reasons, the contribution
from rotation of atoms on their axes is essentially zero in monatomic gases, because the
energy-spacing of the associated quantum levels is too large for signicant thermal energy
to be stored in rotation of systems such small moments of inertia. For similar reasons, axial
rotation around bonds joining atoms in diatomic gases (or along the linear axis in a linear
molecule of any length) can also be neglected as a possible "degree of freedom" as well, since
such rotation is similar to rotation of monatomic atoms, and so occurs about an axis with a
moment of inertia too small to be able to store signicant heat energy.
In polyatomic molecules, other rotational modes may become active, due to the much
higher moments of inertia about certain axes which do not coincide with the linear axis of
a linear molecule. These modes take the place of some translational degrees of freedom for
individual atoms, since the atoms are moving in 3-D space, as the molecule rotates. The
narrowing of quantum mechanically-determined energy spacing between rotational states
results from situations where atoms are rotating around an axis that does not connect them,
and thus form an assembly that has a large moment of inertia. This small dierence between
energy states allows the kinetic energy of this type of rotational motion to store heat energy
at ambient temperatures. Furthermore (although usually at higher temperatures than are
able to store heat in rotational motion) internal vibrational degrees of freedom also may
become active (these are also a type of translation, as seen from the view of each atom). In
summary, molecules are complex objects with a population of atoms that may move about
within the molecule in a number of dierent ways (see animation at right), and each of these
ways of moving is capable of storing energy if the temperature is sucient. (from Wikipedia)
2.4. EXERCISES 39
D
e (T, n) = kT + (−an) (in J/molecule). (2.41)
2
Eq. (2.40) is, in fact, the denition of temperature for systems composed
of molecules with only translational degrees of freedom. For a van der Waal
equation of state, the intermolecular potential energy per unit volume becomes
Φ = −an2 and is always negative. When the number density of molecules
increases, meaning a larger number of molecules in the same volume, this energy
increases in absolute value, meaning that the molecules are strongly linked,
When the number density decreases, Φ → 0. For an ideal gas, the molecules
are free from the intermolecular forces and Φ = 0.
With these restrictions Eqs (2.22-2.24) give for the equilibrium distribution,
the Maxwell-Boltzmann distribution
m D/2 − (ξ−u)2
2kT
f (eq) = n e m . (2.42)
2πkT
2.4 Exercises
1- Consider a two-dimensional system of particles and the transformation
ξx , ξy → ξf x = ξx − ux , ξf y = ξy − uy . We know that dξx dξy = Jdξf x dξf y
where J is the Jacobian of the transformation
∂ξ ∂ξx
x
∂ (ξx , ξy )
∂ξf x ∂ξf y
J= = ∂ξy ∂ξy
∂ (ξf x , ξf y ) ∂ξ ∂ξ
fx fy
Show that J = 1. ( Hint: remember that the molecular velocity and the
position are independent variables )
2
ξf
f eq = Ae− B ,
where A and B can be determined by considering that the number density of
molecules is given by
∞ ∞ ∞ 2 +ξ2 +ξ2
ξf,x f,y f,z
n= eq
f dξ = A e− B dξx dξy dξz ,
−∞ −∞ −∞
and that the kinetic energy related to the `molecular agitation' is given by
40 CHAPTER 2. THE BOLTZMANN EQUATION
1
ec,f = f eq mξf2 dξ
2
∞ ∞ ∞ 2 +ξ2 +ξ2
ξf,x f,y f,z 1
Ae− 2 2 2
= B m ξf,x + ξf,y + ξf,z dξx dξy dξz .
−∞ −∞ −∞ 2
FindA and B in terms of the number density n and the mean kinetic energy
of uctationsec,f . What are the independent variables of n and ec,f ?
Hint. Dene new integration variables such as ξ√f,x
B
ξ
= x, √f,y
B
ξ
= y, √f,z
B
=z
and remember that
∞ ∞
2 √ 2 1√
e−x dx = π, e−x x2 dx = π.
−∞ −∞ 2
2 2
f (eq) = ea e−b(ξ−c) = de−b(ξ−c) .
Find d, b and c using
n = f (eq) dξ = number density of molecules,
1
u = f (eq) ξdξ = mean molecular velocity,
n
1 1 2
ec,f = f (eq) m (ξ − u) dξ = mean peculiar kinetic energy.
n 2
N= n (x, t) dV,
V (t)
dN d
= ndV = 0,
dt dt V (t)
d ∂n
ndV = dV + nu · en dA
dt V (t) V (t) ∂t A(t)
where A (t) is the area of the surface around V(t) and en is the unitary vector
pointing outside A (t), to show that
2.4. EXERCISES 41
∂ρ
+ ∇ · (ρu) = 0 (2.43)
∂t
where ρ = nm and m is the mass of each molecule.
∂ρ
+ u · ∇ρ = −ρ∇ · u
∂t
Consider now a Taylor expansion of ρ (x + ∆x, t + ∆t) around ρ (x, t)
∂ρ
ρ (x + ∆x, t + ∆t) = ρ (x, t) + ∆t + ∇ρ · ∆x + ...
∂t
In the limit ∆t → 0 we get
Pαβ = P δαβ
and reduces to the thermodynamic pressure P = nkT (the Clapeyron equation ).
b) Under the light of Eq. (2.44) try to describe the meaning of an isotropic
tensor, using your own words.
c) Consider a 2D case and the integral
∞ ∞
f eq mξf x ξy dξx dξy
ξy =0 ξfx =−∞
42 CHAPTER 2. THE BOLTZMANN EQUATION
giving the momentum mξf x that is transported into the positive direction ξy
and the integral
0 ∞
f eq mξf x ξy dξx dξy
ξy =−∞ ξfx =−∞
giving the momentum mξf x that is transported into the negative direction ξy
and answer the question:
In equilibrium, the momentum mξf x is transferred along direction y ?...Justify
your answer.
Chapter 3
∂t f + ξ · (∇x f ) = Ω, (3.1)
2π σ
Ω(sd) = (f 0 f10 − f f1 ) bdεdbgdξ1 .
ξ1 ε=0 b=0
(ξ − u)
∇ξ f = ∇ξ f eq = − kT
f eq .
m
So we get a kinetic model for Ω
(ξ − u)
Ω = Ω(sd) + g(e) + g(`d) · kT f eq .
m
By `kinetic model' we mean that the Boltzmann equation is replaced by a
model that is able to retrieve the main, or some of the main properties of the
1 Readers well acquainted with the rudiments of the LB method can avoid this lecture.
43
44 CHAPTER 3. A FIRST INSIGHT INTO LBM
Boltzmann equation. In LBM, the more widely used model for Ω(sd) is the
2
BGK model [4]
f eq − f
Ω(sd) = , (3.3)
τ
meaning that the Boltzmann collision term is replaced by a relaxation term
with a single relaxation time τ. Therefore, this model preserves the main eect
of collisions leading the distribution of particles towards equilibrium, when a
uid is in a non-equilibrium state.
We will talk more about kinetic models in the following lectures. In this
lecture we focus our attention on LBM discretization and consider Ω as a known
function Ω (f, f eq ).
Discretization means to replace the entire continuous physical space, repre-
sented by the continuous variable x by some points xi and the entire velocity
Figure 3.2: D2Q9 [79] and RD2Q9 [74] sets of lattice vectors.
∂t fi + ξi · (∇x fi ) = Ωi ,
where fi = f (x, ξi , t) indicates the value of f for ξ = ξi , i.e., the packet of
particles with velocity ξi that are found in the point x at time t.
We further consider that the uid is free from external (no gravity ) and
intermolecular forces (ideal gas assumption ), or that
fieq − fi
Ωi =
τ
At the time t + δ, the population fi of particles along direction i can be
found using a Taylor series
∂fi 1 ∂ 2 fi
fi (x, t + δ) = fi (x, t) + δ + δ 2 2 + ...,
∂t 2 ∂t
and will be known from the value this packet had at time t, when all the time
∂fi ∂ 2 fi
derivatives
∂t , ∂t2 , ...are known at time t.
This is not the case. Nevertheless, when δ → 0, it is possible to neglect all
the derivatives of order 2 and higher.
∂fi
+ O δ2
fi (x, t + δ) = fi (x, t) + δ
∂t
giving
∂fi ∂fi
2
fi (x + ∆x, t + δ) = fi (x, t) + δ + ∆x · + O (∆x) , δ 2 .
∂t ∂x
Now consider that the physical and velocity space are coupled through
∆x = ξi δ,
meaning that after each time step δ , the population fi of particles is transferred
from site x to site x + ∆x This is the main feature of almost all LB methods.
In this case,
∂fi ∂fi
+ O δ2 ,
fi (x + ξi δ, t + δ) = fi (x, t) + δ + ξi ·
∂t ∂x
or
∂fi ∂fi fi (x + ξi δ, t + δ) − fi (x, t)
+ ξi · = + O (δ) .
∂t ∂x δ
The discrete Boltzmann equation will, then, be written as
fieq − fi
+ O δ2 ,
fi (x + ξi δ, t + δ) = fi (x, t) + ∗
(3.4)
τ
where τ∗ is a dimensionless relaxation time
∗
τ = τ /δ .
The term
|ξi | δ
,
|∆x|
3.1. FROM THE BOLTZMANN EQUATION TO LB SCHEMES 47
3
is the Courant-Friedrich-Lewy number , CF L, and in the present numerical
scheme CF L = 1. This condition implies that at the time t + δ the particles
with velocity ξi will be found at the site x + ∆x , at the time t + δ . This scheme
is at the origin of the collision propagation LB method, which is an explicit
numerical method because the interacton term Ωi is evaluated at time t.
In LB simulation, Eq. (3.4) is usually replaced by
eq
fi − fi
+ O δ2 ,
f i (x + ei h, t + δ) = f i (x, t) + ∗
(3.5)
τ
where
3
fi ξ
fi = ,
n0
is a dimensionless population, n0 being a number density of reference and
r
kT
ξ= ,
m
4
is related to a mean molecular speed .
Now, we need to relate the molecular velocity ξi with the lattice vectors ei
and the Maxwell-Boltzmann equilibrium distribution f eq (ξi ) with the discrete
eq
equilibrium populations fi .
A fundamental quantity in LBM is the lattice speed c dened as
h
c= ,
δ
h being the orthogonal distance between any two contiguous sites, meaning that
the particles will be transported along direction i with a velocity cei . Since
these particles have a molecular velocity ξi , we can conclude that the lattice
speed is such that
ξi = cei .
3 In mathematics, the CourantFriedrichsLewy condition (CFL condition) is a necessary
condition for convergence while solving certain partial dierential equations (usually hyper-
bolic PDEs) numerically (it is not in general a sucient condition). It arises when explicit
time-marching schemes are used for the numerical solution. As a consequence, the time
step must be less than a certain time in many explicit time-marching computer simulations,
otherwise the simulation will produce wildly incorrect results. The condition is named after
Richard Courant, Kurt Friedrichs, and Hans Lewy who described it in their 1928 paper [21].
√
r
kT
D .
m
48 CHAPTER 3. A FIRST INSIGHT INTO LBM
ξi
ξ0,i = ,
ξ
with the lattice vectors ei ,
ξ0,i = as ei .
The scaling factor as can be shown to be a lattice parameter and will have
dierent values when we change the discrete set of lattice vectors. It is more
usual to write it as
1
cs = ,
as
since cs can be shown to be the dimensionless speed of sound in isothermal ow
in LBM simulations.
From the above equations,
ξ
ξi = cei = ξξ0,i = ei
cs
so
r
kT
ξ= = ccs .
m
Now, we drive our attention to the relation between the equilibrium popu-
eq
lations fi and the Maxwell-Boltzmann equilibrium distribution f eq (ξ),
m D/2 − (ξ−u)2
eq 2kT
f (ξ) = n e m .
2πkT
Since the MB equilibrium distribution refers to an innite velocity space,
eq
we cannot require that the discrete populations fi reproduce this distribution.
eq
What it is possible to require [69] is that the rst moments of fi are the same
rst moments of f eq (ξ).
eq
So, the discrete populations fi are required to retrieve the number density
of molecules
X eq
n= f eq (ξ) dξ = n0 fi ,
i
or
n 1 X eq
ρ∗ = = f eq (ξ) dξ = fi . (3.6)
n0 n0 i
1 X eq
ρ∗ u∗ = f eq (ξ) ξdξ = f i ei . (3.7)
n0 c i
where u∗ = u/c
We also require that
1 X eq
ρ∗ u∗α u∗β ∗
+ P δαβ = f eq (ξ) mξα ξβ dξ = f i ei,α ei,β , (3.8)
mn0 c2 i
where
P n kT 1
P∗ = 2
= = ρ∗ c2s
mn0 c m c2
n0 |{z}
|{z}
=ρ∗ =c2 c2s
which is the dimensionless equation of state for ideal gases the Clapeyron
equation written in dimensionless variables.
The above conditions are all we need to satisfy when dealing with second
order models like the D2Q9 or D3Q27.
1 (ξ0 −u0 )2
f eq = n D
e− 2 ,
D/2
(2π) ξ
f − f eq
Ω=− , (3.9)
τ
where τ is a relaxation time.
eq
Now, the dimensionless form f of f eq may be developed in terms of the
Hermitian basis of H as
D
eq f eq ξ 1 (ξ0 −u0 )2
f = = ρ∗ D/2
e− 2
no (2π)
ξ2
− 20 (θ)
e X aeq,nx ,ny ,nz
= ρ∗ D/2
× Hn(θ)
x ,ny ,nz
(ξ0 ) ,
(2π) nx !ny !nz !
θ
H (0) = 1,
(1)
H1 = Hx(1) = ξ0,x ,
(1)
H2 = Hy(1) = ξ0,y (3.10)
(2) (2)
H1,1 = Hxx = ξ0,x ξ0,x − 1,
(2) (2)
H1,2 = Hxy = ξ0,x1 ξ0,y , (3.11)
(2) (2)
H2,2 = Hyy = ξ0,y ξ0,y − 1 (3.12)
(θ)
aeq(nx ,ny ,nz ) (u0 )
1 (ξ0 −u0 )2
= D/2
e− 2 Hn(θ)
x ,ny ,nz
(ξ0 ) dξ0 .
(2π)
In components notation, the rst equilibrium moments are
a(0)
eq = 1,
a(1)
eq,α = u0,α ,
(2)
aeq,αβ = u0,α u0,β ,
2
ξ0 " #
(0) (1) (1) (1) (1)
eq e− 2 aeq H (0) + aeq,x Hx + aeq,y Hy
f = D ρ∗ 1 (2) (2) 1 (2) (2) 1 (2) (2)
(2π) 2 + 2! aeq,xx Hxx + 2! aeq,yy Hyy + + 1!1! aeq,xy Hxy
2
ξ0
e−
1 + 1u0,x ξo,x + u0,yξo,y
2
∗
= ρ 1 2 2
(2π) 2
D
+ 2! u0,x ξo,x − 1 + 2! 2
u20,y ξo,y − 1 + u0,x u0,y (ξo,x ξo,y )
2
ξ0,i
e−
1 + 1u0,x ξ0,i,x + u0,yξo,i,y
eq 2
∗
fi = D ρ 1 2 2 2 2
(2π) 2 + u
2! 0,x ξ 0,i,x − 1 + 2! 0,y ξ0,i,y − 1 + u0,x u0,y (ξ0,i,x ξ0,i,y )
u
u∗ e u∗ ei,y
1 + xc2i,x + yc2
= W i ρ∗ ∗2
s
1 u∗y2 2
s
u∗x u∗y .
1 ux 2 2 2
+ 2 c4 ei,x − cs + 2 c4 ei,y − cs + c4 ei.x ei.y
s s s
u∗ e
" #
eq ∗ 1 + αc2iα +
fi = Wi ρ 1 ∗ ∗
s
2
, (3.13)
2c4 uα uβ eiα eiβ − cs δαβ
s
1 16 4 1
cs = √ , W0 = , W1 = , W2 = .
3 36 36 36
Since the particles behave as the molecules of an ideal gas as the long-
range intermolecular forces were suppressed the incompressibility condition
must be assured by keeping the Mach number as small as possible. The Mach
number is dened as the ratio between a uid characteristic speed U and the
speed of sound Us ,
U
Ma = ,
Us
which can also be written as
U∗
Ma = ,
cs
where, as before, U ∗ = U/c and cs = Us /c are dimensionless speeds. The LBM
simulation of incompressible ows with second order models requires M a ≤ 0.1.
X 1
ρ∗ = fi = f dξ
i
n0
X 1
ρ∗ u∗ = f i ei = f ξdξ
i
n0 c
52 CHAPTER 3. A FIRST INSIGHT INTO LBM
∗ ∗ 1
ν = c2s τ − .
2
On the other hand, since the Reynolds number is dened as
U ∗ (L/h) U ∗ (L/h)
Re = = ,
ν∗ c2s τ ∗ − 21
where L is the linear dimension of the physical domain. Since U∗ must be
kept as small as possible, if we want to simulate larger Re number ows, the
relaxation time τ∗ 1
2 as possible for reducing the
must be kept as close to
computational cost represented by an excessively linear number of sites N =
L/h.
Another limit in LBM simulation is represented by the Knudsen number.
In kinetic theory it is dened as
`
Kn =
L
where ` is the mean free path,
` = ξτ ∗ δ = cs τ ∗ h.
So, in LBM the Kn number is, usualy, dened as
cs ν ∗
Kn = .
N
3.3. SOME FEATURES OF THE D2Q9 LBE 53
c) Propagation step: after the collision step, the new values of the distri-
out
butions f i are propagated along the i-directions
out
f i (x + ei h, t + δ) = f i (x, t + δ) , (3.15)
The equilibrium distribution for the D2Q9 LBE is given by a second order
expansion in the orthogonal Hermite polynomials [84] H0 =1, H1,α = ξ0i,α =
eiα 1
cs , H2,αβ = ξ0i,α ξ0i,β − δαβ = c2s eiα eiβ − c2s δαβ giving
u∗ e
" #
eq 1 + αc2iα +
fi = W i ρ∗ 1 ∗ ∗
s
2
,
2c4 uα uβ eiα eiβ − cs δαβ
s
4 ∗ 3 ∗2 3 ∗2
f¯0eq = ρ 1 − uy − ux
9 2 2
1 ∗ 3 ∗2
f¯1eq = ∗ ∗2
ρ 3ux + 3ux − uy + 1
9 2
1 ∗ 3 ∗2
f¯2eq = ∗
ρ 3uy − ux + 3uy + 1 ∗2
9 2
1 ∗ 3 ∗2
f¯3eq = ∗2 ∗
ρ 3ux − 3ux − uy + 1
9 2
1 ∗ 3 ∗2
f¯4eq = ∗2
ρ 3uy − ux − 3uy + 1 ∗
9 2
1 ∗
f¯5eq ρ 3u∗x + 3u∗y + 9u∗x u∗y + 3u∗2 ∗2
= x + 3uy + 1
36
1 ∗
f¯6eq ρ 3u∗y − 3u∗x − 9u∗x u∗y + 3u∗2 ∗2
= x + 3uy + 1
36
1 ∗
f¯7eq ρ 9u∗x u∗y − 3u∗y − 3u∗x + 3u∗2 ∗2
= x + 3uy + 1
36
1 ∗
f¯8eq ρ 3u∗x − 3u∗y − 9u∗x u∗y + 3u∗2 ∗2
= x + 3uy + 1
36
It is easy to verify that the main conditions for the discretization are satised
up to second order velocity polynomials
8
X 1
f¯ieq = f¯eq dξ = ρ∗
i=0
n0
8
X 1
f¯ieq ei,x = f¯eq ξx dξ = ρ∗ u∗x
i=0
n0 c
8
X 1
f¯ieq ei,y = f¯eq ξy dξ = ρ∗ u∗y
i=0
n0 c
8
X 1
f¯ieq ei,x ei,y = f¯eq ξx ξy dξ = ρ∗ u∗x u∗y
i=0
n0 c2
8
X 1 1
f¯ieq ei,x ei,x = f¯eq ξx ξx dξ = ρ∗ + ρ∗ u∗2
x
i=0
n0 c2 3
8
X 1 1
f¯ieq ei,y ei,y = f¯eq ξy ξy dξ = ρ∗ + ρ∗ u∗2
y
i=0
n0 c2 3
as it would to be expected.
3.4. EXERCISES 55
3.4 Exercises
1- In Figure 3.3 the dimensionless lattice vectors are such that
√ |ei | = 1
for i = {1, 2, 3, 4} and |ei | = 2 i = {5, 6, 7, 8}. In the same way, the
for
quadrature weights can be found as W0 = 16/36, Wi = 4/36 for i = 1, 2, 3, 4
and Wi = 1/36 for i = 5, 6, 7, 8. Consider that you initialize a LB simulation
attributing a macroscopic velocity u∗ = 0 and a dimensionless density ρ∗ = 1
for all the 1000×1000 sites with a D2Q9 symmetry. Calculate the equilibrium
eq
populations fi in this case.
√
2- The sound speed for this lattice is cs = 1/ 3. Suppose that you are
performing a LB simulation and that, at a given time step, we nd the following
dimensionless populations at a given site x
fi = 0.1 for i = 0,
fi = 0.08 for i = 1, 2, 5,
fi = 0.06 for i = 3, 4, 6, 7, 8,
f −i (xb , t + δ) = f i (xb , t) .
Therefore, in the standard bounce-back scheme (MBB) the collision process is
not carried out on the boundary nodes.
A second bounce-back scheme is a modication of the rst when the colli-
sion is also imposed. In this scheme, the pre-collisional populations are recon-
structed in accordance with the following rule
f −i (xb , t) = f i (xb , t) .
So, in this modied bounce-back scheme all pre-collisional populations be-
comes determined on the boundary nodes and it is then possible to perform
the collision step on these boundary nodes.
The half-way bounce back (HWBB) is the most popular boundary condi-
tion in LBM. As for the BB scheme, for each boundary site xb the outgoing
populations f−i at time t+δ are made equal to the incoming populations f i,
at time t, but now the wall surface is supposed to be located in the half-way
between the boundary and the uid sites (Figures 3.5, 3.6).
This boundary condition assures, for plane walls, no mass leaking and zero
slip. Roughness and surface iregularity are sources of errors in numerical sim-
ulations.
Bounce-back schemes can be extended to complex geometries. Their main
advantage is their simplicity and easy implementation and, being local models,
are frequently used for ows in complex geometries such as ows in porous me-
dia. Actually, some studies have shown that the half-way bounce-back scheme
3.6. BOUNCE BACK SCHEMES 57
Figure 3.4: In the standard bounce-back scheme the boundary nodes are wet
nodes.
Figure 3.5: Half-way bounce-back boundary condition: the wall surface is sup-
posed to be located in the half-way between the boundary and the uid sites.
is of second-order accuracy, while the standard one is only of rst order . Sim-
ilar to the half-way bounce-back scheme, the modied bounce-back scheme is
also of second-order accuracy [38]. It is clear that in the bounce-back schemes,
the momentum of a particle is just reversed before and after it hits the wall,
which means that the macroscopic velocity at the wall is zero. Therefore, these
bounce-back schemes work for stationary walls. As the wall moves with a given
velocity, these schemes should be modied to include the contribution of the
wall motion.
eq
f i = f i (ρ0 , u∗ , v ∗ + v 0 ), i = 1, 5, 8 (3.16)
X
f i = ρW , (3.17)
X
f i eix = ρW u∗ , (3.18)
X
f i eiy = ρW v ∗ . (3.19)
neq neq
f1 = f3 . (3.20)
This last condition was formulated considering the simmetry of the pres-
sure tensor. Thus, the Zou and He approach replaces only unknown particle
populations. The strong point is numerical accuracy, especially in 2D, and the
weak point is the decient numerical stability at high Reynolds numbers. In
the Zou/He method, a bounce-back principle is applied to o-equilibrium parts
of the particle populations. This boundary condition is easy to implement in
2D and 3D.
X
f i eix = ρ∗ u∗
X
f i eiy = 0
neq neq
f1 = f3
3.10 Exercise
1. Find f 1, f 5, f 8 for Inamuro and Zou and He BC for at solid surfaces
and for Zou and He BC for the outlet of a uid domain.
60 CHAPTER 3. A FIRST INSIGHT INTO LBM
Figure 3.8: Zou and He [94] boundary conditions for the inlet.
The incompressible ow in a square cavity whose top wall moves with a
uniform velocity has been widely used as a model problem for testing and
evaluating numerical techniques, in spite of the singularities that we have at
the two top corners [30, 47, 8, 92, 83, 10]. These ows are of great scientic
interest because they display almost all uid mechanical phenomena in the
simplest of geometrical settings. In this kind of ow problem, a main primary
vortex is formed near the center of the cavity and secondary and even tertiary
vortices appear in the corners and are intensied when the Reynolds number
increases (Figure 3.9). In LBM framework, numerical solutions were based on
the BGK model [47], multi-relaxation times [89], [62], entropic approach [3]
and regularization [72].
3.11. SAMPLE CASE: LID DRIVEN CAVITY FLOW 61
Figure 3.9: Streamlines for a square lid driven cavity ow when Re=1000 [ ? ].
[25]
Figure 3.10: Numerical results for Re=100, 400, 1000 and 5000 showing the:
a) the horizontal component of the velocity along the vertical mid-line ; b) the
vertical component of the velocity along the horizontal mid-line of the cavity
?
[ ]. Results are compared with Ghia et al. [30], Botella and Peyret [8] and
Erturk et al. [25].
Chapter 4
Forces in LBM
Forces play a central role in many hydrodynamic problems. This is, for
instance, the case of the gravitational acceleration g, which acts as a source of
momentum leading to a number of hydrodynamic eects. If two uids with dif-
ferent densities are mixed or if the temperature in a uid is nonhomogeneous,
density gradients in the gravitational eld lead to buoyancy eects and phe-
nomena like the Rayleigh-Bénard instability or the Rayleigh-Taylor instability.
In the Rayleigh-Bénard instability, which is essential in studies of heat transfer,
convection patterns develop when warmed uid rises from a hot surface and
falls after cooling (Figure 4.1). The Rayleigh-Taylor instability can occur when
a layer of denser uid is over a lower-density uid layer (4.2). Gravity waves at
a free water surface are another example. Apart from gravity, there are several
other physical problems where forces are important. In incompressible ows,
the driving mechanism of the pressure gradient eld may be, in some cases,
described by a body force. There exist cases where the problem physics spec-
ify pressure gradients, but where it is convenient to replace these with forces.
One reason for this is that the LB method may lose accuracy when solving
pressure elds due to compressibility errors. While this change is possible in
arbitrarily complex ow geometries, the task of nding an equivalent driving
force eld is only trivial in periodic ow congurations. This is often explored
in LB simulations of porous media ows. Forces are also commonly used to
model multiphase or multi-component ows. Furthermore, some algorithms
for uid-structure interactions, e.g. the immersed boundary method, rely on
forces mimicking boundary conditions.
eq
∂t∗ f i + ei .∇f i = Ω i + g∗ . (ei − u∗ )f i , (4.1)
63
64 CHAPTER 4. FORCES IN LBM
D
where the overbar indicates a dimensionless distribution
p f i = fi ξ /n0 , ξ is
the mean molecular speed, ξ = kT /m and is related to the lattice speed
c through ξ = ccs , g∗ is a dimensionless acceleration, g∗ = gδ 2 /h, Ω i is
eq
the collision term and f i is the LBE discrete approximation to the Maxwell-
Boltzmann equilibrium distribution.
In lattice-Boltzmann framework it is a common practice to simplify the
2
second term on the r..h.s. of Eq. (4.1) by neglecting all the u∗ and higher
order terms. It results
1 ∗ ∗ eq 1 ∗ ∗ 1 2
∗
g . (ei − u )f i u 2 Wi ρ g . ei + 2 ei ei − cs δ .u . (4.2)
cs2 cs cs
where
1 ∗ ∗ 1 ∗ ∗ ∗ ∗
1 2
S̄ig = W i ρ g e
α iα + g u + g u e e
iα iβ − c δ
s αβ . (4.4)
c2s 2 α β β α
c2s
∗ ∗ 1 ∗ ∗ ∗ ∗
above equation we have replaced gα uβ by
In the
2 gα uβ + gβ uα because
this replacement increases the symmetry and showed to have benecial eects
on the algorithm.
The Navier-Stokes equations require the exact retrieval of the viscous stress
tensor, which is a second order moment of the non-equilibrium distribution
fineq . So if the NS equations are to be correctly retrieved in a LBE simulation,
without errors up to the rst order in Knudsen it is required to: a) use a second
order approximation to the streaming term; b) use a third order aproximation
to the equilibrium distribution and a corresponding third order set of lattice
vectors ([75], [82] ). When Ω̄i is the BGK relaxation term, Ω̄i = −fineq /τ ∗ ,
the LBM classical schemes based on Eq. (4.5) are only successful in simu-
lating hydrodynamics phenomena, because the second order errors given by
(∂tt fi + eiα eiβ ∂αβ fi + 2eiα ∂tα fi ) contribute as a numerical viscosity which is
absorbed in the momentum balance equation. When the LB equation includes
a force term as in Eq. (4.3) a second or higher order approximation has to be
used for avoiding the presence of spurious terms in the Navier-Stokes equations
[41].
66 CHAPTER 4. FORCES IN LBM
Second order approximations for the stream term of Eq. (4.3) can be found
by considering a second-order Taylor series expansion of the populations
∂t∗ t∗ f i + eiα eiβ ∂α∗ β ∗ f i + 2eiα ∂t∗ α∗ f i = ∂t∗ ∂t∗ f i + eiα ∂α∗ f i
+ eiα ∂α∗ ∂t∗ f i + eiβ ∂β ∗ f i .
and, so,
1
+ ∂t∗ Ω i + S̄ig + eiα ∂α∗ Ω i + S̄ig . (4.7)
2
Consider, now, a rst-order Taylor series expansion of the r.h.s. of Eq. (4.3)
1
f¯i (x + ei h, t + δ) = f¯i (x, t) + Ω i + S̄ig +
Ω i (x + ei h, t + δ) − Ω i (x, t)
2
1
+ S̄ig (x + ei h, t + δ) − S̄ig (x, t) ,
2
leading to
1
f¯i (x + ei h, t + δ) = f¯i (x, t) +
Ω i (x + ei h, t + δ) + Ω i (x, t)
2
1
+ S̄ig (x + ei h, t + δ) + S̄ig (x, t) (4.10)
2
4.2. A SMART PROCEDURE FOR OVERCOMING THE IMPLICITENESS OF THE N
1
f¯i (x, t) = f¯i (x − ei h, t − δ) +
Ω i (x, t) + Ω i (x − ei h, t − δ)
2
1
+ S̄ig (x, t) + S̄ig (x − ei h, t − δ) . (4.11)
2
2. Since all populations are known in the neighbor sites x − ei h at the
previous time step t − δ, it is possible to nd the density ρ(x, t) by summing
Eq. (4.11) on i,
X
ρ∗ (x, t) = f¯i (x − ei h, t − δ)
i
" ! !#
1 X X
+ Ω i (x − ei h, t − δ) + S̄ig (x − ei h, t − δ) .
2 i i
X
ρ∗ u∗γ (x, t) = f¯i (x − ei h, t − δ)
i
!
1 X
+ Ω i (x − ei h, t − δ)eiγ
2 i
!
1 X
+ ρ∗ gγ∗ + S̄ig (x − ei h, t − δ)eiγ ,
2 i
because
X 1 X
S̄ig eiγ = ρ∗ gα∗ Wi eiα eiγ = ρ∗ gγ∗ .
i
c2s i
4. We can then calculate the equilibrium distribution fieq (ρ∗ , u∗ ) and the
source term S̄ig (x, t) from Eq. (4.4) .
5. Population f¯i (x, t) can be nally calculated by using a slight modication
of Eq. (4.11)
68 CHAPTER 4. FORCES IN LBM
¯1
1 + ∗ f (x, t) = f¯i (x − ei h, t − δ)
2τ i
1 1 ¯eq
+ f + Ω i (x − ei h, t − δ)
2 τ∗ i
1
+ S̄ig (x, t) + S̄ig (x − ei h, t − δ) .
2
1
f˜i = f¯i −
Ω i + S̄ig , (4.12)
2
which for a BGK collision term, can be also written as
1 1 ¯eq
f˜i = f¯i − ¯
f − fi + S̄ig ,
2 τ∗ i
enabling to retrieve the original distribution f¯i from known values of f˜i , f¯ieq (ρ∗ , u∗ ),
∗
S̄ig (ρ )
1 ¯eq
f˜i + 1
2τ ∗ fi + 2 S̄ig
f¯i = . (4.13)
1 + 2τ1∗
With the transformation Eq. (4.12) , Eq. (4.10) becomes
f¯eq − f˜i
1
f˜i (x + ei h, t + δ) = f˜i (x, t) + i + 1− S̄ig (x, t), (4.14)
τ̃ 2τ̃
where τ̃ = τ ∗ + 1/2. Eq. (4.14) represents an explicit numerical scheme for the
modied distribution f˜i .
From Eq. (4.12)
X
f˜i = ρ∗ , (4.15)
i
X 1X
f˜i eiγ = ρ∗ u∗γ − S̄ig eiγ ,
i
2 i
1X 1
S̄ig eiγ = ρ∗ gγ∗ ,
2 i 2
because,
4.4. SIMULATION WORK: PLANE POISEUILLE FLOW 69
X
Wi eiα eiγ = c2s δαγ .
i
Therefore,
P ˜
fi eiγ 1
u∗γ = i ∗ + gγ∗ . (4.16)
ρ 2
2 !
y∗
3
u∗x = hu∗x i 1 − , (4.17)
2 H∗
70 CHAPTER 4. FORCES IN LBM
1 P.S. Laplace, Théorie analytique des probabilités (1812) livre 2, 321-323; Oeuvres VII
2 P.L.Chebyshev: Sur le développement des fonctions à une seule variable Bull. Acad.
Sci. St. Petersb. I 1859 193-200; Oeuvres I 501-508.
71
72 CHAPTER 5. THE DISCRETIZATION PROBLEM
work was overlooked and they were named later after Charles Hermite who
3
wrote on the polynomials in 1864 describing them as new . In his later 1865
papers, Hermite was the rst to dene the multidimensional polynomials.
dθ
ξ2 2
(θ) θ − ξ2
H (ξ) = (−1) e 2 e , (5.1)
dξ θ
where θ is the polynomial degree and ξ is a one-dimensional independent vari-
able. The rst Hermite polynomials are
H (0) (ξ) = 1,
(1)
H (ξ) = ξ,
H (2) (ξ) = ξ 2 − 1,
H (3) (ξ) = ξ ξ2 − 3 ,
H (4) (ξ) = ξ 4 − 6ξ 2 + 3.
Exercise
Derive the above expressions for the Hermite polynomials using Eq. (5.1)
∞
1 ξ2
f ∗g = √ e− 2 f gdξ,
2π −∞
ξ2
where e− 2 is a weight function.
Hermite polynomials are orthogonal, in the sense that
∞
1 ξ2
√ e− 2 H (θ) (ξ) H (η) (ξ) dξ = 0,
2π −∞
when θ 6= η .
When θ = η,
∞ 2
1 ξ2
√ e− 2 H (θ) dξ = θ!
2π −∞
3 C. Hermite: Sur un nouveau développement en série de fonctions C. R Acad. Sci. Paris
58 1864 93-100; Oeuvres II 293-303
4 Square ∞
integrable functions f : (−∞, ∞) → R are functions such that
−∞ |f (x)|2 dx is
a nite number
5.1. HERMITE POLYNOMIALS IN ONE DIMENSION 73
X a(θ)
f (ξ) = H (θ) (ξ) .
θ!
θ
ξ2 H (η) (ξ)
Mutiplying both sides by e− 2 √
2π
and integrating,
∞ X a(θ) ∞
1 ξ2 1 ξ2
√ e− 2 H (η) (ξ) f (ξ) dξ = √ e− 2 H (η) (ξ) H (θ) (ξ) dξ ,
2π −∞ θ! 2π −∞
θ | {z }
=θ!δηθ
Exercise
Show that
∞ 2 ∞
1 − ξ2
2 1 ξ2
√ e H (3)
dξ = 3! , √ e− 2 H (3) H (1) dξ = 0.
2π −∞ 2π −∞
2
f eq 1 (ξ−u)
− 2kT
= 1/2 e m .
n 2π kT
m
2 2
(ξ − u) (ξ0 − u0 ) T0
2kT
= ,
m
2 T
(ξ0 −u0 )2
eq f eq ξ 1 1 −
2 T
f = = ρ∗ 1/2 1/2
e T0
,
n0 (2π)
T
T0
where ρ∗ = n/n0 .
Dening
74 CHAPTER 5. THE DISCRETIZATION PROBLEM
T
Θ= − 1,
T0
as a relative temperature deviation with respect to T0 , we get
eq 1 1 (ξ0 −u0 )2
f = ρ∗ 1/2
√ e− 2(Θ+1) .
(2π) Θ+1
or
(ξ0 −u0 )2
eq 1 ξ2 √ 1
Θ+1
e− 2(Θ+1)
− 20
f = ρ∗ e .
1/2 ξ2
(2π) e − 20
5
The function
(ξ0 −u0 )2
√ 1
Θ+1
e− 2(Θ+1)
2
ξ0
,
e− 2
1 (ξ0 −u0 )2 2 X
ξ0 1 (θ) (θ)
√ e− 2(Θ+1) = e− 2 a H (ξ0 ) .
Θ+1 θ!
θ
1
Multiplying both members by √
2π
H (η) (ξ0 ) and integrating, we get the
Hermite coecients of the expansion
∞
1 1 (ξ0 −u0 )2
a(η) = √ √ e− 2(Θ+1) H (η) (ξ0 ) dξ0 .
2π −∞ Θ+1
ξ0 −u0
The variable ξ0 is now changed to x = √
Θ+1
resulting in the following
a(0) = 1,
a(1) = u0 ,
a(2) = u20 + Θ,
a(3) = u0 u20 + Θ
and so on.
So, in one-dimension the MB distribution becomes a series of Hermite poly-
nomials with the following form
2 ∞
eq 1 ξ0 X 1 (θ)
f = ρ∗ 1/2
e− 2 a (u0 , Θ) H (θ) (ξ0 ) .
(2π) θ!
θ=0
5 When Θ = 0, this function is the generating function of the Hermite polynomials.
5.2. HERMITE POLYNOMIAL TENSORS 75
Exercise
Starting from Eq. (5.2) show that the rst Hermitian moments in one-
dimensional spaces are
H (0) (ξ) = 1.
n
(n−1)
X
Hα(n+1)
0 α1 ...αn
= ξα0 Hα(n)
1 ...αn
− δα0 αk Hα1 ...αn /αk (5.3)
k=1
H (0) = 1,
Hα(1) = ξα ,
(2)
Hαβ = ξα ξβ − δαβ ,
(3)
Hαβγ = ξα ξβ ξγ (5.4)
Exercise
Derive the expressions for Hermite polynomial tensors in Eq. (5.4) using
the recurrence equation, Eq. (5.3).
1 ξ2
f ∗g = D/2
e− 2 f gdξ.
(2π)
76 CHAPTER 5. THE DISCRETIZATION PROBLEM
Hermite polynomial tensors are also orthogonal with respect to the above
internal product and their norm is given by
1 ξ2 2
D/2
e− 2 H (θ) dξ = n1 !...nD !,
(2π)
where n1 is the number of times the index x appears repeated in the polynomial
tensor,n2 is the number of times the index y appears repeated in the polynomial
tensor and D is the dimension of the velocity space (D = 1, 2, 3).
Let
X a(θ)
f (ξ) = H (θ) (ξ) .
n1 !...nD !
θ
1 ξ2
a(η) = D/2
e− 2 f (ξ) H (η) (ξ) dξ.
(2π)
Exercise
Considering a two-dimensional space, show that
1 ξ2 2
D/2
e− 2 (2)
Hxx dξ = 2!
(2π)
and that
1 ξ2 2
D/2
e− 2 (2)
Hxy dξ = 1!
(2π)
eq 1 1 (ξ0 −u0 )2
f = ρ∗ D/2 D e− 2(Θ+1) . (5.5)
(2π) (Θ + 1) 2
The function
(ξ0 −u0 )2
√ 1
Θ+1
e− 2(Θ+1)
2
ξ0
,
e− 2
1 (ξ0 −u0 )2 2
ξ0 1
e−
X
2(Θ+1) = e− 2 a(θ) H (θ) (ξ0 ) . (5.6)
(Θ + 1)
D
2 n1 !...nD ! rθ
θ
1
Multiplying both members by
(2π)D/2
H (η) (ξ0 ) and integrating
5.3. VELOCITY DISCRETIZATION 77
1 1 (ξ0 −u0 )2
ξ0 −u0
Changing the variable ξ0 by r= √
Θ+1
we get
√
1 r2
a(η) = D/2
e− 2 H (η) Θ + 1r + u0 dr. (5.7)
(2π)
In components notation, the rst equilibrium moments are
a(0)
eq = 1,
a(1)
eq,α = u0,α ,
(2)
aeq,αβ = (u0,α u0,β + Θδαβ ) ,
(3) u0,α u0,β u0,γ
aeq,αβγ = ,
+Θ (δβγ u0,α + δαγ u0,β + δαβ u0,γ )
(4)
aeq,αβγδ = u0,α u0,β u0,γ u0,δ
δαβ u0,γ u0,δ + δαγ u0,β u0,δ
+Θ +δαδ u0,β u0,γ + δβγ u0,α u0,δ
+δβδ u0,α u0,γ + δγδ u0,α u0,β
+Θ2 (δαβ δγδ + δαγ δβδ + δαδ δβγ ) .
Exercise
Starting from Eq. (5.7), show that
b
f (x) dx,
a
X b
Wi f (xi ) = f (x) dx.
i a
for ϕ (ξ0 ) = 1, ξ0,α , ξ0,α ξ0,β , ξ0,α ξ0,β ξ0,γ , .... In fact, as it was demonstrated by
McNamara [69], the preservation of the equilibrium moments in the discrete
space is the main condition for the exact retrieval of the macroscopic hydrody-
namic equations and this is so, because when the Knudsen number Kn → 0, all
the non-equilibrium moments can be expressed in terms of spatial derivatives
of equilibrium moments.
Nevertheless, we are, here, asked to decide the order N of ϕ (ξ0 ) up to
which the quadrature condition must be satised and N will give the order of
approximation of the LBE to the continuous Boltzmann equation.
So, considering the velocity discretization as a quadrature problem, the
quadrature equation becomes
b −1
nX
eq eq
ωi f (ξ0,i ) ϕ (ξi,0 ) = f (ξ0 ) ϕ (ξ0 ) dξ0 .
i=0
2
ξ0,i
D/2
Further it is convenient, although not mandatory, to replace ωi by Wi (2π) e 2 .
b −1
nX 2
ξ0 eq eq
D/2
Wi (2π) e 2f (ξ0,i )ϕ (ξi,0 ) = f (ξ0 ) ϕ (ξ0 ) dξ0 . (5.8)
| {z }
i=0
ωi
| {z }
=fieq
eq
and, so, our discrete populations fi becomes
2
ξ0
eq D/2 eq
f i = Wi (2π) e 2 f (ξ0,i )
Therefore, as before, solving this quadrature problem is what is wanted,
leading to a set of discrete velocities ξ0,i and weights Wi in such a manner that
all the equilibrium moments of interest are preserved.
5.3. VELOCITY DISCRETIZATION 79
2
eq 1 ξ0 X 1
f (ξ0 ) = ρ∗ D e− 2 a(θ) (u0 , Θ) H (θ) (ξ0 ) , (5.9)
(2π) 2 n1 !...nD !
θ
Replacing Eq. (5.9) into Eq. (5.8), the following equation is obtained
nXb −1
X 1 (θ)
a (u0 , Θ) Wi H (θ) (ξi,0 ) ϕ (ξi,0 )
n1 !...nD ! i=0
θ
2
X 1 1 ξ0
= a(θ) (u0 , Θ) D e− 2 H (θ) (ξ0 ) ϕ (ξ0 ) dξ0 . (5.10)
n1 !...nD ! (2π) 2
θ
Nevertheless, the velocity monomials ϕ = 1, ξ0,α , ξ0,α ξ0,β ... can be, also,
6
expanded in a nite sum of Hermite polynomial tensors
N
X
ϕ (ξ0 ) = b(η) H (η) (ξ0 ) . (5.11)
η=0
N nXb −1
XX 1
b(η) a(θ) (u0 , Θ) Wi H (θ) (ξi,0 ) H (η) (ξi,0 )
η=0
n1 !...nD ! i=0
θ
N 2
XX 1 1 ξ0
= b(η) a(θ) (u0 , Θ) D e− 2 H (θ) (ξ0 ) H (η) (ξ0 ) dξ0 ,
η=0
n1 !...nD ! (2π) 2
θ
or
N
XX 1
b(η) a(θ) (u0 , Θ)
η=0
n1 !...nD !
θ
Pnb −1
i=0 Wi H (θ) (ξi,0 ) H (η) (ξi,0 )
× 2
ξ0
− 1
D e− 2 H (θ) (ξ0 ) H (η) (ξ0 ) dξ0
(2π) 2
= 0
6 In eect
1 = H0 ,
ξ0,α = H1,α ,
ξ0,α ξ0,β = H2,αβ + δαβ H0 ,
and so on.
80 CHAPTER 5. THE DISCRETIZATION PROBLEM
2
1 ξ0
D/2
e− 2 H (θ) (ξ0 ) H (η) (ξ0 ) dξ0
(2π)
b −1
nX
= Wi H (θ) (ξ0,i ) H (η) (ξ0,i ) , (5.12)
i=0
for all θ, η smaller or equal than N . In fact, in the discrete space, the nite
expansion of ϕ (ξ0 ) up to order N restricts the expansion of f eq to the same
order, meaning that the poles ξi and weights Wi are only required to satisfy a
nite set of equations.
Solving a discrete LBE gives an approximation to the full kinetic equation,
whose accuracy can be as high as it can be achieved by increasing the order of
approximation of the Hermite expansion and, consequently, the number nb of
discrete velocities used in the representation of the continuous velocity space.
Nevertheless, since the main interest in numerically solving a LBE is to
describe the behaviour of a physical system whose macroscopic equations are
known, or supposed to be known, the expansion of the MB distribution in a
nite set of orthogonal polynomial tensors of order N of approximation has the
important feature of preserving the moments hϕp i of order p, smaller or equal
than N,
eq
fN ϕp dξ = f eq ϕp dξ for p ≤ N.
By dening the inner product in the discrete space ξ (N ) that maps the
velocity space onto the real numbers as
b −1
nX
f ∗g = Wi f (ξ0,i ) g (ξ0,i ) , (5.13)
i=0
it can be seen that Eq. (5.12) requires that the norm and orthogonality of
Hermite polynomial tensors are preserved in the discrete space ξ (N ) for all θ, η
smaller or equal than N. This means that the metrics of the continuous velocity
space must be preserved for all velocity functions whose order is smaller than
N.
In Philippi et al. [75], it is shown that when the discrete velocity space is
invariant under π/2 rotations and reections about the x, y and z axis, the
norm preservation of the Hermite polynomial tensors in ξ (N )
2 i2 b −1
nX i2
1 ξ0
h h
D/2
e− 2 H (θ) (ξ0 ) dξ0 = Wi H (θ) (ξ0,i ) , (5.14)
(2π) i=0
Finaly, from Eqs. (5.9) and (5.8), it is seen that the projections of the MB
distribution along the discrete poles ξ0,i are,
N
eq X 1
f i = ρ∗ Wi a(θ) (u0 , Θ) H (θ) (ξ0,i )
n1 !...nD !
θ=0
H (0) = 1,
(1)
H = ξi,0 = as ei ,
(2) 2
H = ξi,0 − 1 = a2s e2i − 1.
b−1 i2 2 i2
X h 1 ξ0
h
Wi H (θ) (ξ0,i ) = D/2
e− 2 H (θ) (ξ0 ) dξ0 ,
i=0 (2π)
for θ = 0, 1, 2, giving
∞ 2
1 ξ0
W0 + W1 + W1 = 1/2
e− 2 dξ0 , for θ = 0,
(2π) −∞
2 2 2
W0 (as e0 ) + W1 (as e1 ) + W1 (as e−1 )
∞
1 ξ2
− 20 2
= 1/2
e ξ0 dξ0 , for θ = 1,
(2π) −∞
2 2 2
W0 a2s e20 − 1 + W1 a2s e21 − 1 + W1 a2s e2−1 − 1
∞
1 ξ2 2
− 20
= 1/2
e ξ02 − 1 dξ0 , for θ = 2.
(2π) −∞
Since
∞
1 x2
1/2
e− 2 dx = 1,
(2π) −∞
∞
1 − x2
2
1/2
e x2 dx = 1,
(2π) −∞
∞
1 x2 2
1/2
e− 2 x2 − 1 dx = 2,
(2π) −∞
W0 + W1 + W1 = 1,
W1 a2s + W1 a2s = 1,
2 2
W0 + W1 a2s − 1 + W1 a2s − 1
= 2,
2 1
√
whose solution is: W0 = , W1 = , as = 3.
3 6
The discrete equilibrium distributions are
eq ∗ (0) (0) (1) (1) 1 (2) (2)
f0 = W0 ρ a H (as e0 ) + a H (as e0 ) + a H (as e0 )
2
∗ 1 2
= W0 ρ 1 + u0 as × e0 + (u0 u0 + Θ) as e0 e0 − 1
2
1
W0 ρ∗ 1 − u2 + Θ ,
=
2 0
eq ∗ (0) (0) (1) (1) 1 (2) (2)
f1 = W1 ρ a H (as e1 ) + a Hα (as e1 ) + a H (as e1 )
2
∗ 1 2
= W1 ρ 1 + u0 as × e1 + (u0 u0 + Θ) as e1 e1 − 1
2
∗ 1 2 2
= W1 ρ 1 + u0 as + u + Θ as − 1 ,
2 0
5.3. VELOCITY DISCRETIZATION 83
1
eq
f−1 = W1 ρ∗ a(0) H (0) (as e−1 ) + a(1) Hα(1) (as e−1 ) + a(2) H (2) (as e−1 )
2
1
= W1 ρ∗ 1 + u0 as × e−1 + (u0 u0 + Θ) a2s e−1 e−1 − 1
2
1 2
= W1 ρ∗ 1 − u0 as + u0 + Θ a2s − 1 .
2
Writing u0 = as u∗ , the above expressions become
1 ∗2
f0eq = W0 ρ∗ 1 − 3u + Θ ,
2
h 2 i
f1eq = W1 ρ 1 + 3u∗ + 3u∗ + Θ ,
∗
h 2 i
eq
f−1 = W1 ρ∗ 1 − 3u∗ + 3u∗ + Θ .
We see that
X
fieq = ρ∗ ,
i
X
fieq ei = ρ∗ u∗ ,
i
1 X eq 2 1 ∗ 1 2
f e = ρ (1 + Θ) + ρ∗ u∗ .
2 i i i 6 2
n kT0 T
P = nkT = mn0 .
n0 m To
But
kT0 2 1
= ξ = c2 c2s with cs = .
m as
So
P
P∗ = = ρ∗ c2s (Θ + 1) .
mn0 c2
For the kinetic energy of uctuations
D D kT0 n T
nec,f = nkT = n0 m
2 2 m n0 T0
| {z }
=c2 c2s
84 CHAPTER 5. THE DISCRETIZATION PROBLEM
n ec,f 1
ρ∗ e∗c,f = = ρ∗ c2s (Θ + 1) (5.15)
n0 mc2 2
√
Parameter cs Θ + 1 can be shown to be related to the dimensionless speed
of sound in the lattice. In fact, the speed of sound us is given by
∂T
∂s ∂s
∂P ∂ρ ∂T P ∂ρ
u2s = = − ∂s P = − ∂s ∂T P
∂ρ s ∂P ρ ∂T ρ ∂P ρ
dT dρ dT m ∂ρ
ds = cv −k = cp + 2 dP
T ρ T ρ ∂T P
ρ
Since P = m kT ,
∂T
cp ∂ρ cp
kT0 T D+2 2 2
u2s =− P = = c c (Θ + 1)
cv ∂T
∂P ρ
cv m T0 D } s
| {z
|{z} | {z }
= D+2 =c2 c2s =3
D
2
u 2
s
u∗s = = 3c2s (Θ + 1)
c
H (0) = 1,
Hx(1) = ξ0x = as ei,x ,
Hy(1) = ξ0y = as ei,y ,
(2) 2
Hxx = ξ0x − 1 = a2s e2i,x − 1,
(2) 2
Hyy = ξ0y − 1 = a2s e2i,y − 1,
(2)
Hxy = ξ0x ξ0y = a2s ei,x ei,y .
b −1
nX i2 2 i2
h 1 ξ0
h
Wi H (n) (ξ0,i ) = D/2
e− 2 H (n) (ξ0 ) dξ0 , (5.16)
i=0 (2π)
5.3. VELOCITY DISCRETIZATION 85
giving
W0 + 4W1 + 4W2 = 1,
4
X 8
X
W0 a2s e20,x + W1 a2s e2i,x + W2 a2s e2i,x = 1,
i=1 i=5
4
X 8
X
W0 a2s e20,y + W1 a2s e2i,y + W2 a2s e2i,y = 1,
i=1 i=5
4 8
2 X 2 X 2
W0 a2s e20,x − 1 + W1 a2s e2i,x − 1 + W2 a2s e2i,x − 1 = 2,
i=1 i=5
4 8
2 X 2 X 2
W0 a2s e20,y − 1 + W1 a2s e2i,y − 1 + W2 a2s e2i,y − 1 = 2,
i=1 i=5
4 8
2 X 2 X 2
W0 a2s e0,x e0,y + W1 a2s ei,x ei,y + 4W2 a2s ei,x ei,y = 1.
i=1 i=5
Only four of the six above equations are linearly independent, giving the
16 4 1
√
values of
36 , W1 = 36 , W2 = 36 and the scaling factor as =
W0 = 3.
The equilibrium distribution written in terms of discrete variables reads
86 CHAPTER 5. THE DISCRETIZATION PROBLEM
2
eq X 1
f i = Wi ρ∗ a(θ) (u0 , Θ) H(θ) (as ei ) .
n1 !...nD !
θ=0
a(0) + a(1) 1
a(2)
a2s ei,x ei,x − 1
|{z} α as ei,α + 2 xx
=1 |{z} |{z}
u0,α =as u∗ =a2s u∗ 2
eq α x +Θ
= Wi ρ∗
fi + 1 a(2) 2
(2) 2
(5.17)
2 yy as ei,y ei,y − 1 + axy as ei,x ei,y
|{z} |{z}
=a2s u∗ 2
y +Θ
=a2s u∗ ∗
x uy
Figure 5.4: Lattices D2Q17 and D2Q21. For the D2Q21 lattice there will
be seven unknowns as , W0 , W1 , W2 , W3 , W4 .W5 for six norm restrictions, after
eliminating identical equations. Letting as be a free variable, the system gives
a solution with real positive roots when as is inside the interval 0.659836 and
1.16208. The values as =0.659836 and as =1.16208 are roots of the polynomials
W0 (as ) and W3 (as ) , respectively. Therefore, when the value as =1.16208 is
chosen, W3 =0, corresponding to the lattice vectors in red, giving a modication
of the D2Q17 lattice, which was named D2V17.
Therefore, when the value as =1.16208 is chosen, W3 =0 and the lattice loses
an energy level, giving a modication of the D2Q17 lattice, which was named
D2V17 [75]. The weights, with six signicant digits, are W0 =0.402005, W1 =0.116155,
W2 =0.0330064, W3 =0, W4 =0.0000790786, and W5 =0.000258415.
5.3.3 Summary
Consider the subspace HN of H generated by the rst Hermite polynomial
tensors
(2) (N )
H (0) , Hα(1) , Hαβ , ...Hαβ...θN ,
There is a relationship between the order N of the Hilbert space HN and the
discrete and nite velocity set ei , i = 0, ..., nb − 1. Figure 5.5 shows this rela-
tionship for the two-dimensional D2Q9 lattice, a second order LBE and Figure
5.6 this same relationship for the D2V17 LBE. In fact, velocity discretization
means to replace the MB equilibrium distribution by its projection onto the
space HN and the key question in the velocity discretization problem is to nd
the set ei , i = 0, ..., nb −1 in accordance with the order N of HN . This question
has not a single answer and the prescribed abcissas method above presented
leads to dierent velocity sets, with the same order N, when they are required
to be space lling or not. Further details can be found in [75], [82], [84], [87]
and [81].
88 CHAPTER 5. THE DISCRETIZATION PROBLEM
ei = {0, 1, −1} .
The D2Q9 set of lattice vectors is the Cartesian product of two D1Q3 LBE
along the directions x and y
(0, 0) , (1, 0) , (0, 1) , (−1, 0) , (0, −1) ,
ei = .
(1, 1) , (−1, 1) , (−1, −1) , (1, −1) ,
√
Now, for the D1Q3 LBE we found as = 3 and W0 = W (ke0 k) = 64 , W1 =
1
√
W (ke1 k) = 6 . For the D2Q9 LBE the scaling factor is, also, as = 3 and the
weights are the products of the corresponding one-dimensional weights
4 4 16
W0 = W (ke0 k) × W (ke0 k) = × = ,
6 6 36
4 1 4
W1 = W (ke0 k) × W (ke1 k) = × = ,
6 6 36
1 1 1
W2 = W (ke1 k) × W (ke1 k) = × = .
6 6 36
The Hermitian representation of this LBE can be also retrieved as the Carte-
sian product of the corresponding one-dimensional representations (Table 5.1).
(2)
In fact, consider, e.g., the Hermite polynomial Hαβ . From the recurrence
equation, Eq. (5.3)
(2) (1)
Hαβ = Hα(1) Hβ − 1.
So, when α=β and γ = δ, but dierent from α (or β)
(4) (2) (2)
Hαβγδ = Hαβ Hγδ .
This result can be generalized for higher order Cartesian products between
Hermite polynomials.
Therefore
(3) (2)
(eix ) × Hy(1) (eiy ) = a2s e2ix − 1 as eiy ,
Hxxy (eix , eiy ) =
Hxx
(3)
(eix , eiy ) = Hx(1) (eix ) × Hyy
(2)
(eiy ) = a2s e2iy − 1 as eix ,
Hxyy
(4) (2) (2)
(eiy ) = a2s e2ix − 1 a2s e2iy − 1 .
Hxxyy (eix , eiy ) = Hxx (eix ) × Hyy
5.4. CARTESIAN PRODUCTS 91
Figure 5.7: Hermitian representations for the D2Q9 LBE. (a) Second order Her-
mitian basis; (b) Full Hermitian basis as obtained from the Cartesian product
of one-dimensional representations.
eq ∗ ∗ 9 ∗ 2 3 ∗2
f i,q=6 = Wi ρ 1 + 3u · ei + (u · ei ) − u , (5.19)
2 2
where q represents the number of Hermite polynomials in the basis.
When the D2Q9 LBE is considered as the cartesian product of two D1Q3
one-dimensional LBE, the equilibrium distribution is given by
2 2
1 + 3u∗ · ei + 29 (u∗ · ei ) − 23 u∗
∗2 ∗
+ 27 1
2
2 ux uy eix − 3 eiy
eq
f i,q=9 = Wi ρ∗ , (5.20)
∗2 ∗
+ 27 2 1
2 uy ux eiy − 3 eix
81 ∗2 ∗2
+ 4 ux uy eix − 3 e2iy − 31
1
2
eix = {0, 1, −1} , eiy = {0, 1, −1} and eiz = {0, 1, −1}
Figure 5.8: D3Q27 full Hermitian basis. All the Hermite polynomials in red
n o n o
(1) (2) (1) (2)
results from the Cartesian product H (0) , Hx , Hxx × H (0) , Hy , Hyy ×
n o
(1) (2)
H (0) , Hz , Hzz
(0, 0, 0) ,
(1, 0, 0) , (0, 1, 0) , (0, 0, 1) , (−1, 0, 0) , (0, −1, 0) , (0, 0, −1)
(1, 1, 0) , (1, 0, 1) , (0, 1, 1) (−1, 1, 0) , (−1, 0, 1) , (0, −1, 1)
ei = (−1, −1, 0) , (−1, 0, −1) , (0, −1, −1) , (1, −1, 0) , (1, 0, −1) (0, 1, −1) .
(1, 1, 1) ,
(1, 1, −1) , (1, −1, 1) , (−1, 1, 1) , (1, −1, −1) , (−1, 1, −1) , (−1, −1, 1) ,
(−1, −1, −1)
√
For the D3Q27 LBE the scaling factor is, also, as = 3 and the weights are
the products of the corresponding one-dimensional weights
4 4 4 64
W0 = W (ke0 k) × W (ke0 k) × W (ke0 k) = × × = ,
6 6 6 216
4 4 1 16
W1 = W (ke0 k) × W (ke0 k) × W (ke1 k) = × × = ,
6 6 6 216
4 1 1 4
W2 = W (ke0 k) × W (ke1 k) × W (ke1 k) = × × = ,
6 6 6 216
1 1 1 1
W3 = W (ke1 k) × W (ke1 k) × W (ke1 k) = × × = .
6 6 6 216
The D3Q27 full Hermitian representation is shown in Figure 5.8.
ei = {−2, −1, 0, 1, 2} .
There are three weights W0 = W (0), W1 = W (1) and W2 = W (2) and a
scaling factor as to be found and the norm preservation condition Eq. (5.12)
2 2 b −1
nX 2
1 ξ0
D/2
e− 2 H (θ) (ξ0 ) dξ0 = Wi H (θ) (as ei ) , (5.21)
(2π) i=0
W0 + 2W1 + 2W2 = 1,
2
2W1 a2s + 2W2 (2as ) = 1,
W0 + 2W1 a4s − 2a2s + 1 + 2W2 a4s 24 − 2a2s 22 + 1
= 2,
2 2
2 2
2W1 (as ) − 3 as + 2W2 (as 2) − 3 2as = 6.
ei = {−3, −1, 0, 1, 3} .
W0 + 2W1 + 2W2 = 0!
2
2W1 a2s + 2W2 (3as ) = 1!
W0 + 2W1 a4s − 2a2s + 1 + 2W2 a4s 34 − 2a2s 32 + 1
= 2!
2W1 9a2s − 6a4s + a6s + 2W2 9 × 32 a2s − 6 × 34 a4s + 36 a6s
= 3!
This system has two solutions with positive weights and scaling factor.
One of such solutions is
√
4 4 + 10
W0 = ,
45√
3 8 − 10
W1 = ,
80√
16 − 5 10
W2 = ,
s 720 √
5 + 10
as = .
3
94 CHAPTER 5. THE DISCRETIZATION PROBLEM
2 √
X
(4) 16 + 5 10
Wi Hxxxx (as eix ) = 4!,
i
6
and not 4!
Therefore, we rewrite the fourth order Hermite polynomial as
(4)
(4) Hxxxx
Kxxxx = ,
γ
where
s √
16 + 5 10
γ= . (5.22)
6
Figure 5.9 presents the D2Q25-ZOT LBE and Figure 5.10 its corresponding
Hermitian basis.
5.5. THE MOMENTS SPACE 95
(θ)
X
fi = Ki a(θ) , (5.23)
(θ)
Hi ∈Hq
(θ)
(θ) Wi Hi (θ)
where Ki = nx !ny ! , with Hi = H (θ) (as ei ),
(0) (8) 1
K0 = W (ke0 k) H (0) (as e0 ) , ..., K0 = (4)
W (ke0 k) Hxxyy (as e0 )
2!2!
(0) (8) 1
K1 = W (ke1 k) H (0) (as e1 ) , ..., K1 = (4)
W (ke1 k) Hxxyy (as e1 )
2!2!
(0) (8) 1
K2 = W (ke2 k) H (0) (as e2 ) , ..., K2 = (4)
W (ke2 k) Hxxyy (as e2 )
2!2!
.............................................................
(0) (8) 1
K8 = W (ke8 k) H (0) (as e8 ) , ..., K8 = (4)
W (ke8 k) Hxxyy (as e8 )
2!2!
This gives a 9×9 matrix for the D2Q9 LBE, which can be written as
96 CHAPTER 5. THE DISCRETIZATION PROBLEM
4
0 0 − 29 − 29 0 0 0√ 1
9 √ 9
3
1
9 9 0
√
1
9
1
− 18 0 0√ − 183 1
− 18
3
1
9 0√ 9
1
− 18 1
9 0 − 183 0
√
1
− 18
1 3 1 1 3 1
9 − 9 0√ 9 − 18 0 0
√ 18 − 18
1
−√ 93 3
1 1
1
K=
9 √
0 − 18 9 0 18
√ √
0 − 18
1 3 3 1 1 1 3 3 1
36 36
√ 36
√ 36 36 12 36
√ 36
√ 36
1 3 3 1 1 1 3 3 1
36 −√36 36
√ 36 36 − 12 36
√
−√36 36
1
36 −√363 −√36
3 1
36
1
36
1
12 −√36
3
−√363 1
36
3
1
36 36 − 363 1
36
1
36
1
− 12 − 363 3
36
1
36
Since
X Wi H (θ) H (η)
i i
= δθη
i
nx !ny !
due to the orthogonality condition for the Hermite polynomials, each moment
a(η) can be written as
(η)
X
a(η) = Hi f i (5.24)
i
H (0) (e0 ) ... H (0) (e8 )
(1) (1)
H = Hx (e0 ) ... Hx (e8 ) ,
(4) (4)
Hxxyy (e0 ) ... Hxxyy (e8 ) 9×9
or
H = K−1 =
1 √1 1 1
√ 1 √1 1
√ 1
√ √1
0
3 √0 − 3 √0 √3 −√ 3 −√ 3 √3
0
0 3 0 − 3 3 3 − 3 − 3
−1 2 −1 2 −1 2 2 2 2
−1 −1 2 −1 2 2 2 2 2
0
0 √0 0 √ 0 √3 −3
√ 3√ −3
√
0
0
√ − 3 √0 3 2√3 2 √3 −2√3 −2√ 3
0 − 3 0 3 0 2 3 −2 3 −2 3 2 3
1 −2 −2 −2 −2 4 4 4 4
Matrix H can be considered as a linear transformation from the space gen-
T
erated by the 9 populations fb = f (ei ) , i = 0, ..., 8 to the space generated
by the 9 moments
h iT
a = a(0) , a(1)
b (1) (2) (2) (2) (3) (3) (4)
x , ay , axx , axy , ayy , axxy , , axyy , axxyy
5.5. THE MOMENTS SPACE 97
and K as its inverse H−1 , retrieving the populations f i from a set of 9 moments,
a = Hfb
b fb = Kb
a
Therefore, since H is independent of the position and time, it can be cal-
culated with its inverse K and stored in the computer memory, allowing to
compute on each point x, for each time step, the moments a(θ) from the in-
coming populations fi and vice-versa. The three rst moments are equilibrium
eq
moments and dene the equilibrium populations f i . The non-equilibrium
neq eq
populations are given by f i = f i − f i and the non-equilibrium moments by
neq
aneq = Hfb .
b
The collision step can thus be written as
out eq 1 neq
fb = fb + 1 − ∗ fb ,
τ
where
fb eq = aeq
Kb (5.25)
fb neq = Kb
aneq
. (5.26)
In this procedure, the collision step is thus performed with the information
we have for the equilibrium and non equilibrium moments at (x, t). It is a regu-
larized kinetic scheme because the non equilibrium populations are recalculated
by Eq. (5.26) previously to the collision step, ltering the high-order moments
that are outside the Hq=9 Hermitian representation. The moments method
together with the idea of replacing the BGK collision step by the relaxation of
the non-equilibrium moments with dierent relaxation times was introduced
into the LBM framework by D'Humières [23]. Nevertheless, contrary to nd-
ing a set of orthogonal polynomials by the Gram-Schmidt orthogonalization
method as in D'Humières methodwhich can be a cumbersome process, espe-
cially for high order LBEthese polynomials are here considered as discrete
forms of the Hermite polynomials. The main condition is that these polyno-
mials satisfy Eq. (5.14), preserving, in the discrete space, the same metrics of
the continuous space.
This procedure can be extended to all LBE that are Cartesian products of
one-dimensional LBE.
98 CHAPTER 5. THE DISCRETIZATION PROBLEM
Chapter 6
Chapman-Enskog analysis
In this chapter we investigate the necessary conditions for a LBE for retrieving
the macroscopic transport equations. First we remember some main points of
the Chapman-Enskog analysis [11], [12], restricting ourselves to a system of
material points without external forces.
∂t f + ξ · ∇f = Ω. (6.1)
τ /Γ Ω+
∂t+ f + ξ + · ∇+ f = . (6.2)
`/L `/L
When collisions are dominant as it happens in continuous systems, we ex-
pect that
` L,
τ Γ.
`
= .
L
99
100 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
The dimensionless group multiplying the rst term on the l.h.s. of Eq. (6.2)
can be written as
τ /Γ χ
= ,
`/L
where χ = τ / Γ. The term χ/ is expected to be of O(1), although it can be an
important dimensionless number in problems with high oscillating macroscopic
1
variables. This term is, in fact, usually identied as the Strouhal number ,
L
= Sr.
Γξ
Now, when the molecular collisions prevail over advection, we can consider
that the distribution function f is close to the equilibrium distribution and
write it as an asymptotic expansion
∂t (ρu) + ∇ · ρuu + P = 0. (6.4)
Pαβ = mf ξα ξβ dξ,
and so, the asymptotic expansion Eq. (6.3) will induce the decomposition of
the pressure tensor in accordance with the order of the Knudsen number
(0) (1) 2
Pαβ = mf ξα ξβ dξ + mf ξα ξβ dξ + mf (2) ξα ξβ dξ + ...,
| {z } | {z } | {z }
(0) (1) (2)
Pαβ Pαβ Pαβ
meaning that the time derivative of the momentum in Eq. (6.4) is due to
2
contributions with dierent orders in the Knudsen number. Enskog proposed
to decompose this time derivative in accordance with
1 Strouhal, V. (1878) "Ueber eine besondere Art der Tonerregung" (On an unusual sort of
sound excitation), Annalen der Physik und Chemie, 3rd series, 5 (10) : 216251.
2 Enskog, D. Kinetische theorie der vorgänge in mässig verdünnten gasen. I. Allgemmeiner
teil. Inaugural Dissertation, Published by Alqvist & Wiksells, 1917, Uppsala, 1917
6.1. THE CONTINUOUS CASE 101
(0)
∂0 (ρu) + ∇ · ρuu + P = 0, (6.6)
(1)
∂1 (ρu) + ∇ · P = 0, (6.7)
(2)
2 2
∂2 (ρu) + ∇ · P = 0, (6.8)
and so on.
When Eqs. (6.3) and (6.5) are replaced into Eq. (6.2), we get the following
equation
χ
(∂0+ + ∂1+ + ∂2+ ...) f (0) + f (1) + 2 f (2) + ...
+ξ + · ∇+ f (0) + f (1) + 2 f (2) + ...
(0)+
Ω (1)+ (2)+
= +Ω + Ω + ....
By rearranging the terms in accordance with their order
χ χ χ
∂0 f (0) + ξ + · ∇+ f (0) + ∂0 f (1) + ∂1 f (0) + ξ + · ∇+ f (1) +
| {z } | {z }
order 1 order
χ χ χ +
∂0 2 f (2) + ∂1 f (1) + 2 ∂2 f (0) + 2 ξ · ∇+ f (2) + ...
| {z }
order 2
(0)+
Ω (1)+ (2)+
= +Ω + Ω + ....
The above equation is then split into the following equations,
(0)+
Ω = 0,
χ (1)+
∂0 f (0) + ξ + · ∇+ f (0) = Ω ,
χ (1) χ (2)+
∂0 f + ∂1 f (0) + ξ + · ∇+ f (1) = Ω ,
and so on.
We can then return to the original variables, obtaining
Ω(0) = 0, (6.9)
(0) (0) (1)
∂0 f + ξ · ∇f = Ω , (6.10)
(0) (1)
∂1 f + ∂0 f + ξ · ∇ f (1) = Ω 2 (2)
, (6.11)
102 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
and so on.
Therefore, from Eq. (6.9), the order-zero solution f (0) is the Maxwell-
Boltzmann equilibrium distribution
f (0) = f eq .
Multiplying Eq. (6.10) by the mass m and integrating in the velocity space
∂0 mf (0) dξ + ∂β mξ f (0) dξ = 0,
β
or
∂0 ρ + ∂β (ρuβ ) = 0, (6.12)
(0)
∂0 (ρuα ) + ∂β Pαβ = 0. (6.13)
where P = nkT is the ideal gas pressure. The hydrodynamic equations are the
Euler equations and there is no viscous transfer of momentum.
Consider now that we multiply Eq. (6.10) by the molecular kinetic energy
1 2
2 mξ
1 2 (0) 1 2
∂0 mξ f dξ + ∂β mξ ξ f (0) dξ = 0. (6.14)
2 2 β
1 2
We know that the molecular kinetic energy
2 mξ may be decomposed,
following
1 2 1 2 1
mξ = m (ξα − uα ) + mξα uα − mu2α .
2 2 2
In addition,
ξα ξβ = (ξα − uα ) (ξβ − uβ ) + ξα uβ + ξβ uα − uα uβ .
6.1. THE CONTINUOUS CASE 103
1 1 2
∂0 ρec,f + ρu2 + ∂β ρec,f + ρu uβ + P uβ = 0, (6.15)
2 2
where
1 2
ρec,f = m (ξα − uα ) f (0) dξ,
2
is the peculiar kinetic energy per unit volume of the system of particles and,
at the zeroth-order, Eq. (6.15) gives the balance of the total kinetic energy,
1 2
considered as composed by ρec,f 2 ρu .
and the advective energy
At order zero in Knudsen, the balance equation for the internal energy can
be found by subtracting the balance equation for the momentum at this order
from Eq. (6.15).
So, when we perform the internal product of Eq. (6.13) by the macroscopic
velocity,
1 2 1
uα ∂0 (ρuα ) = ∂t ρu + u2 ∂0 ρ,
2 2
whereas
1 2 1
uα ∂β (ρuα uβ ) = ∂β ρu uβ + u2 ∂β (ρuβ ) ,
2 2
in such a manner that
1 2 1 2
uα ∂0 (ρuα ) + uα ∂β (ρuα uβ ) = ∂0 ρu + ∂β ρu uβ .
2 2
On the other hand
δαβ uα ∂β (P ) = ∂β (P uβ ) − P ∂β (uβ ) .
So, the balance equation for the advection energy becomes,
1 2 1 2
∂0 ρu + ∂β ρu uβ + P uβ = P ∇ · u. (6.17)
2 2
Eq. (6.17) is then subtracted from Eq. (6.15) giving
which is the balance equation for the internal energy at the Kn zeroth-order.
104 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
Order 1 in Knudsen
At the much faster time scale t1 = t0 , the solution f (1) is retrieved by solving
Eq. (6.10). Considering for simplicity a BGK collision model [4], then
f (1)
Ω(1) = − ,
τ
and, in this case the rst order approximation to the distribution function f (1)
and the viscous stress tensor ταβ are found by writting the l.h.s. of Eq. (6.10)
in terms of spatial derivatives of the hydrodynamic elds.
So, when Eq. (6.11) is multiplied by the mass m and integrated in the
velocity space, it gives
∂1 ρ = 0. (6.19)
Consider now, multiplying Eq. (6.11) by the momentum mξα and integrat-
ing the result in the velocity space
∂1 mξα f (0) dξ + ∂β mξα ξβ f (1) dξ = 0,
or
ταβ = mξα ξβ f (1) dξ.
1 2
Finally, we multiply Eq. (6.11) by the kinetic energy
2 mξα and integrate
the result in the velocity space
1 2 (0) 1 2
∂1 mξ f dξ + ∂β mξ ξβ f (1) dξ = 0. (6.21)
2 2
| {z }
=ρec,f + 21 ρu2
1 2 1 2 1 2 (1)
mξ ξβ f (1) dξ = mξ (ξβ − uβ ) f (1) dξ+uβ mξ f dξ .
2 2 2
| {z }
=0
But
1 2 1
mξ = m (ξα − uα ) (ξα − uα ) + m (ξα − uα ) uα − muα uα
2 2
and so,
6.1. THE CONTINUOUS CASE 105
1 2 1 2
mξ ξβ f (1) dξ = mξ (ξβ − uβ ) f (1) dξ
2 2
1
= m (ξα − uα ) (ξα − uα ) (ξβ − uβ ) f (1) dξ
2
| {z }
(1)
=qβ
+uα m (ξα − uα ) (ξβ − uβ ) f (1) dξ
| {z }
(1)
=ταβ
− (uα uα ) m (ξβ − uβ ) f (1) dξ ,
| {z }
=0
where the vector giving the ux of heat q is given by the diusive transport of
the peculiar kinetic energy
1 2
q= m (ξ − u) (ξ − u) f neq dξ.
2
Finally
1
(1) (1)
∂1 ρec,f + ρu2 + ∂β qβ + uα ταβ = 0, (6.22)
2
or
1
∂1 ρec,f + ρu2 + ∂β (qβ + uα ταβ ) = 0 (6.23)
2
∂t ρ + ∂β (ρuβ ) = 0. (6.24)
Finally, we sum Eqs. (6.15) and (6.23) for retrieving the equation for the
energy conservation,
1 2 1 2
∂t
2
ρu + ρec,f + ∂β 2 ρu + ρec,f uβ + P uβ + ταβ uα + qβ = 0.
(6.26)
106 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
1 2 1 2
∂t ρu + ∂β ρu uβ + uβ P + uα ταβ = P ∇ · u + ταβ ∂β uα . (6.27)
2 2
The balance equation for the internal energy can be then obtained by sub-
tracting Eq. (6.27) from Eq. (6.26),
The last term on the r.h.s. is the viscous dissipation term φ = −ταβ ∂β uα ,
which can be shown to be always non-negative and represents the irreversible
transformation of advective energy into internal energy.
6.1.2 Exercise
When the collision operator Ω is modelled following a BGK relaxation term
f − f eq
Ω=−
τ
this relaxation term is decomposed in accordance with
f (0) − f eq
Ω(0) = −
τ
f (1)
Ω(1) = −
τ
f (2)
Ω(2) = −
τ
and so on.
a) Using Eq. (6.10), show that the viscous stress tensor components, in
(1)
order 1 in Knudsen, ταβ = ταβ , when calculated as the moment
Ω(1) ξα ξβ dξ,
is given by
b) Using Eq. (6.10), nd the heat ux vector , in order 1 in Knudsen,
(1)
qα = qα , when calculated as the moment
1 2
Ω(1) m (ξβ − uβ ) (ξα − uα ) dξ.
2
f i (x + ei h, t + δ) − f i (x, t) = Ωi + O h2 .
(6.29)
3
where, as before, f i = fni ξ̄0 is the population of particles fi = f (ξi ), in dimen-
ξ¯ = kT0 /m is the mean molecular speed, and T0 and n0 are,
p
sionless form,
respectively, a reference temperature and a reference number density.
We develop f i (x + ei h, t + δ) in a Taylor series around f i (x, t),obtaining
1 1
δ∂t f i + heiα ∂α f i + δ 2 ∂tt f i + h2 eiα eiβ ∂αβ f i + hδeiα ∂tα f i + ... = Ωi . (6.30)
|2 2 {z }
second order errors
So, in the rst-order LB, the second-order terms in the above equation must
be considered as errors with respect to the continuous Boltzmann equation.
Let L be a macroscopic length and Γ a macroscopic time and dene
h
= ,
L
and
δ
χ= .
Γ
Although we are using the same symbols and χ as before, their meanings
are not the same. Symbol is now being used for the spatial resolution and the
symbol χ for the ratio between the time step δ and the macroscopic scale of
time. In LBM, the Knudsen number continues to be given by the ratio between
the mean free path and a macroscopic characteristic length,
` ξτ h
Kn = = ,
L h L
and since ξ = ccs = hδ cs , it can also be written as
ν∗
Kn = cs τ ∗ = .
cs
108 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
Remark that, in addition to the geometric resolution and the scaling factor
cs a lattice constant the Knudsen number depends on the dimensionless
relaxation time τ ∗ = τ /δ . This relaxation time is a measure of the importance
of the collision term with respect to the stream term. Small values of τ∗ means
a fast relaxation to an equilibrium state and corresponds to a low kinematic
viscosity. So, in low-viscous uids the interactions between the particles are
so numerous and eective that momentum and energy between neighboring
particles is quickly balanced; the system remains close to the equilibrium state.
As a result, the net ux of momentum is small in the moving frame of reference
or, in other words, transport of momentum is sustained mainly by advection
rather than by diusion.
Both and χ are required to be small parameters, the rst one being related
to the geometric resolution of the physical system and the second one to the
ratio between the time step δ and the time Γ that gives a scale for the time
variation of the local elds. We consider χ ≈ and replace t by t+ = t/Γ and
x by x+ = x/L in Eq. (6.30),
2
1χ
2 ∂t t f i
+ +
χ Ωi
∂t+ f i + eiα ∂α+ f i + + 12 eiα eiβ ∂α+ β + f i + ... = .
| {z } +χeiα ∂t+ α+ f i |{z}
O(1) | {z } O(−1 )
O()
χ (0) (0)
∂0+ f i + eiα ∂α+ f i +
| {z }
O(1)
(1) (1) (0) 2 (0)
!
1χ
χ∂0+ f i + eiα ∂α+ f i + χ∂1+ f i +
2 ∂0 0 f i
+ +
(0) (0) + ...
+ 12 eiα eiβ ∂α+ β + f i + χeiα ∂0+ α+ f i
| {z }
O()
(0)
Ωi (1) (2)
= + Ωi + Ωi + ... . (6.31)
| {z } |{z} | {z }
O(1) O()
O(1/)
So, in O(1/):
6.3. ZERO-TH ORDER TRANSPORT EQUATIONS 109
(0)
Ωi = 0.
(0)
and this equation gives the zero-th order approximation fi to the solution as
related to the equilibrium solution
(0) eq
fi = fi ,
usually written as a polynomial approximation to the Maxwell-Boltzmann equi-
librium distribution.
In O(1) :
χ (0) (0) (1)
∂ + f + eiα ∂α+ f i = Ωi , (6.32)
0 i
or, returning to the LB variables,
X (0) 1
fi = ρ∗ = f eq dξ, (6.36)
i
n0
X (0) ∗ ∗ 1
f i ei =ρ u = f eq ξdξ, (6.37)
i
cn0
and
X
Ωi = Ωdξ =0. (6.38)
i
(0) eq
Therefore, since fi = fi , we see that, for retrieving the zero-th approx-
imation to the mass conservation equation, the Hermite polynomial approxi-
eq
mation for fi must satisfy the conditions given by Eqs. (6.36-6.37) and, in
addition, the modelled collision term must be written in such way as to sat-
isfy Eq. (6.38), meaning that mass must be preserved by the colision model
considering the discrete set ei , i = 0, ..., b − 1 of lattice vectors.
X (0) 1
f i eiα eiβ = ρ∗ u∗α u∗β ∗
+ P δαβ = f eq ξα ξβ dξ, (6.40)
i
n 0 c2
and
X
Ωi ei = Ωξdξ = 0. (6.41)
i
P ∗ = ρ∗ c2s (Θ + 1)
and corresponds to the Clayperon pressure for an ideal gas, P = nkT . Eq.
(6.41) means that the collision model and the set ei , i = 0, ..., b − 1 of lattice
vectors are required to satisfy the momentum preservation in collisions.
(0) eq
Eq. (6.40) is only satised when fi = fi is, at least, a second order
Hermitian representation.
6.3. ZERO-TH ORDER TRANSPORT EQUATIONS 111
Exercise
Show that the D2Q9 LBE with an equilibrium distribution given by
a4s
eq
h i
f i,q=6 = Wi ρ∗ + a2s ρ∗ u∗α eiα + 2 ρ∗ u∗α u∗β + Θ eiα eiβ − 1
a2s δαβ ,
(6.42)
satises the condition given by Eq. (6.40).
! !
1 X (0) 1 X (0) X (0) 1
∂0∗ f eiβ eiβ + ∂α∗ f eiα eiβ eiβ = Ωi eiβ eiβ
2 i i 2 i i i
2
which gives the zero − th order aproximation of the energy conservation equa-
tion
1 ∗ ∗2 1 ∗ ∗2
∂0∗ ρ u + ρ∗ e∗c,f + ∇∗ · ρ u + ρ∗ e∗c,f u∗ + P ∗ u∗ = 0, (6.43)
2 2
3
where the internal energy per unit volume is given in its dimensionless form
by
D ∗ 2
ρ∗ e∗c,f = ρ cs (Θ + 1) ,
2
For retrieving the energy conservation equation at order zero, Eq. (6.43),
the following conditions must be satised
1 X (0) 1 2 1 1
f eiβ eiβ = ρ∗ u∗ + ρ∗ e∗c,f = f eq ξβ ξβ dξ, (6.44)
2 i i 2 n0 c2 2
and
1X (0) 1 ∗ ∗2 1 1
mf i eiα eiβ eiβ = ρ u + ρ∗ e∗c,f u∗α +P ∗ u∗α = f eq ξα ξβ ξβ dξ,
2 i 2 n0 c3 2
(6.45)
which means a null net heat ow vector at this order,
3 The internal energy is the kinetic energy resulting from the velocity uctuations of the
molecules,
1
ρe = m (ξ − u)2 dξ.
2
112 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
(0)
q = 0.
In addition, the following condition must also be satised,
X
Ωi eiβ eiβ = Ωξ 2 dξ = 0, (6.46)
i
meaning that the collision model Ωi and the set ei , i = 0, ..., b − 1 of lattice
vectors are required to satisfy the preservation of kinetic energy in collisions.
Exercise
It is important to remark that Eq. (6.45) requires more than a second
eq
order Hermitian representation of fi . Verify that the D2Q9 LBE with the
equilibrium distribution given by Eq. (6.42) do not satisfy the condition given
by Eq. (6.45) .
But
1 ∗ ∗2 1
u∗β ∂0∗ ρ∗ u∗β = ∂0∗ + u∗2 ∂0∗ ρ∗ ,
ρ u
2 2
and
1 ∗ ∗2 ∗ 1
u∗β ∂α∗ ρ∗ u∗α u∗β = ∇∗ · + u∗2 ∇∗ · (ρ∗ u∗ ) .
ρ u u
2 2
Finally,
u∗ · ∇∗ P ∗ = ∇∗ · (P ∗ u∗ ) − P ∗ ∇∗ · u∗ ,
giving the balance equation for the advection energy at the zero-th order,
1 ∗ ∗2 1 ∗ ∗2 ∗
∂0∗ ρ u + ∇ · ρ u u + P u = |P ∗ ∇{z
∗ ∗ ∗ ∗
· u∗}. (6.47)
2 2
source term
Thus, the pressure P∗ that is advected with the ow contributes, to the
1 ∗ ∗2
balance of the advective energy
2 ρ u , at this order of approximation. The
∗ ∗ ∗
source of the advection energy is σ (Eadv ) = P ∇ · u . Since the potential
energy, related to external forces is not being considered, the total energy Et
has only two contributions, the advection Eadv and the internal energy E
Et = Eadv + E.
6.4. FIRST-ORDER TRANSPORT EQUATIONS 113
Consequently,
σ (E) = −σ (Eadv ) = −P ∗ ∇∗ · u∗ ,
giving for the internal energy, the following balance equation at order zero
The term between brackets in the l.h.s. of the above equation can be,
further rearranged
So, we obtain
114 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
(1)
(1)
(0)
∂0∗ f i + ∂α∗ eiα f i + ∂1∗ f i
1h (1) (1)
i
+ ∂0∗ Ωi + eiβ ∂β ∗ Ωi
2
(2)
= 2 Ω i . (6.49)
∂1 ρ∗ = 0, (6.50)
X (1) X (1)
fi = eiα f i = 0,
i i
eq
f ϕ (ξ) dξ = f ϕ (ξ) dξ,
(1)
(1) f
Ωi = − i∗ . (6.51)
τ
Eq. (6.49) becomes
(0)
∂1∗ f i
(1)
!
f (1)
1
∂0∗ i∗ + eiα ∂α∗ f i
+ 1− 2τ ∗ τ
(2)
= 2 Ωi . (6.52)
Multiply Eq. (6.52) by eiβ and sum on i, using the previous relations, and
the condition
6.4. FIRST-ORDER TRANSPORT EQUATIONS 115
X (2)
Ωi eiβ = 0. (6.53)
i
We obtain
1
ρ∗ u∗β
∂1∗ + ∂α∗ 1− ∗ ταβ = 0, (6.54)
2τ
4
where
(1)
X (1)
ταβ = ταβ = f i eiα eiβ ,
i
(1)
(0) (0) f i
∂0∗ f i + eiα ∂α∗ f i =− ,
τ∗
and so, making the inner product of the above equation by eiα eiβ ,
X (0) (0)
ταβ = −τ ∗ eiα eiβ ∂0∗ f i + eiγ ∂γ ∗ f i . (6.55)
i
The following conditions must be satised by the LBE for correctly retriev-
ing the viscous stress tensor
X (0) 1
eiα eiβ f i = f eq ξα ξβ dξ, (6.56)
i
n 0 c2
X (0) 1
eiα eiβ eiγ f i = f eq ξα ξβ ξγ dξ. (6.57)
i
n 0 c3
The rst condition Eq. (6.56) is, in fact, the condition given by Eq. (6.44)
and results in
X (0) 2 ∗ ∗
eiα eiβ f i = ρ e δαβ + ρ∗ u∗α u∗β .
i
D
The second condition, Eq. (6.57) requires a third order Hermite polynomial
(0)
expansion for fi . When this condition is satised it gives
(1) (2)
ταβ = ταβ + 2 ταβ + ...
Considering
(1)
ταβ = ταβ ,
corresponds to a Knudsen rst-order Chapman-Enskog approximation.
116 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
X (0)
eiα eiβ eiγ f i
i
2 ∗ ∗
ρ e δαβ u∗γ + δαγ u∗β + δβγ u∗α + ρ∗ u∗α u∗β u∗γ .
= (6.58)
D
After some cumbersome calculations the viscous stress tensor is given by
where
2
∗ ∗ 1 ∗ ∗ ∗ ∗ 1
µ = τ − ρ e =ρ τ − c2s (Θ + 1) .
D 2 2
is the shear viscosity and
2 ∗
κ∗ = µ ,
D
is the bulk viscosity for a mechanical system of material points, without internal
degrees of freedom.
Replacing ταβ in Eq. (6.54) by its expression given by Eq. (6.59), the
momentum balance equation at order 1 is obtained as
So, the second order error terms resulting from a rst order approximation
of the stream term in Eq. (6.29) are absorbed into the momentum balance
equation as a numerical viscosity. The only restriction is that the dimensionless
relaxation time τ∗ must be greater than
1
2.
X !
1 1 (1) 1
∂1∗ ρ∗ e∗c,f + ρ∗ u∗2 + ∂α∗ 1− ∗ f i eiα eiβ eiβ = 0.
2 2τ i
2
Using
eiα eiβ eiβ = (eiα − u∗α ) eiβ eiβ + u∗α eiβ eiβ
= (eiα − u∗α ) eiβ − u∗β eiβ − u∗β + 2 eiβ − u∗β u∗β
we obtain
∗
=qα
z }| {
X (1) 1 1 X (1) ∗ 1 ∗ 2
f i eiα eiβ eiβ = f i (eiα − uα ) eiβ − uβ
i
2 2 i 2
X (1)
+u∗β f i (eiα − u∗α ) eiβ − u∗β .
i
| {z }
∗
=ταβ
1 1
ρ∗ e∗c,f + ρ∗ u∗2 qα∗ + u∗β ταβ
∗
∂1∗ + ∂α 1− = 0.
2 2τ ∗
The heat ow vector is given by
1 X (1) 2
qα∗ = f i (eiα − u∗α ) eiβ − u∗β .
2 i
From Eq. (6.33)
(1)
(0) (0) f
∂0∗ f i + eiγ ∂γ ∗ f i = − i∗ ,
τ
Therefore
1 X 2 (0)
qα∗ = −τ ∗ ∂0∗ (eiα − u∗α ) eiβ − u∗β f i
2 i
1 X 2 (0)
−τ ∗ ∂γ ∗ (eiα − u∗α ) eiβ − u∗β eiγ f i ,
2 i
meaning that the following conditions must be satised for correctly retrieving
the heat ow vector
X 2 (0) 1 2
(eiα − u∗α ) eiβ − u∗β fi = f eq (ξα − uα ) (ξβ − uβ ) dξ = 0,
i
n0 c3
(6.61)
X 2 (0)
(eiα − u∗α ) eiγ eiβ − u∗β fi
i
1 2
= f eq (ξα − uα ) (ξγ − uγ ) (ξβ − uβ ) dξ. (6.62)
n0 c4
5
Condition given by Eq. (6.62) means that
5 Note that
118 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
X 2 (0) 2
(eiα − u∗α ) eiγ − u∗γ eiβ − u∗β = ρ∗ c4s (Θ + 1) δαγ
fi
i
1
qα∗ = −τ ∗ ∂α∗ ρ∗ c2s (Θ + 1) c2s (Θ + 1)
2
∗1 2
= −τ cs ∂α∗ (P ∗ (Θ + 1))
2
τ∗ τ∗
= − (Θ + 1) c2s ∂α∗ P ∗ − ρ∗ c4s (Θ + 1) ∂α∗ Θ.
2 2
ρ∗ c4s
1
λ∗ = τ∗ − (Θ + 1) ,
2 2
1 ∗ ∗2 1 δ
u∗β ταβ ταβ ∂α∗ u∗β ,
∂1∗ ρ u + ∂α∗ 1− ∗ = 1−
2 2τ 2τ
which is the kinetic energy balance equation. Therefore, the internal energy
balance equation will be in the rst order in Knudsen,
1 1
∂1 ρ∗ e∗c,f + ∂α∗ qα = − 1 − ∗ ταβ ∂α∗ u∗β .
1− (6.63)
2τ ∗ 2τ
1 2
f eq (ξα − uα ) (ξγ − uγ ) ξβ − uβ dξ
n0 c4
2 2
1 kT0 T 1 C2
= ρ∗ D
e− 2 Cα Cγ C 2 dC
c4 m T0 (2π) 2
1 C2
= ρ∗ c4s (Θ + 1)2 D
e− 2 Cα Cγ C 2 dC.
(2π) 2
| {z }
=δαγ
6.5. MACROSCOPIC EQUATIONS 119
∂1∗ ρ∗ = 0,
giving
1
∂1∗ ρ∗ u∗β + ∂α∗
1− ταβ = 0,
2τ ∗
giving
1
ρ∗ u∗β ρ∗ u∗α u∗β ∗
∂t∗ +∂ α∗ + P δαβ = −∂ α∗ 1− ∗ ταβ .
2τ
1 1
∂1∗ ρ∗ e∗c,f + ∂α∗ 1 − ∗ qα = − 1 − ∗ ταβ ∂α∗ u∗β ,
2τ 2τ
giving
1 1
∂t ρ∗ e∗c,f ρ∗ e∗c,f u∗β ∗
∂β u∗β −∂α∗ qα − 1 − ∗ ταβ ∂α∗ u∗β .
+∂β = −P 1− ∗
2τ 2τ
120 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
X
Ωi = 0,
i
X
Ωi eiβ = 0,
i
X
Ωi e2i = 0,
i
X (0)
fi = ρ∗ , (6.65)
i
X (0)
f i eiα = ρ∗ u∗α , (6.66)
i
X (0)
f i eiα eiβ = ρ∗ u∗α u∗β + P ∗ δαβ , (6.67)
i
X (0) 1
eiα eiβ eiγ f i = f eq ξα ξβ ξγ dξ. (6.68)
i
n0 c3
X 2 (0)
(eiα − u∗α ) eiγ eiβ − u∗β fi
i
1 2
= f eq (ξα − uα ) (ξγ − uγ ) (ξβ − uβ ) dξ (6.69)
n 0 c4
When conditions given by Eqs. (6.65-6.68) are satised, the momentum
balance equation becomes
(6.70)
6.7. THE VISCOUS STRESS TENSOR 121
!
X (0) 2
∂0∗ (ρ∗ e∗ ) δαβ + ∂0∗ ρ∗ u∗α u∗β .
∂0∗ eiα eiβ f i = (6.71)
i
D
From Eq. (6.48)
2
∂0∗ (ρ∗ e∗ ) = −∂γ ∗ ρ∗ e∗ u∗γ − ρ∗ e∗ ∂γ ∗ u∗γ .
D
On the other hand
∂0∗ ρ∗ u∗α u∗β = u∗β ∂0∗ (ρ∗ u∗α ) + u∗α ∂0∗ ρ∗ u∗β − u∗α u∗β ∂0∗ ρ∗ .
2 ∗ ∗
∂0∗ (ρ∗ u∗α ) ∗ ∗ ∗
= −∂γ ∗ ρ uα uγ + ρ e δαγ .
D
So, it results for Eq. (6.71)
!
X (0)
∂0∗ eiα eiβ f i
i
2 2
−u∗β ∂γ ∗ρ∗ u∗α u∗γ − u∗β ∂α∗ (ρ∗ e∗ ) − u∗α ∂γ ∗ ρ∗ u∗β u∗γ − u∗α ∂β ∗ (ρ∗ e∗ )
=
D D
∗ ∗ ∗ ∗
2 ∗ ∗ ∗
2 ∗ ∗ ∗
+uα uβ ∂γ ∗ ρ uγ − ∂γ ∗ ρ e uγ + ρ e ∂γ ∗ uγ δαβ .
D D
122 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
!
X (0)
∂γ ∗ eiα eiβ eiγ f i
i
2 2 2
∂γ ∗ ρ∗ e∗ u∗γ δαβ + u∗β ∂α∗ (ρ∗ e∗ ) + ρ∗ e∗ ∂α∗ u∗β
=
D D D
2 2
+ u∗α ∂β ∗ (ρ∗ e∗ ) + ρ∗ e∗ ∂β ∗ u∗α + ∂γ ∗ ρ∗ u∗α u∗β u∗γ
D D
Adding the two last equations
! !
X (0) X (0)
∂0∗ eiα eiβ f i + ∂γ ∗ eiα eiβ eiγ f i
i i
−u∗β ∂γ ∗ ρ∗ u∗α u∗γ − u∗α ∂γ ∗ ρ∗ u∗β u∗γ + u∗α u∗β ∂γ ∗ ρ∗ u∗γ
=
2
∗ ∗ ∗ ∗ 2 2
ρ∗ e∗ ∂γ ∗ u∗γ δαβ + ρ∗ e∗ ∂α∗ u∗β + ∂β ∗ u∗α
+∂γ ∗ ρ uα uβ uγ −
D D
But
So, the viscous stress tensor, Eq. (6.55) is, nally, given by
(1)
ταβ = −µ∗ ∂α∗ u∗β + ∂β ∗ u∗α + κ∗ ∇∗ · u∗ δαβ .
(6.72)
where
2 ∗ ∗
µ∗ = τ ∗ ρ e
D
and
2 ∗
κ∗ = µ
D
6.8 Exercises
1. Show that when the D2Q9 LBE is used with an equilibrium distribution
given by
ρ∗ + ρ∗a2su∗α eiα +
" #
eq ∗
f i,q=6 = Wi ρ a4s , (6.73)
2 ρ∗ u∗α u∗β + Θ eiα eiβ − 1
a2s δαβ
f (1)
∂0 f (0) + ∂α ξα f (0) = − . (6.74)
τ
the Knudsen rst order approximation to the viscous stress tensor is given by
(1)
τβγ = f (1) (ξβ − uβ ) (ξγ − uγ ) dξ = f (1) ξβ ξγ dξ, (6.75)
f i (x + ei h, t + δ) − f i (x, t) = Ωi + S̄i uO h2 ,
(6.76)
eq
S̄i = g∗ . (ei − u∗ ) f i ,
gis a dimensionless acceleration, g∗ = gδ 2 /h and Ωi is a collision term. usually
a BGK relaxation term [4].
By expanding the equilibrium distribution in Hermite polynomials and ne-
2
glecting all the terms of order u∗ and higher, the source term S̄i can be also
written as
1 ∗ ∗ 1 ∗ ∗ ∗ ∗
2
S̄i ≈ W i ρ gα e iα + g u + g u e iα e iβ − c δ αβ (6.77)
c2s 2c2s α β β α s
f i (x + ei h, t + δ)
in Eq. (6.76) leads to
124 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
1 1
δ∂t f i + heiα ∂α f i + δ 2 ∂tt f i + h2 eiα eiβ ∂αβ f i + hδeiα ∂tα f i + ... = Ωi + S̄i .
2
| 2 {z }
second order errors
(6.78)
h
= ,
L
and
δ
χ= .
Γ
Both and χ are required to be small parameters, the rst one being related
to the geometric resolution of the physical system and the second one to the
ratio between the time step δ and the time Γ that gives a scale for the time
variation of the local elds. We consider χ ≈ and replace t by t+ = t/Γ and
x by x+ = x/L in Eq. (6.78),
2
1χ
2 ∂t t f i
+ +
χ Ωi S̄i
∂t+ f i + eiα ∂α+ f i + + 12 eiα eiβ ∂α+ β + f i + ... = + ,
| {z } +χeiα ∂t+ α+ f i
|{z}
|{z}
O(1) | {z } O(−1 ) O(1)
O()
where we have considered that the source term S̄i is O(). Therefore, when
→ 0, the collision term on the right-hand side of the above equation becomes
dominant with respect to the stream terms on the left-hand side and the source
term.
we obtain
6.9. LBE WITH A FORCING TERM 125
χ (0) (0)
∂0+ f i + eiα ∂α+ f i +
| {z }
O(1)
(1) (1) (0) 2 (0)
!
1χ
χ∂0+ f i + eiα ∂α+ f i + χ∂1+ f i +
2 ∂0 0 f i
+ +
(0) (0) + ...
+ 12 eiα eiβ ∂α+ β + f i + χeiα ∂0+ α+ f i
| {z }
O()
(0)
Ωi S̄i (1) (2)
= + + Ωi + Ωi + ... . (6.79)
} |{z}
| {z |{z} | {z }
O(1) O()
O(1/) O(1)
So, in O(1/):
(0)
Ωi = 0.
(0)
and this equation gives the zero-th order approximation fi to the solution as
related to the equilibrium solution
(0) eq
fi = fi ,
usually written as a polynomial approximation to the Maxwell-Boltzmann equi-
librium distribution.
In O(1), returning to the LB variables,
Finally, in O()
(1) (1) (0) (0)
∂0∗ f i + eiα ∂α∗ f i + ∂1∗ f i + 12 ∂0∗ 0∗ f i
(0) (0) (2) (6.81)
+ 12 eiα eiβ ∂α∗ β ∗ f i + eiα ∂0∗ α∗ f i = 2 Ω i .
Eq. (6.81) can be rearranged using
(1)
(1) f
Ωi = − i∗ ,
τ
and so, Eq. (6.81) becomes
126 CHAPTER 6. CHAPMAN-ENSKOG ANALYSIS
1 (1) (0) 1 S̄i (2)
1− ∗ (∂0∗ + eiα ∂α∗ ) f i +∂1∗ f i + (∂0∗ + eiα ∂α∗ ) = Ωi . (6.82)
2τ 2
1
∂1∗ ρ∗ = − ∂α∗ (ρ∗ gα∗ ) . (6.84)
2
Summing Eqs (6.83) and (6.84),
1
∂t∗ ρ∗ + ∂α∗ (ρ∗ u∗α ) = − ∂α∗ (ρ∗ gα∗ ) , (6.85)
2
which is the mass conservation equation with a spurious, nonphysical, term on
6
its right hand side .
The momentum balance equation can be also obtained from Eqs. (6.80)
and (6.82), giving
1
∂t∗ ρ∗ u∗β + ∂α∗ ρ∗ u∗α u∗β + P ∗ δαβ = ρ∗ gβ∗ − ∂α∗ ∗
1− ταβ + Eβ ,
2τ ∗
(6.86)
where Eβ is, also, a spurious nonphysical term given by
1
∂0∗ ρ∗ gβ∗ + ∂α∗ ρ∗ gα∗ u∗β + gβ∗ u∗α .
Eβ = −
2
Therefore, we see that although the second order errors were absorbed by
the viscous term, the use of a rst order approximation for the stream term is
at the origin of the spurious term into the momentum balance equation, Eq.
(6.86), and is responsible for breaking the mass conservation, Eq. (6.85). In
the following we consider a second order scheme for the stream term.
Exercise
Try to derive the momentum balance equation, Eq. (6.86) from Eqs.
(6.80) and (6.82).
δ
∂t ρ + ∂α (ρuα ) = − ∂α (ρgα )
2
.
6.9. LBE WITH A FORCING TERM 127
f¯ieq − f¯i 1
f i (x + ei h, t + δ) − f i (x, t) = + S̄i (x, t) + S̄i (x + ei h, t + δ) .
τ∗ 2
(6.87)
In this case, we must evaluate S̄i (x + ei h, t + δ) in terms of S̄i (x, t). We
get
f¯ieq − f¯i 1
f i (x + ei h, t + δ) − f i (x, t) = + S̄i + (∂t∗ + eiα ∂α∗ ) S¯i ,
τ∗ 2
and the equation analogous to Eq. (6.79) is given by.
χ (0) (0)
∂0+ f i + eiα ∂α+ f i +
| {z }
O(1)
(1) (1) (0) 2 (0)
!
1χ
χ∂0+ f i + eiα ∂α+ f i + χ∂1+ f i +2 ∂0 0 f i
+ +
(0) (0) + ...
+ 12 eiα eiβ ∂α+ β + f i + χeiα ∂0+ α+ f i
| {z }
O()
¯(0) (1) (2)
1 f¯ieq − fi S̄i fi fi 1
= ∗
+ − ∗
+ − ∗
+ (∂0∗ + eiα ∂α∗ ) S̄i . (6.88)
| τ
{z } | {z τ } | τ 2 {z }
O(1/) O(1) O()
(1)
(1) (1) (0) 1 fi
∂ 0∗ fi + eiα ∂ α∗ fi + ∂ 1∗ fi + 2 (∂0
∗ + eiα ∂ ) − τ ∗ + S̄i .
α∗
(2)
fi
= −2 τ∗ + 12 (∂0∗ + eiα ∂α∗ ) S̄i .
(6.90)
The above equation reduces to
(1)
(1) (1) (0) f
∂0∗ f i + eiα ∂α∗ f i + ∂1∗ f i + 12 (∂0∗ + eiα ∂α∗ ) − τi∗ .
(6.91)
(2)
fi
= −2 τ∗ ,
Eq. (6.91) has the same form of Eq. (6.34): the second order expansion of
the stream term canceled the spurious non-physical terms related to the mo-
mentum source term S̄i and the macroscopic equations are correctely retrieved.
Chapter 7
Beyond BGK:
moments-based methods
Instability issues are frequent in LB simulation with the BGK relaxation
term and increasing the stability ranges is a very actual subject in LBM re-
search. In this section we review the theory that was developed to increase our
understanding on these issues and the methods to enhance the stability limits.
7.1 Introduction
Picturing a uid as a mechanical system of particles has a long and suc-
cessful history, written by outstanding philosophers and scientists, and leading
to the solution of important challenges to the human knowledge. Indeed, the
principle of energy conservation was only postulated by Helmholtz [43], when
matter was recognized as composed by molecules. The works of Joule [49],
Clausius [19], Maxwell [67], Boltzmann [5], [6] and Gibbs [31], were funda-
mental for the borning of the `Mechanical Theory of Heat ' and for clarifying the
meaning of the second principle of thermodynamics and entropy for a system
of particles ruled by the deterministic Newtonian mechanics.
`Dissipation ' means a one-way transformation of macroscopic energy into
heat the mechanical energy of the moleculesand this happens because
molecules collide between themselves exchanging momentum and kinetic en-
ergy.
When a uid ows, these molecular properties are transferred from one
point to another by two kind of mechanisms: diusion and advection. Diu-
sion is related to the random motion of molecules and is responsible for entropy
growth. In the reverse sense, advection means organization: molecular proper-
ties are transferred with the bulk ow, along the streamlines.
Diusion is dominant at the small Reynolds regime, but when the Reynolds
number increases, diusion becomes less eective and properties are mainly
transferred by advection.
129
130 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
Until presently, the great majority of LB models are second order models
whose moment space are tailored to a particular discrete velocity set (D2Q9,
D3Q27, D3Q19) by, e.g., resorting to the Gram-Schmidt orthogonalization pro-
cedure.
Improving stability of these LB schemes has been addressed by several
authors and some solutions have been proposed. For recent reviews, see e. g.
Brownlee et al. [9], Golbert et al. [32] and references therein.
The rst solutions were based on the use of multiple relaxation times (MRT)
tuned with the help of a linear stability analysis, D'Humieres [23], D'Humières
et al. [24], Dellar [22], Lallemand and Luo [57], and Xu and Sagaut [90]. For a
recent paper dealing with MRT coupled with the Kupershtokh [53], [54] exact
dierence method (EDM) in the study of non-isothermal droplet evaporation,
see Albernaz et al. [2] and references therein.
The entropic LBGK method appears as a second solution and was conceived
based on the maximization of the entropy, by locally tuning the BGK single
relaxation time, Karlin et al. [51]. Recently, a new extension of LB schemes
was proposed, namely the entropic stabilizer, Karlin et al. [50]. Unlike the
entropic LBGK model, this Karlin et al. extension do not locally alter the
viscosity, but rather relies on modifying the relaxation time for the higher-order
moments (i.e. the moments beyond the stress tensor) which do not contribute
to the viscosity. In this respect, this extension is akin to the already mentioned
relaxation parameter tuning for MRT schemes.
Finally, the cascaded model and its subsequent improvements, factorized
and cumulant LB schemes [28], [78], [29], base its approach on peculiar mo-
ments and particular choices for the relaxation times for assuring Galilean in-
variance. Nevertheless, these models are, also, restricted to second order D2Q9
and D3Q27 LBE.
Considering the Boltzmann equation in fact, a continuous kinetic model
able to fulll the main conditions that are required for the description of a given
physical system a lattice Boltzmann equation (LBE) can be considered as
a discrete form of a kinetic projection of this kinetic model onto a nite Hq
subspace of H.
The connection between the LBE and such a projection was rst established
by He and Luo [40] and Abe [1] in 1997, who directly derived the LBE from
the Boltzmann equation for some widely known lattice-Boltzmann equations:
D2Q9, D2Q6, D2Q7 and D3Q27. This was performed by the discretization
of the velocity space, using the Gauss-Hermite and Gauss-Radau quadrature.
Excluding the above mentioned lattices, the discrete velocity sets obtained with
this kind of quadrature do not generate regular space lling lattices.
Shan et al. [82] and Philippi and co-workers [75], [84] established a sys-
tematic link between Hermitian representations and the sets of LBE discrete
velocities. This approach led to recognize a LB equation as an approximation
of the Boltzmann or any kinetic equation based on a nite and discrete
set of nb molecular velocities ξi , i = 0, ..., nb − 1.
This set is directly related to the projection of the kinetic equation onto a
subspace generated by the rst q Hermite polynomials and the order N of this
132 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
set gives the order of approximation of the LBE to its continuous counterpart.
As considered by Higuera and Jimenez [45] in 1989, the main purpose
of LBM continues to be, nowadays, to solve hydrodynamic problems, but,
presently, the conception of accurate and stable LB equations enabling to
achieve high Reynolds ows and to describe uid systems under strong tem-
perature non-equilibrium and/or complex interactions are the main challenges
for this kinetic method. Instability issues are frequent in LB simulation and
increasing the stability ranges is a very actual subject in LBM research.
As mentioned above, LBGK equations are low-order approximations to the
continuous BGK equation and, so, their solutions suer from errors related to
uncontrolled high-order moments. Consequently, instability issues are to be
expected.
Therefore, considering the nature of these issues, three main alternatives
appear at hand for enlarging the stability ranges without the help of free pa-
rameters, ad-hoc assumptions and/or an optimization step. The rst one is to
use lattices with increased dimensionality, i.e., based on high-order LB equa-
tions with increased velocity sets. This alternative was investigated by Siebert
et al. [85]. The second alternative is to add high order Hermite polynomial
tensors to the equilibrium distribution, trying to reduce the eect of their re-
lated moments on stability. This alternative was, also, utilized by Ref. [85] for
improving the stability of the D2Q9 LBE.
The third alternative is to work with moments-based LB equations, i.e., to
rewrite the LB equation in such a manner as to lter the undesirable ghost
moments from the numerical scheme. This alternative was investigated by
Ladd [55], Chen et al. [15], Zhang et al. [91] and by Latt and Chopard [58] for
the D2Q9 model and was very recently compared with the entropic stabilizer
method of Ref. [50] by Mattila et al. [66]. Although it is also a moments based
method, it diers from MRT because instead of being attributed with a given
relaxation time, all the ghost moments are ltered,
In this chapter, we present a method based on Mattila et al. [65] for
dealing with general ow problems which require high-order LBE, based on a
systematic improvement of the standard regularization idea.
In this method, for each Hermitian representation, both the equilibrium,
f eq , and non-equilibrium, f neq , parts are expanded on the nite Hermitian
basis of the subspace Hq of H.
Nevertheless, the high-order non-equilibrium moments do not t into the
Hermitian representation and, consequently, are sources of errors.This is the
case of the viscous stress tensor ταβ , which requires a full third order Hermitian
representation and is the source of O u3 errors in the macroscopic equations
when a second order LBE such as the D2Q9 or D3Q27 is used for the ow
simulation. This is also the case of the heat ow vector qα , which requires a
full fourth order Hermitian representation.
In Mattila et al. proposal, [65], the expansion of the non-equilibrium distri-
bution is carried out on a pair with the equilibrium distribution, but the diu-
sive parts are retained only up to the last physically relevant non-equilibrium
moment. Beyond that, the diusive parts are ltered out which leads to spe-
7.2. KINETIC PROJECTIONS 133
∂t f + ξ·∇x f = Ω, (7.1)
134 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
may be considered, for each (x, t) as a map from the continuous velocity space
ξD onto the space R of real numbers (Figure 7.1). In fact, it belongs to the
Hilbert space H of square integrable functions f : ξ D −→ R and may be
written in terms of an orthogonal basis of H that will be considered as the
innite set of Hermite polynomial tensors,
X a(θ) (x, t)
f (x, ξ, t) = H (θ) (ξ0 ) , (7.2)
nx !ny !nz !
θ
where θ indicates the order of tensor H (θ) , nα is the number of times the index
α = x, y or z appears repeated, nx + ny + nz = θ, ξ0 = ξ/ξ is a dimension-
1
less velocity, ξ = (kT0 /m) 2 is a mean microscopic speed, k is the Boltzmann
constant, T0 is a reference temperature and m is the mass of each particle.
Using the more common components notation for the Hermite polynomials
(1) (N )
H (0) , Hα0 , ..., Hα0 α1 ...αN , where αi = x, y or z, the rst Hermite polynomial
tensors are,
H (0) = 1,
Hα(1) = ξ0,α ,
(2)
Hαβ = ξ0,α ξ0,β − δαβ ,
(3)
Hαβγ = ξ0,α ξ0,β ξ0,γ (7.3)
n
(n−1)
X
Hα(n+1)
0 α1 ...αn
= ξα0 Hα(n)
1 ...αn
− δα0 αk Hα1 ...αn /αk , (7.4)
k=1
where the notation α1 ...αn /αk means that for a given k, only the indices α1 ,
..., αk−1 , αk+1 , ..., αn are taken into account.
The Boltzmann equation, Eq. (7.1) has an equilibrium solution, the Maxwell-
Boltzmann (MB) distribution [67], [5], [7],
(ξ0 −u0 )2
eq 1 e− 2(Θ+1)
f =n D/2
, D
D/2
(Θ + 1)
(2π) ξ
where n is the number density of particles, D is the Euclidean dimension of the
velocity space, D = 1, 2 or 3 and u0 = u/ξ is a dimensionless local velocity,
and Θ = T /T0 − 1 is the relative deviation from the reference temperature T0 .
7.2. KINETIC PROJECTIONS 135
In its simplest form, the interaction term Ω in Eq. (7.1) is modeled with a
BGK single relaxation term [4]
f − f eq
Ω=− , (7.5)
τ
where τ is a relaxation time.
Now, recall that f eq belongs to the Hilbert space H and its dimensionless
eq
form f may be developed in terms of the Hermitian basis of H as
2
ξ0
D
eq f eq ξ ∗ e− 2
f = =ρ D/2
no (2π)
X a(θ)
eq (u0 , Θ)
× H (θ) (ξ0 ) ,
nx !ny !nz !
θ
a(θ)
eq (u0 , Θ)
(ξ0 −u0 )2
1 e− 2(Θ+1)
= D/2 D/2
Hn(θ)
x ,ny ,nz
(ξ0 ) dξ0 .
(2π) (Θ + 1)
a(0)
eq = 1,
a(1)
eq,α = u0,α ,
(2)
aeq,αβ = (u0,α u0,β + Θδαβ ) ,
(3) u0,α u0,β u0,γ
aeq,αβγ = , (7.6)
+Θ (δβγ u0,α + δαγ u0,β + δαβ u0,γ )
(4)
aeq,αβγδ = u0,α u0,β u0,γ u0,δ
δαβ u0,γ u0,δ + δαγ u0,β u0,δ
+ Θ +δαδ u0,β u0,γ + δβγ u0,α u0,δ
+δβδ u0,α u0,γ + δγδ u0,α u0,β
+ Θ2 (δαβ δγδ + δαγ δβδ + δαδ δβγ ) .
2
ξ0
D
fξ e− 2
f= = D/2
n0 (2π)
X a(θ) (x, t)
× H (θ) (ξ0 ) . (7.7)
(θ)
n x !n y !n z !
H ∈Hq
H (θ) ∗ H (η) = H (θ) ∗ H (η) , (7.9)
d c
between any two Hermite polynomials H (θ) that form the basis of Hq , where
H (θ) ∗ H (η)
c
2
1 ξ0
= D e− 2 H (θ) (ξ0 ) ∗ H (η) (ξ0 ) dξ0 , (7.10)
(2π) 2
7.3. INCREASING THE ORDER OF APPROXIMATION 137
H (θ) ∗ H (η)
d
b −1
nX
= Wi H (θ) (as ei ) ∗ H (η) (as ei ) . (7.11)
i=0
b −1
nX h i2
Wi H (θ) (as ei )
i=0
2 i2
1 ξ0
h
= D/2
e− 2 H (θ) (ξ0 ) dξ0 . (7.12)
(2π)
2 ∗
" #
eq ∗
1+
as uα eiα +
f i,q=6 = Wi ρ a4s , (7.13)
2 u∗α u∗β eiα eiβ − 1
a2s δαβ
Figure 7.3: Hermitian basis of the Hilbert subspaces Hq=6 in (a) and Hq=9 in
(b), related to the D2Q9 LBE.
(θ)
where, as usual, the indexes nx and ny in Hnx ,ny indicates, respectively, the
number of times the index x and y appear repeatedly in the polynomial tensor,
with nx + ny = θ .
Recall [75], [84], that the norm preservation condition of these polynomial
tensors in the discrete space
b −1
nX 2
Wi Hn(θ)
x ,ny
(as ei )
i=0
2 2
1 ξ0
= D/2
e− 2 Hn(θ)
x ,ny
(ξ0 ) dξ0 = nx !ny !, (7.15)
(2π)
2 eq(θ)
eq
X anx ,ny (θ)
fi,q=6 = Wi H (as ei ) (7.16)
nx !ny ! nx ,ny
θ=0
where
7.3. INCREASING THE ORDER OF APPROXIMATION 139
1 (ξ0 −u0 )2 nx ny
aeq(θ)
nx ,ny = D/2
e− 2 Hn(θ)
x ,ny
(ξ0 ) dξ0 = ρ∗ u∗x u∗y aθs .
(2π)
(3) (3) (4)
Nevertheless, two third order H2,1 and H1,2 and one fourth order H2,2
polynomial tensors have their norm, Eq. (7.15) unaltered when the D2Q9
weigths Wi and scaling factor as are used. This allows the inclusion of the
related moments in the equilibrium distribution, which takes on the following
modied form
aeq(θ)
P2 nx ,ny (θ)
θ=0 nx !ny ! Hnx ,ny (as ei )
eq eq(3) eq(3)
fi,q=9 = Wi ρ∗ + a2,1 H (3) (as ei ) + a1,2 H (3) (as ei ) . (7.17)
2!1! 2,1 1!2! 1,2
eq(4)
a2,2 (4)
+ 2!2! H2,2 (as ei )
This full set of q=9 Hermite polynomial tensors is show in Figure 7.3.b
and can be understood [17], as the cartesian product of the sets
n o
(0) (1) (2)
H0 (as eix ) , H1 (as eix ) , H2 (as eix )
and
n o
(0) (1) (2)
H0 (as eiy ) , H1 (as eiy ) , H2 (as eiy ) .
Consider, now, the more usual Hermitian second order basis of the D2Q9
LBE, composed by the rst 6 two-dimensional Hermite polynomial tensors
(1) (1) (2) (2) (2)
H (0) , Hx , Hy , Hxx , Hxy and Hyy (Figure 7.3.a).
The inner product of ξβ ξγ by the second term on the l.h.s. of Eq. (6.33)
(3) (3) (3) (3)
requires the full set Hxxx , Hxxy , Hxyy and Hyyy , to belong to Hq . Since the
140 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
D2Q9 is a nine velocity set, the best it is possible to do is to add the polynomials
(3) (3) (4)
Hxxy , Hxyy and Hxxyy to the Hq subspace, because the discretization condition
given by Eq. (7.12),
b −1
nX h i2
Wi H (θ) (as ei ) = nx !ny !,
i=0
√
is satised by all these added polynomials when 3, W0 = 49 , W1 = 19
as =
1
and W2 =
36 (see Refs. [75], [85] for further details)
From Figure 7.3.b it is clearly seen that τxy is the only non-equilibrium
second-order moment that can be retrieved from the Hermitian basis of the
(3) (3)
subspace Hq=9 because Hxxy and Hxyy belong to the Hermitian basis of Hq=9 .
(3) (0)
The moments τxx and τyy are also dependent on the inner products Hxxx ∗ f
(3) (0) (3)
and Hyyy ∗ f , respectively and these inner products are zero because Hxxx
(3)
and Hyyy do not belong to the Hermitian basis of Hq=9 . So, τxx and τyy are
3
retrieved with errors O M a .
Therefore, the D2Q9 LBGK is a second order approximation to the con-
tinuous BGK equation and will retrieve the momentum balance equation with
third order errors O M a3 in both the two versions of the Hermitian basis,
Hq=6 and Hq=9 , restricting this LBE to be used for isothermal, nearly incom-
pressible ows. This order of approximation is given in Figure 7.3.a and b, by
the last row with a full set of Hermite polynomial tensors.
Nevertheless, in the rst version, all second order non-equilibrium moments
are ghost, in the sense that they cannot be represented in terms of the Her-
mitian basis of the subspace Hq=6 , while, in the second version, the viscous
stress tensor τxy can be represented in terms of the spatial derivatives of the
(3) (3)
equilibrium moments aeq,xxy and aeq,xyy .
It is, here, important to note that the above reasoning leads to the con-
clusion that contrary to what is usually believed, the O M a3 errors that
characterize the D2Q9 LBE and, also, the D3Q27 are not due to the dis-
cretization process itself, but to a poor second order Hermitian representation.
Indeed, this conclusion reects the dependence of the discrete velocity set on the
order of approximation of the Hermitian representation (and not the inverse ).
Better representations can be built without changing the set of discrete
velocities and the LBE parameters as and Wi . So, more accurate and stable
eq
LBE can be achieved by only replacing the equilibrium distribution fi , Eq.
(7.13), by its expansion in terms of the full set of 9 Hermitian polynomial
tensors shown in Figure 7.3.b.
The question that remains to be answered is whether the inclusion of these
third-order Hermite polynomials has any eect on the LBE stability.
The D2Q9 stability was analyzed by comparing the D2Q9 LBGK with the
Hq=6 Hermitian representation, Eq. (7.16), and a Hermitian representation
(3) (3)
where the third-order polynomials Hxxy and Hxyy were included into the
equilibrium distribution, Siebert et al. [85]. These models were also compared
with the Lallemand and Luo [56], multiple relaxation time MRT model, since
7.3. INCREASING THE ORDER OF APPROXIMATION 141
Figure 7.4: The eect on stability of adding third order Hermite polynomials to
the second order equilibrium distribution in a D2Q9 LBGK [85]. Comparison
with the MRT model of Ref. [56]
this model was also built with the aim of improving the LBE stability.
The present analysis can be found in Figure 7.4 and is focused on values
1
of the relaxation time very close to its singular limit
2 . The abscissa was
∗
chosen as 1/τ in order to compare with previous results from Ref. [56]. It can
be observed that both the second and the third-order LBGK models present a
homogeneous decrease in the local speed stability limit when the relaxation time
approaches 1/2, whereas this limit remains insensitive to the τ∗ variation in the
∗
MRT model, up to τ =0.50251. Although the results presented by Lallemand
and Luo are not related to quasi-incompressible models, as the present LBGK
ones are, these authors attributed the better performance of the MRT model
with respect to the second-order LBGK to the use of high frequency relaxation
terms in modeling the collision term.
142 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
Figure 7.4 shows, nevertheless, that the third-order LBGK model has a
considerably better performance when compared with the second-order one
and with the MRT model in what concerns its stability limits. So, the addition
of third-order velocity polynomials largely improve the stability range and this
improvement is due to the equilibrium distribution representation itself and
not to the use of extra relaxation terms in the collision model. This is an
important conclusion, since it avoids the use of MRT dispersion relations for
the adjustable parametersrelated to the short wavelength nonhydrodynamic
momentsto increase numerical stability.
X a(θ) (x, t)
f i = Wi H (θ) (as ei ) , (7.18)
nx !ny !nz !
θ
X
f i H (η) (as ei ) = a(η) .
i
because
For isothermal second-order models, the Hermitian moments that are needed
are
X
a(0) = f i = ρ∗ ,
i
X
a(1)
α = f i as eiα = as ρ∗ u∗α ,
i
(2)
X
f i a2s eiα eiβ − δαβ = a2s ρ∗ u∗α u∗β + ταβ
∗
aαβ = .
i
1 (2) 2
fbi = Wi a(0) + a(1)
α a e
s iα + a a e e
iα iβ − δ αβ ,
2 αβ s
or
1 1
fbi = Wi ρ∗ + 2 ρ∗ u∗α eiα + 4 ρ∗ u∗α u∗β + ταβ
∗
eiα eiβ − c2s δαβ .
(7.19)
cs 2cs
Although fbi does not alter the mass and momentum of site x, satisfying
144 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
X Xb
fi = f i = ρ∗ ,
i i
X Xb
f i eiα = f i eiα = ρ∗ u∗α ,
i i
X Xb
f i eiα eiβ = f i eiα eiβ = P ∗ δαβ + ρ∗ u∗α u∗β + ταβ
∗
.
i i
these reconstructed populations dier from the populations fi that were origi-
nally propagated to site x. Some authors attribute this dierence to the noise
produced by numerical errors due to non-controlled higher-order moments. In
this case, the reconstruction of fbi based on its rst momentsthe ones that t
into the LBE Hermitian representation may be seen as a regularization proce-
dure, performed for each site and at each time step [58]. So, when we rewrite
the populations as fbi ρ∗ , u∗α , ταβ
∗
as in Eq. (7.19), we are implicitly requiring
(θ)
that the inner products f ∗ H = 0 for every θ > 2. In other words, we are
b
i
out 1 eq 1 b
fi (x, t + δ) = f (x, t) + 1 − f i (x, t) . (7.20)
τ∗ i τ∗
eq
where fi is given by
eq 1 1
f i = Wi ρ∗ + 2 ρ∗ u∗α eiα + 4 ρ∗ u∗α u∗β eiα eiα − c2s δαβ
cs 2cs
The main idea behind this method is to lter the high-order non-equilibrium
moments that cannot be represented in terms of the Hermitian basis of the
Hilbert subspace Hq and so, are sources of errors aecting both the accu-
racy and the stability of the method. This idea was investigated by Latt and
Chopard [58] for the second-order D2Q9 LBE, who observed for this method,
an stability range 7.7 times the stability range of the BGK numerical scheme,
in the simulation of lid driven cavity ows.
It is, perhaps, important to emphasize that the distribution given by Eq.
(7.19) is the projection of fi on the subspace Hq=6 and that all the calculations
are based on this projection at each time step. So all the spurious information
7.4. MOMENTS-BASED METHODS 145
The initial conditions are shown in Figure 7.8 and are given by
146 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
y
∗ 1
L
uo tanh λ L − 4 for |y| ≤ 4
u∗x = ,
y
u∗o tanh λ 3 L
− for |y| >
4 L 4
and
x 1
u∗y = εu∗o sin 2π + .
L 4
In the above equationsu∗o is the velocity of the shear layer. The prole
is controlled by parameter λ, which was considered in present sample case as
λ = 80. With this value of λ, the velocity prole, although continuous, is nearly
constant in the central layer and has a step gradient in the contact lines with
the lateral layers, producing a shear ow. An initial sinusoidal velocity pertur-
bation u∗y is imposed in the vertical direction and controlled by parameter ε.
Parameter ε = 0.05 in present simulations. Boundary conditions are periodic.
The simulation starts from these imposed velocity proles and follows the ve-
locity eld till the end of the ow, when all the kinetic energy is dissipated by
viscous forces.
The Mach number, M = u∗ /cs was kept very small (M = 5.4127 × 10−2
) and the simulations were performed with the second-order D2Q9 LBE, in
despite of its errors. At each time step the average kinetic energy is measured
2
and normalized with respect to ρu∗o /2. Fig. 7.9 shows the time decay of the
average kinetic energy comparing the simulation results between the LBGK
model, given by Eq. (7.5) with the moments-based LBE, Eq. (7.19) , Mattila
et al. [66]. The initial Reynolds number is Re = 32000, the numerical domain
has a length L = 128 lattice unities, and velocity u∗0 =1/32. The dimensionless
∗
kinematic viscosity is very low ν = 4/30000. The plotted average values of
∗2
the kinetic energy are normalized with respect to ρuo /2. It is clearly seen that
∗
the LBGK scheme becomes unstable very early, around time t = 0.3, corre-
sponding to the rst 1230 time steps, whreas the moments-based LB scheme
remain stable during the whole simulation.
7.4. MOMENTS-BASED METHODS 147
Figure 7.8: Initial velocity proles used for the simulation of the double periodic
shear ow.
Figure 7.9: Decay of average kinetic energy for an initial Reynolds number,
Re = 32000, L = 128 lattice unities, u∗0 =1/32 and ν ∗ = 4/30000. The average
∗2
values are normalized with respect to ρuo /2. The LBGK scheme becomes
unstable early on, around time t∗ = 0.3, corresponding to 1230 time steps.
The moments-based LB scheme remain stable during the whole simulation,
[66]
148 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
Figure 7.10: Vorticity eld for t∗ = 1 showing the simulation results for the
LBGK model (left) and the moments-based method (right). It is clearly seen
that the LBGK model is plagued by spurious secondary vortex even for Re =
32000 whereas the moments-based scheme produces accurate and stable results
up to Re = 80000 [66]
Being a global measure, the average kinetic energy, does not capture well
the dierences between local ow features. Fig. 7.10 shows the simulated
t∗ = 1, comparing the LBGK and the moments-based scheme
vorticity eld, at
at this time step. Now Re = 80000 for a simulation domain with L = 256
∗
and a dimensionless velocity u0 =1/16. The dimensionless kinematic viscosity
∗ −4
is ν = 2 × 10 . The LBKG case was simulated with Re = 32000 since
beyond this limit the simulation was unstable. Simulation with LBGK shows
small spurious secondary vortexes. The moments-based LB scheme remained
stable during the whole simulation and was shown to be accurate without the
appearance of unphysical secondary vortex.
∗
X neq
παβγ = fi eiα eiβ eiγ . (7.21)
i
∗ ∗
παβγ =π
eαβγ + πadv,αβγ ,
∗
where π
eαβγ is the diusive component
∗
X neq
(eiα − u∗α ) eiβ − u∗β eiγ − u∗γ ,
π
eαβγ = fi
i
∗
πadv,αβγ = u∗α τβγ
∗
+ u∗β ταγ
∗
+ u∗γ ταβ
∗
.
A Chapman-Enskog analysis shows that, in isothermal problems, the vis-
cous stress tensor ταβ is the highest non-equilibrium moment that is needed for
closing the system of hydrodynamic equations at the Navier-Stokes level. In
isothermal problems, the appropriate truncation point for the Hermite polyno-
mial expansion is at N = 3, when the M a3 errors are, also, to be avoided. That
is, for moments up to third-order, both equilibrium and non-equilibrium part,
should be included into the expansion. However, as becomes evident from this
analysis, the diusive component of the third-order non-equilibrium moment
is dened in terms of fourth-order equilibrium moments which do not belong
∗
to the chosen Hermite basis. Hence this third-order diusive component π
eαβγ
must be considered as a spurious or ghost moment, and thus should be
ltered from the representation in isothermal hydrodynamics, resulting in the
following recurrence relation
∗
παβγ = u∗α τβγ
∗
+ u∗β ταγ
∗
+ u∗γ ταβ
∗
,
150 CHAPTER 7. BEYOND BGK: MOMENTS-BASED METHODS
c0 = ξ0 − u0
c2
0
eq ρ∗ e− 2(Θ+1)
f = D D , (7.23)
(Θ + 1) 2 π2
c2
− 0 (θ)
∗e aeq
eq 2 X
f =ρ D
e e (θ) ,
H
π 2 nx !ny !nz !
θ
c2
0
1 e− 2(Θ+1) e (θ)
a(θ)
eeq = D D H (c0 ) dc0 .
(2π) 2 (Θ + 1) 2
Since the Hermite polynomials are written in terms of the peculiar velocity
c0 and considering Eq. (7.23):
c2
0
1 e− 2(Θ+1)
a(0)
eeq = D D 1dc0
(2π) 2 (Θ + 1) 2
1 c2
= D e− 2 dc = 1,
(2π) 2
p
(Θ + 1) c2
a(1)
eeq,α = D e− 2 cα dc = 0,
(2π) 2
(2) 1 c2
aeq,αβ
e = D e− 2 (cα cβ Θ + cα cβ − δαβ ) dc =Θδαβ
(2π) 2
an so on.
Consider now the non-equilibrium distribution and its expansion in Hermite
polynomials e (θ) ,
H
c2
0 (θ)
neq e− 2 X aneq
e e (θ) ,
f = D H (7.24)
π 2 nx !ny !nz !
θ
7.5. TOWARD HIGH-ORDER MOMENTS-BASED LBE: RECURRENCE RELATION
(θ)
where parameters aneq
e are given by
neq
a(θ)
eneq = f e (θ) dc0 .
H
(n−1)
X
e α(n)...α (c0 ) = Hα(n)...α (ξ0 ) −
H u0,αθ Hα1 ...αn /αθ (ξ0 )
1 n 1 n
θ
(n−2)
X
+ u0,αθ u0,αη Hα1 ...αn /αθ αη (ξ0 )
θ,η
n
+ ... + (−1) u0,α1 ...u0,αn H (0) (ξ0 ) , (7.25)
where the notation α1 ...αn /αθ ...αη means that all the indices are considered
except the indices αθ ... αη . For each rth term in the sum, r = 1, ..., n + 1, the
n
above equation considers all the possible Cr−1 combinations of the n indices
{α1 , ..., αn } taken as sets of r − 1 elements. For instance, if n = 4, the third
4
term on the right hand side of Eq. (7.25) will be a sum of C2 = 6 terms,
Eq. (7.25) gives rise to a general relationship for the non-equilibrium mo-
neq
ments. By multiplying Eq. (7.25) by f and integrating in the velocity space
(n−1)
X
a(n)
e (n)
neq,α1 ...αn = aneq,α1 ...αn − u0,αθ aneq,α1 ...αn /αθ
θ
(n−2)
X
+ u0,αθ u0,αη aneq,α1 ...αn /αθ αη
θ,η
n
+ ... + (−1) u0,α1 ...u0,αn a(0)
neq . (7.26)
a(0) a(0)
neq = eneq = 0,
a(1) a(1)
neq,α = eneq,α = 0,
(2) (2) 1 ∗
aneq,αβ = e
aneq,αβ = τ ,
c2s αβ
∗ ∗
uα τβγ + u∗β ταγ
∗
(3) (3) 1
aneq,αβγ = e
aneq,αβγ + 3 , (7.27)
cs +u∗γ ταβ
∗
(4) (4)
aneq,αβγδ = e
aneq,αβγδ
!
(3) (3)
1 u∗α aneq,βγδ + u∗β aneq,αγδ
+ (3) (3)
cs +u∗γ aneq,αβδ + u∗δ aneq,αβγ
u∗α u∗β τγδ
∗ ∗
+ u∗α u∗γ τβδ
1
− 4 +u∗α u∗δ τβγ∗
+ u∗γ u∗δ ταβ
∗ .
cs ∗ ∗ ∗ ∗ ∗ ∗
+uβ uδ παγ + uβ uγ παδ
(3) 1
u∗α τβγ
∗
+ u∗β ταγ
∗
+ u∗γ ταβ
∗
aneq,αβγ = 3
, (7.28)
cs
∗ ∗ ∗
u u τ + u∗α u∗γ τβδ ∗
(4) 1 α∗ β∗ γδ∗
aneq,αβγδ = 4 +uα uδ τβγ + u∗β u∗γ ταδ
∗
. (7.29)
cs ∗ ∗ ∗ ∗ ∗ ∗
+uβ uδ ταγ + uγ uδ ταβ
Eqs. (7.28) and (7.29) are recurrence relations for the third and fourth
order non-equilibrium moments in terms of the second order viscous stress
tensor. Malaspinas [63] found similar relations, but using a dierent approach
and not considering the role of the diusive component of these moments.
1 In the lattice Boltzmann framework, the Knudsen number is related to the ratio of
the lattice spacing h and a characteristic length L. It is a nite number and cannot be
considered as close to zero. So, errors due to high-order Knudsen terms are to be expected in
the hydrodynamic equations. The inuence of nite Knudsen numbers on the error formation
in isothermal hydrodynamics is, nevertheless, outside the scope of this paper.
7.6. MOMENTS-BASED LBE WITH RECURRENCE RELATIONS 153
eq
f i (x, t)
(θ)
X aeq
= Wi H (θ) (as ei ) , (7.30)
nx !ny !nz !
H (θ) ∈Hq
and
f i (x, t)
X a(θ)
= Wi H (θ) (as ei ) . (7.31)
nx !ny !nz !
H (θ) ∈Hq
∗ (2) ∗ (2)
a(3)
neq,xxy = uy aneq,xx + 2ux aneq,xy ,
∗ (2) ∗ (2)
a(3)
neq,xyy = 2uy aneq,xy + ux aneq,yy ,
2
∗ (2) ∗ ∗ (2) ∗2 (2)
a(4)
neq,yyy = ux aneq,yy + 2ux uy aneq,xy + uy aneq,yy .
1 ∗ 1 ∗ ∗
2
1+ c2s uα eiα + 2c4s uα uβ eiα eiα − cs δαβ
27 ∗2 ∗ 1
2
+ 2 ux uy eix − 3 eiy
eq
f i = W i ρ∗ , (7.32)
∗2 ∗
+ 27 1
2
2 uy ux eiy − 3 eix
∗2 ∗2
+ 81 e2ix − 13 e2iy − 31
4 ux uy
and
fˆ i = f i
eq
9 ∗ 1
2 ταβ eiα eiβ − 3 δ2αβ
∗ ∗ ∗ ∗
+ 92 1
uy τxx + 2ux τxy eix − 3 eiy
+ 92 u∗x τyy
∗
+ 2u∗y τxy
∗
e2iy − 1
3 eix
+ Wi (7.33)
∗2 ∗
ux τyy +
2u∗x u∗y τxy∗
27
e2ix − 13 1
2
+4 eiy − 3
2
+u∗y τxx ∗
Eqs. (7.32) and (7.33) are to be compared with the standard moments-
based D2Q9 LBE. This new form of the D2Q9 moments-based LBE, also re-
trieves the isothermal Navier-Stokes equations with errors O M a3 and, as
commented above, is also restricted to very low Mach number ows, in the in-
compressible limit. The addition of third and fourth order ltered terms in the
non-equilibrium distribution has the purpose of enhancing its stability limits,
for high Reynolds number ows, without the interference of third and higher
order diusive ghost moments.
Figure 7.11: Third-order Hermitian basis of the subspace Hq=10 and its related
D2V17 set of lattice vectors.
∗ (2)
a(3)
neq,xxx = 3ux aneq,xx ,
∗ (2) ∗ (2)
a(3)
neq,xxy = uy aneq,xx + 2ux aneq,xy , (7.34)
a(3)
neq,xyy = 2u∗y a(2)
neq,xy + u∗x a(2)
neq,yy ,
a(3)
neq,yyy = 3u∗y a(2)
neq,yy .
eq
fi
1 + c12 u∗α eiα
s
+ 1 u∗ u∗ e e − c2s δαβ
= Wi ρ∗ 2c4s α β iα iβ
u∗ ∗ ∗
eiα eiβ eiγ ,
α uβ uγ
+ 6c6s −c2s (δβγ eiα + δαγ eiβ + δαβ eiγ )
fˆ i = f i
eq
∗
ταβ
e e − c2s δαβ
2c4s iα iβ
+ Wi
+ 6c16 u∗α τβγ∗
+ u∗β ταγ
∗
+ u∗γ ταβ
∗ .
s
× eiα eiβ eiγ − c2s (δβγ eiα + δαγ eiβ + δαβ eiγ )
31 neq (3)
qα = ρ0 ξ f H e
αββ (c0 ) dc0
2
3 1 (3)
= ρ0 ξ e a .
2 neq,αββ
2
Remembering that ξ = c2 c2s , in its dimensionless form the heat ux qα∗
(3)
= qα / ρ0 c3 can be aneq,αββ in its discrete form by,
related to e
(3) 2qα∗
aneq,αββ =
c3s
e
1 X neq 2
= 3 f eiβ − u∗β (eiα − u∗α ) . (7.35)
cs i i
(3)
Therefore, from Eq. (7.27) the sum aneq,αββ is related to the heat ux qα∗
by
(3) qα∗ 1
+ 3 u∗α τββ
∗
+ 2u∗β ταβ
∗
aneq,αββ = 2 3
.
cs cs
Third-order diusive moments require fourth-order Hermite representations.
th
Figure 7.12 shows a 4 -order Hermitian representation and its corresponding
D2V37 LBE.
(4)
The difusive moments aneq,αβγ
e are to be ltered from the representation
(4)
and each moment aneq,αβγδ is calculated as in Eq. (7.27), in terms of the
7.6. MOMENTS-BASED LBE WITH RECURRENCE RELATIONS 157
(3) (3)
aneq,αβγ . The peculiar moment aneq,αβγ
e is calculated at each site and time
step in accordance with
(3) 1 X neq
f i (eiα − u∗α ) eiβ − u∗β eiγ − u∗γ ,
aneq,αβγ =
e 3
cs i
related to the heat ow vector. In two dimensions and in accordance with Eq.
(3) (3)
(7.35) the sum c3s eaneq,αxx + e
aneq,αyy is twice the heat ux component along
In this regard, the lattice Boltzmann equation (LBE) has been successfully
used in the last decades as a computational uid dynamics tool, solving a
discrete analogue of the Boltzmann equation for the mass, momentum and
energy balance equations ([52], [86], [35]) . The LBE has proved to be valuable
in solving simple and complex ows.
A lot of eort has been devoted to the boundary conditions for the lattice
Boltzmann framework, but the most recent thorough comparison of dierent
models for boundary conditions was published a decade ago [58], [62]. In this
work, we focus the extension of moments-based populations to boundary con-
ditions (BC's). Schemes based on moments-based method generally leads to
improved general stability of LB equations, but, as we will see in this chapter,
there is no single solution to solve the boundary condition problem.
159
160 CHAPTER 8. BOUNDARY CONDITIONS: PART II
X
f¯i eiy = ρ∗ u∗y .
i
where u∗x and u∗y are the dimensionless velocity components for x = xb , that
are supposed to be known. A third equation can be written as
X
f¯i = ρ∗ , (8.1)
i
giving the mass density ρ∗ at the boundary node, which is here to be reinter-
preted as the pressure since P ∗ = ρ∗ c2s in the LB incompressible limit. For the
node depicted in Figure 8.1, we have thus 3 equations for the unknowns ρ∗ , f¯4
, f¯7 , f¯8 . This is the main diculty for establishing local boundary conditions
in LB framework since the number of unknowns is greater than the number
of equations at our disposal. This is especially important when we deal with
high-order LB equations, with a great number of discrete lattice vectors [75],
[84], [82].
In addition, although the collision process does not aect the mass density
at the boundary node and we can always assure that
X X
f¯i = f¯iout , (8.2)
i i
8.1. MASS CONSERVATION 161
there is no guarantee that that there is no mass-leakage, i.e., that the amount
of particles that reach the boundary node from the uid phase is the same
amount that leaves this node back to the uid phase, after collision.
To be clear, let us consider the set
I = {0, 1, 2, 3, 5, 6} ,
in Figure 8.1, as the set of incoming directions from the uid phase to the solid
surface and that
O = {0, 1, 3, 4, 7, 8} ,
is the set of outgoing directions from the solid to the uid sites. For avoiding
loss or gain of particles to or from the external environment, instead of Eq.
(8.2) the mass preservation condition requires that the total amount of particles
incoming from the uid phase before collision be the same as the total amount
of particles leaving the solid after collision. In other words,
X X
f¯i = f¯iout . (8.3)
i∈I i∈O
X X X X
f¯i + f¯i = f¯iout + f¯iout , (8.4)
i∈I i∈I
/ i∈O i∈O
/
X X
f¯i = f¯iout . (8.5)
i∈I
/ i∈O
/
For the boundary node shown in Figure 8.1, the above equation reduces to
meaning that the amount of particles entering the uid before the collision step
must compensate the amount of particles leaving the uid after the collision
step. This is necessary for avoiding gain or loss of particles by the uid phase.
It should be remarked that the bounce-back boundary condition shown in
Section ........... for solid surfaces at rest,
f −i (xb , t + δ) = f i (xb , t) ,
also avoids the mass leaking, by making
f¯4 (xb , t + δ) = f¯2 (xb , t) , f¯7 (xb , t + δ) = f¯5 (xb , t) , f¯8 (xb , t + δ) = f¯6 (xb , t) .
Nevertheless, the above condition must not be confused with the mass
preservation condition given by Eq. (8.6). While Eq. (8.6) gives post-collisional
populations that are dependent on the relaxation time, BB condition gives post-
collisional populations f¯i (xb , t + δ) independent on it.
162 CHAPTER 8. BOUNDARY CONDITIONS: PART II
X X
f¯i = f¯iout , (8.7)
i∈I i∈O
X
f¯i eix = ρ∗ u∗x , (8.8)
i
X
f¯i eiy = ρ∗ u∗y , (8.9)
i
X 2
f¯i eix eix = ρ∗ u∗x + ρ∗ c2s + τxx
∗
, (8.10)
i
X
f¯i eix eiy = ρ∗ u∗x u∗y + τxy
∗
, (8.11)
i
X 2
f¯i eiy eiy = ρ∗ u∗y + ρ∗ c2s + τyy
∗
, (8.12)
i
fˆ i = f i = Wi
h i
ρ∗ + 1 ∗ ∗
c2s ρ uα eiα + 1
2c4s ρ∗ u∗α u∗β + ταβ
∗
eiα eiβ − c2s δαβ .
Therefore, the system of equations (8.7)-(8.12) does not have a single so-
lution and dierent models were proposed for solving this boundary condition
problem
fˆ
¯neq τ ∗ , τ ∗ , τ ∗ = f¯2 − f¯eq (ρ∗ ) ,
4 xx xy yy 2
fˆ
¯neq τ ∗ , τ ∗ , τ ∗ = f¯5 − f¯eq (ρ∗ ) ,
7 xx xy yy 5
8.2. BOUNDARY NODES ON PLANE SURFACES 163
fˆ
¯neq τ ∗ , τ ∗ , τ ∗ = f¯6 − f¯eq (ρ∗ ) ,
8 xx xy yy 6
∗ ∗ ∗
and the following expressions are obtained for τxx , τxy , τyy ,when cs is replaced
√
by 1/ 3
∗
= 4 f¯5 + f¯6 − 2f¯2 − ρ∗ u∗x ,
τxx
1
∗
= 2 f¯5 − f¯6 − ∗
+ uy ρ∗ u∗x ,
τxy
3
1
∗
= 2 f¯2 + f¯5 + f¯6 − 1 + u∗y ρ∗ u∗y + ,
τyy
3
∗
where the density ρ can be evaluated as
1 ¯
ρ∗ = f0 + f¯1 + f¯3 + 2 f¯2 + f¯5 + f¯6 .
(8.13)
u∗y +1
as in Zou & He [93] model.
The unknown populations can be then calculated as
¯4 = f¯2 − 2 ρ∗ u∗ ,
fˆ y
3
¯7 = f¯5 − 1 ρ∗ u∗ + u∗ ,
fˆ
x y
6
¯8 = f¯6 + 1 ρ∗ u∗ − u∗ ,
fˆ
x y
6
The above equations give the following adherence conditions
X
f¯i eiy = ρ∗ u∗y , (8.14)
i
X 1
f¯i eix = ρ∗ u∗x , (8.15)
i
3
which can only be satised for the x-component when the known populations f¯1
and f¯3 are replaced by their moments-based forms fˆ¯1 and fˆ¯3 , whose dierence
gives
fˆ ¯3 = 2 ρ∗ u∗ .
¯1 − fˆ
x
3
Thus the main particularity of Latt et al. model is the replacement of the
known populations by reconstructed ones, based on the values of the viscous
stress tensor components. Nevertheless, Latt et al. model does not satisfy
Eq.(8.3) and, so, it is unable to prevent the leaking of particles across the
boundaries.
164 CHAPTER 8. BOUNDARY CONDITIONS: PART II
Exercise
Demonstrate Eqs. (8.14) and (8.15).
eq
f i = f i (ρ∗ ) + fˆ
¯neq τ ∗ , τ ∗ , τ ∗ , i ∈ I,
i xx xy yy
X
ρ∗I = f i = ρ∗I ρ∗ , τxx
∗ ∗ ∗
, τxy , τyy
i∈I
Therefore, for the boundary node depicted in Figure 8.1, the above equa-
tions give a system of 7 equations and 4 unknowns, because fi is know for each
i ∈ I. A minimization procedure is used for solving this overdetermined system
of equations.
X
f¯i eix = ρ∗ u∗x , (8.16)
i
X
f¯i eiy = ρ∗ u∗y , (8.17)
i
X 2
f¯i eix eix = ρ∗ u∗x + ρ∗ c2s , (8.18)
i
∗
and requiring that τxx = 0, because u∗x is constant along direction x.
ˆ
¯ ∗ 1 ∗ ∗ 1 ∗ ∗ ∗ ∗
2
fi = Wi ρ + 2 ρ uα eiα + 4 ρ uα uβ + ταβ eiα eiβ − cs δαβ .
cs 2cs
Since mass must be conserved and node (xb , yb ) is adjacent to a solid sur-
face, the sum of the moments-based populations of particles outgoing from the
boundary node to the uid sites, after the collision step, must be identical to
the sum of the populations incoming from uid sites before the collision step,
or
fˆ
X X
f¯i = ¯out .
i (8.19)
i∈I i∈O
So, the physics underlying Eq. (8.19) and Eq. (8.3) is the same. Never-
theless, in Eq. (8.19) all the outgoing post-collisional populations are kinetic
projections dependent on the dimensionless quantities ρ∗ , τxx
∗ ∗ ∗
, τxy , τyy in ac-
cordance with
fˆ
X X
f¯i eiα + ¯i eiα = ρ∗ u∗ ,
α (8.20)
i∈I i∈I
/
X Xˆ
f¯i eiα eiβ − c2s δαβ + f¯i eiα eiβ − c2s δαβ = ρ∗ u∗α u∗β + ταβ
∗
. (8.21)
i∈I i∈I
/
As before, there are dierent methods for solving the overdetermined system
of Eqs. (8.19)-(8.21) for the boundary node on plane walls such as the one
depicted in Figure 8.1, leading to dierent solutions. We present the results of
two of these methods presented in Bazarin et al. [ ? ]
∗ 5 1 ∗ ∗
1 ∗
ρ − u 1 + uy + (ω − 1) τyy = ρI . (8.22)
6 2 y 2
166 CHAPTER 8. BOUNDARY CONDITIONS: PART II
∗
On the other hand, Eq. (8.21), for τ22 , results in the following relationship
2 ∗ 1 2
τyy + ρ∗ 6u∗y + 3u∗y − 1 = (myy )I , (8.23)
3 9
ρI = iI fi and (ρmαβ )I = i f¯iI eiα eiβ − c2s δαβ
P ¯ P
where are known quan-
tities, since they are only dependent on the incoming populations that reachs
the boundary node, coming from the uid phase.
Eqs. (8.22) and (8.23) can be solved for the density ρ∗ and the normal
∗
component τyy of the viscous stress, giving
4ρI + 3 (myy )I (1 − ω)
ρ∗ = 3 ,
ω + 9 − 3uy (1 + ω) − 6ωu2y
∗
2ρI + 15 (myy )I − 3 2ρI + 3 (myy )I uy − 3 4ρI + 3 (myy )I u2y
τyy = .
ω + 9 − 3uy (1 + ω) − 6ωu2y
∗ 6 4ρI − 3 (myy )I (ω − 1)
τxx = (mxx )I − 3u2x ,
5 ω − 3uy (ω + 1) − 6ωu2y + 9
∗ ux 4ρI − 3 (myy )I (ω − 1) (3uy + 1)
τxy = 2 (mxy )I − .
ω − 3uy (ω + 1) − 6ωu2y + 9
As expected all the moments become dependent on the relaxation frequency
ω.
¯fˆ ρ∗ , τ ∗ , τ ∗ , τ ∗ .
X X
ρ∗ = f¯i +
i xx xy yy (8.24)
i∈I i∈I
/
For the node depicted in Figure 8.1 the above equation, Eq, (8.19) also
gives a relationship between the dimensionless mass density and the normal
component of the viscous stress
1 ∗ ∗ 2
1
ρ 3uy − 3u∗y + 5 − τyy ∗
= ρI ,
6 2
and Eq. (8.21) for α = β = 2 closes the system of equations for ρ∗ and
∗
τyy
1 ∗ ∗2 2
ρ 6uy + 3u∗y − 1 + τyy ∗
= (myy )I .
9 3
∗
Solving this system gives the following relationship for ρ ,
8.3. CORNER NODES 167
1
ρ∗ =
4ρI + 3 (myy )I .
3 1 + u∗y
The remaing second order moments can be written in terms of ρ∗ ,
∗ 3 1 2
τyy = (myy )I − ρ∗ 6u∗y + 3u∗y − 1 ,
2 6
∗ 1
= 2 (mxy )I − ρ∗ u∗x 3u∗y + 1 ,
τxy
3
∗ 6 2
τxx = (mxx )I − ρ∗ u∗x .
5
1
f¯5 − f¯7 = ρ∗ u∗x + u∗y ,
6
being identical only for the case when the corner boundary node is at rest,
u∗x = u∗y =0.
In Latt et al. [59] approach, Eq. (8.13) is useless for evaluating the density
at corner nodes. Therefore Latt et al. propose to extrapolate the density of
the next neighbors with a second order precision.
Extending the approach to concave non-slip corners, Mohammed and Reis
[71] propose a system of 6 equations to determine the 5 unknown populations
168 CHAPTER 8. BOUNDARY CONDITIONS: PART II
and density. The system is given by Eqs (8.7)-(8.12) considering the tensions
∗ ∗ ∗
τxx , τxy and τyy null.
For nding the unknown populations in Hegele et al model [42], the proce-
dure is exactly the same as presented in Section 8.2.4 for the plane surfaces.
We solve the conservation of mass Eq.(8.19) requiring
X bout X
fi = f¯i ,
i∈O i∈I
with I = {0, 2, 3, 6} and O = {0, 1, 4, 8} and Eq. (8.21) is used for the
∗ ∗ ∗
moments τxx , τxy and τyy .
The solution for ux = u∗y = 0
∗
is obtained as
12
ρ∗
= 3ρI + (1 − ω) 3 (myy )I + 3 (mxx )I + 7 (mxy )I
9ω + 16
∗ 2
τxx = 12ρI − 12 (myy )I (ω − 1) + 15 (mxx )I (ω + 4) + 10 (mxy )I (ω − 6)
3 (9ω + 16)
∗ 2
τxy = 42ρI − 15 (myy )I (ω − 6) − 15 (mxx )I (ω − 6) − 50 (mxy )I (2ω + 9)
9 (9ω + 16)
∗ 2
τyy = 12ρI + 15 (myy )I (ω + 4) − 12 (mxx )I (ω − 1) + 10 (mxy )I (ω − 6)
3 (9ω + 16)
Alternatively, when the density concept replaces the mass conservation con-
dition, we solve Eq. (8.24), instead of Eq. (8.19), and Eqs.(8.21). The solution
for u∗x = u∗y = 0 is, now, independent on the collision frequency and is given by
∗ 9 7
ρ = ρI + (myy )I + (mxx )I − (mxy )I
4 3
∗ 1
τxx = ρI + (myy )I + 5 (mxx )I − 5 (mxy )I
2
∗ 1 7
τxy = ρI + 5 (myy )I + 5 (mxx )I − 25 (mxy )I
4 3
∗ 1
τyy = ρI + 5 (myy )I + (mxx )I − 5 (mxy )I
2
This procedure is the same for convex corner sites and, therefore, can be
extended to staircase surfaces.
Exercises
1. For the boundary node shown in Fig. 8.2, nd the populations f¯1 , f¯4 ,
f¯8 , f¯5 and f¯7 and the density ρ ∗
by solving Eqs. (8.7)-(8.12), following the
method of Mohammed and Reis [71].
2. Using Hegele et al. [42] method, presented in Secion 8.2.4, nd the
expressions for the density ρ∗ and second order moments
∗
τxx ∗
, τxy and
∗
τyy
∗
when the corner boundary node is moving with a velocity ux 6= 0.
8.4. COMPARATIVE ANALYSIS 169
8.4.1 Accuracy
The accuracy of the presented methods was tested by Bazarin et al.[ ? ],
comparing the results obtained with moments-based methods with the classical
results of Ghia et al. [30], and the more recent benchmarks of Botella and
Peyret [8] and Erturk, Corke and Gökçöl [25]. Ghia et al. and Erturk et
170 CHAPTER 8. BOUNDARY CONDITIONS: PART II
Figure 8.4: Numerical results for Re=400, 1000 and 5000 showing the relative
velocity u∗x /u∗0 along the mid-line y ∗ = L/2. Results are compared with Ghia et
al. [30], Botella and Peyret [8] and Erturk, Corke and Gökçöl [25]. Continuous
lines are from Bazarin et al. [ ? ].
al. solved this problem using discretization schemes based on the vorticity-
stream formulation of the Navier-Stokes equations. The work of Botella and
Peyret uses a highly-accurate solution based on Chebyshev collocation method
for the spatial discretization in an assymptotic expansion of the Navier-Stokes
equations.
Comparing the velocity proles with other works, Figures 8.4 and 8.5 shows
for Re = 100, 400, 1000 and 5000 the numerical results for the relative velocity
u∗x /u∗0 along the mid-line y ∗ = L/2 and the relative velocity u∗y /u∗0 along the
∗
mid-line y = L/2 of the cavity, respectively. Results were based on simulations
performed using a 512 x 512 numerical domain. Adherence conditions are
veried for all methods.
Figure 8.6 shows plots of the stream function and the vorticity isolines.
These stream function isolines give a very clear picture of the complex ow
structure that develops in the cavity ow and the eect of the Reynolds number
on the patterns of the steady recirculating eddies in the cavity. The increas-
ingly complexity of the ow structure with the Reynolds number can be also
appreciated in the second row of Figure 8.6, which shows the vorticity isolines
for the lid driven cavity when Re varies from 100 to 5000.
A primary vortex and a pair of counterrotating eddies of much smaller
strength develop in the lower corners of the cavity are observed for all values
of Re presented. At Re = 1000 and 5000 a second pair of tertiary vortices are
seen in the lower corners of the cavity and exclusively for Re = 5000 a left top
vortex is observed. Additionally, as shown in Table 3, for low Re (Re=100), the
center of the primary vortex is located in the upper-right quadrant of the cavity
and moves down towards the geometric center of the cavity as the Reynolds
number increases.
The velocity proles shown in Figures 8.4 and 8.5, and stream function and
vorticity elds in Figure 8.6 were obtained using the method based on the mass
8.4. COMPARATIVE ANALYSIS 171
Figure 8.5: Numerical results for Re=400, 1000 and 5000 showing the relative
velocity u∗y /u∗0 along the mid-line x∗ = L/2. Results are compared with Ghia et
al. [30], Botella and Peyret [8] and Erturk, Corke and Gökçöl [25]. Continuous
lines are from Bazarin et al. [ ? ].
Figure 8.6: Stream function and vorticity isolines for (a) Re = 100, (b) Re =
400, (c) Re = 1000, (d) Re = 3200 and (e) Re = 5000. The top row illustrate
the stream function and the bottom row the vorticity, Bazarin et al. [ ? ].
172 CHAPTER 8. BOUNDARY CONDITIONS: PART II
¯
P ∗
x,y ρ
ρ= ,
L∗2
was measured for each time step for all the models that were analyzed in this
work. Results are shown in Figures 8.7 and 8.8.
As expected, both methods (I and II), based on Eq. (8.3), are clearly mass
conservative but Method IV also appears as mass conservative. For clarifying
this issue Bazarin et al. (2021) changed the average density scale in Figure 8.8,
measuring the total mass variation with a more precise weight balance. While
in both methods based on Eq. (8.3) the total mass is constant with time,
method IV based on density, presents the same linear behavior as observed in
Malaspinas et al. (2011) and Latt et al. (2008) models, although with a much
less mass variation. Curiously, Mohammed & Reis (2017) model preserves the
total mass although being a density based model.
Therefore, for investigating this issue more carefully, local mass leaking was
also veried by computing
at each time step and at each boundary node, using the information on the
pre-collisional populations f¯i that were propagated to the boundary sites xb
from the uid phase and the information on the populations ¯fˆ out that leave
i
these sites after collision and that were calculated based on the information we
have for the rst moments ρ∗ , τxx
∗ ∗
, τxy and
∗
τyy .
Figure 8.9 shows the local mass leakage for Mohammed and Reis (2017)
model after 50000 time steps. It is clear that Mohammed and Reis model is
8.4. COMPARATIVE ANALYSIS 173
Figure 8.7: Average density in terms of the number of time steps. We compare
the results of Latt et al. (2008), Malaspinas et al. (2011) and Mohammed
and Reis (2017) with the solution methods presented by Bazarin et al. (2021).
Both Malaspinas et al. and Latt et al. models fail in preserving the total mass
of the system. Method IV based on the concept of density is neither a mass
preserving method, but mass leaking is much smaller, Bazarin et al. [ ? ].
Figure 8.8: Average density measured with a more accurate weigth balance in
terms of the number of time steps. method IV based on density, presents the
same linear behavior as observed in Malaspinas et al. (2011) and Latt et al.
(2008) models, although with a much less mass variation, Bazarin et al. [ ? ].
174 CHAPTER 8. BOUNDARY CONDITIONS: PART II
Figure 8.9: Local mass leakage for Mohammed and Reis (2017) model after
50000 time steps: (a) north and south boundaries; (b) east and west boundaries;
8.4. COMPARATIVE ANALYSIS 175
not mass conservative from a local standpoint. In fact, although the model is
globally mass conservative there is a local mass leaking on the upper corners.
Nevertheless, the addition of mass on the left upper corner is compensated by
a loss of mass on the right upper corner.
176 CHAPTER 8. BOUNDARY CONDITIONS: PART II
Bibliography
[1] T. Abe, Derivation of the Lattice Boltzmann Method by Means of the
Discrete Ordinate Method for the Boltzmann Equation, Journal of Com-
putational Physics, 131 (1997), pp. 241246.
[7] , Vorlesungen uber Gas Theory, Part II, J.A. Barth, Leipzig, 1896.
[10] C.-H. Bruneau and M. Saad, The 2D lid-driven cavity problem revis-
ited, Computers and Fluids, 35 (2006), pp. 326348.
177
178 BIBLIOGRAPHY
[19] , Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich
daraus für die Wärmelehre selbst ableiten lassen, Annalen der Physik und
Chemie, 155 (1850), pp. 500524.
[20] , The mechanical theory of heat, with its applications to the steam-
engine and to the physical properties of bodies, John van Voorst, London,
1867.
[35] Z. Guo, S. Chang, and C. Shu, Lattice Boltzmann Method and its
Applications in Engineering, vol. 5, World Scientic Publishers, Singapore,
2013.
[40] X. He and L.-S. Luo, Theory of the lattice Boltzmann method: From the
Boltzmann equation to the lattice Boltzmann equation, Physical Review E,
56 (1997), pp. 68116817.
[56] P. Lallemand and L.-S. Luo, Theory of the lattice Boltzmann Method:
Dispersion, Dissipation, Isotropy, Galilean Invariance, and Stability,
Physical Review E, 61 (2000), pp. 65466562.
[71] S. Mohammed and T. Reis, Using the Lid-Driven Cavity Flow to Vali-
date Moment-Based Boundary Conditions for the Lattice Boltzmann Equa-
tion, Archive of Mechanical Engineering, 64 (2017), pp. 5774.
[86] S. Succi, The Lattice Boltzmann Equation for Fluid Dynamics and Be-
yond (Numerical Mathematics and Scientic Computation) 1st Edition,
Clarendon Press, 2001.
[88] J. D. van der Waals, Over de Continuiteit van den Gas en Vloeistoftoe-
stand, PhD thesis, Leiden University, 1873.
[91] R. Zhang, X. Shan, and H. Chen, Ecient kinetic method for uid
simulation beyond the Navier-Stokes equation, Physical Review E, 74
(2006), p. 046703.
184 BIBLIOGRAPHY
[93] Q. Zou and X. He, On pressure and velocity boundary conditions for the
lattice Boltzmann BGK model, Physics of Fluids, 9 (1997), pp. 15911598.
[7] , Vorlesungen uber Gas Theory, Part II, J.A. Barth, Leipzig, 1896.
[10] C.-H. Bruneau and M. Saad, The 2D lid-driven cavity problem revis-
ited, Computers and Fluids, 35 (2006), pp. 326348.
185
186 BIBLIOGRAPHY
[19] , Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich
daraus für die Wärmelehre selbst ableiten lassen, Annalen der Physik und
Chemie, 155 (1850), pp. 500524.
[20] , The mechanical theory of heat, with its applications to the steam-
engine and to the physical properties of bodies, John van Voorst, London,
1867.
[35] Z. Guo, S. Chang, and C. Shu, Lattice Boltzmann Method and its
Applications in Engineering, vol. 5, World Scientic Publishers, Singapore,
2013.
[40] X. He and L.-S. Luo, Theory of the lattice Boltzmann method: From the
Boltzmann equation to the lattice Boltzmann equation, Physical Review E,
56 (1997), pp. 68116817.
[56] P. Lallemand and L.-S. Luo, Theory of the lattice Boltzmann Method:
Dispersion, Dissipation, Isotropy, Galilean Invariance, and Stability,
Physical Review E, 61 (2000), pp. 65466562.
[71] S. Mohammed and T. Reis, Using the Lid-Driven Cavity Flow to Vali-
date Moment-Based Boundary Conditions for the Lattice Boltzmann Equa-
tion, Archive of Mechanical Engineering, 64 (2017), pp. 5774.
[86] S. Succi, The Lattice Boltzmann Equation for Fluid Dynamics and Be-
yond (Numerical Mathematics and Scientic Computation) 1st Edition,
Clarendon Press, 2001.
[88] J. D. van der Waals, Over de Continuiteit van den Gas en Vloeistoftoe-
stand, PhD thesis, Leiden University, 1873.
[91] R. Zhang, X. Shan, and H. Chen, Ecient kinetic method for uid
simulation beyond the Navier-Stokes equation, Physical Review E, 74
(2006), p. 046703.
192 BIBLIOGRAPHY
[93] Q. Zou and X. He, On pressure and velocity boundary conditions for the
lattice Boltzmann BGK model, Physics of Fluids, 9 (1997), pp. 15911598.