Cellular Automata Modeling of Physical Systems
Cellular Automata Modeling of Physical Systems
com
Collection Alea-Saclay:
Monographs and Texts in Statistical Physics
www.Ebook777.com
Cellular Automata
Modeling of Physical
Systems
. . .�! CAMBRIDGE
;:; UNIVERSITY PRESS
www.Ebook777.com
PUBLISHED BY THE PRESS SYNDICATE OF THE UNIVERSITY OF CAMBRIDGE
The Pitt Building, Trumpington Street, Cambridge CB2 1 RP, United Kingdom
A catalogue record for this book is available from the British Library
www.Ebook777.com
To Viviane
Clea, Tristan and Daphne
and Dominique
Anne-Virginie and Philippe
This book provides a self-contained introduction to cellular automata and lattice
Boltzmann techniques.
Beginning with a chapter introducing the basic concepts of this developing
field, a second chapter describes methods used in cellular automata modeling.
Following chapters discuss the statistical mechanics of lattice gases, diffusion
phenomena, reaction-diffusion processes and nonequilibrium phase transitions. A
final chapter looks at other models and applications, such as wave propagation
· and multiparticle fluids. With a pedagogic approach, the volume focuses on the
use of cellular automata in the framework of equilibrium and nonequilibrium
statistical physics. It also emphasizes application-oriented problems such as
fluid dynamics and pattern formation. The book contains many examples and
problems. A glossary and a detailed list of references are also included.
This will be a valuable book for graduate students and researchers working in
statistical physics, solid state physics, chemical physics and computer science.
BASTIEN CHOPARD obtained a Masters degree and a PhD in theoretical physics
from the University of Geneva. After three years as a research associate at MIT
(Cambridge, Mass.) and in the Forschungzentrum in Jtilich (Germany), he is now
assistant professor in the Department of Computer Sciences (Parallel Computing
Group) of the University of Geneva. His research interests concern the modeling
and simulation of complex systems and parallel computing.
MICHEL D Roz obtained a Masters degree and a PhD in theoretical physics
from the University of Geneva. He spent three years as research associate at
Cornell University (Ithaca, NY) and Temple University (Phihi.delphia, Penn.). He
is now at the Department of Theoretical Physics of the University of Geneva
where his main fields of research are equilibrium and non-equilibrium statistical
mechanics, complex systems and cellular-automata modeling of physical systems.
Everything should be made as simple as possible but not simpler
A. Einstein
Contents
Preface Xl
1 Introduction 1
1.1 Brief history 1
1 . 1 . 1 Self-reproducing systems 1
1 . 1 .2 Simple dynamical systems 3
1 . 1 .3 A synthetic universe 4
1 . 1 .4 Modeling physical systems 5
1 . 1 . 5 Beyond the cellular automata dynamics : lattice Boltzmann
methods and multiparticle models 7
1.2 A simple cellular automaton : the parity rule 8
1.3 Definitions 12
1 . 3 . 1 Cellular automata 12
1 .3.2 Neighborhood 14
1 .3.3 Boundary conditions 15
1 . 3.4 Some remarks 16
1 .4 Problems 18
Vll
V111 Contents
References 313
Glossary 327
Index 337
Preface
The cellular automata approach and the related modeling techniques are
powerful methods to describe, understand and simulate the behavior of
complex systems. The aim of this book is to provide a pedagogical
and self-contained introduction to this field and also to introduce recent
developments. Our main goal .is to present the fundamental theoretical
concepts necessary for a researcher to address advanced applications in
physics and other scientific areas.
In particular, this book discusses the use of cellular automata in the
framework of equilibrium and nonequilibrium statistical physics and in
application-oriented problems. The basic ideas and concepts are illustrated
on simple examples so as to highlight the method. A selected bibliog
raphy is provided in order to guide the reader through this expanding
field.
Several relevant domains of application have been mentioned only
through references to the bibliography, or are treated superficially. This
is not because we feel these topics are less important but, rather, because
a somewhat subjective selection was necessary according to the scope
of the book. Nevertheless, we think that the topics we have covered
are significant enough to give a fair idea of how the cellular automata
technique may be applied to other systems.
This book is written for researchers and students working in statistical
physics, solid state physics, chemical physics and computer science, and
anyone interested in modeling complex systems. A glossary is included
to give a definition of several technical terms that are frequently used
throughout the text. At the end of the first six chapters, a selection
of problems is given. These problems will help the reader to become
familiar with the concepts introduced in the corresponding chapter, or will
introduce him to new topics that have not been covered in the text. Some
problems are rather easy, although they usually require some programming
Xl
xii Preface
effort, but other problems are more involved and will demand significant
time to complete.
Most of the cellular automata simulations and results presented in this
book have been produced on the 8k Connection Machine CM-200 of
the University of Geneva. Others have been computed on an IBM SP2
parallel computer, also installed at the University of Geneva. Although
a parallel supercomputer is quite useful when considering large scale
simulations, common workstations and even modern personal computers
are well adapted to perform cellular automata computations, except for
on-line display which is alway very desirable. Dedicated hardware is also
available but, usually, less flexible than a general purpose machine.
Despite our effort, several errors and misprints are still likely to be
present. Please report them to us* (as well as any comment or suggestion).
We would like to thank all the people who have made this book
possible and, in particular Claude Godreche who gave us the opportunity
to write it. Special thanks go to Pascal Luthi and Alexandre Masselot
who made several original and important simulations which are presented
in this book. Other people have played a direct or indirect role in
the preparation of the manuscript. Among them, we thank Rodolphe
Chatagny, Stephen Cornell, Laurent Frachebourg, Alan McKane, Zoltan
Racz and Pierre-Antoine Rey.
Finally we acknowledge the Swiss National Science Foundation who
funded some of the research reported here and the Computer Science
Department and the Theoretical Physics Department of the University of
Geneva for making available the necessary environment and infrastructure
for this enterprise.
1
2 1 Introduction
be simulated by the rule of the automaton. All this shows that quite
complex and unexpected behavior can emerge from a cellular automaton
rule.
After the work of von Neumann, others have followed the same
line of research and the problem is still of interest [14] . In particular,
E.F. Codd [15] in 1 968 and much later C.G. Langton [16] and Byl [17]
proposed much simpler cellular automata rules capable of self-replicating
and using only eight states. This simplification was made possible by giv
ing up the property of computational universality, while still conserving
the idea of having a spatially distributed sequence of instructions (a kind
of cellular DNA) which is executed to create a new structure and then
entirely copied in this new structure.
More generally, artificial life is currently a domain which is intensively
studied. Its purpose is to better understand real life and the behavior of
living species through computer models. Cellular automata have been an
early attempt in this direction and can certainly be further exploited to
progress in this field [18, 19] .
. - •• "(t
· ·
0(10 ·-·- o0o · -·- . - •
0
I I •
0 l""'b-
• • I
..... '"I:.)
...
.) ..........
"'=
. . -·-
..... �
t+lO t+20
Fig. 1 . 1 . The game of life automaton. Black dots represents living cells whereas
dead cells are white. The figure shows the evolution of some random initial
configurations.
the visualization of the fast moving particle shed a different light on the
possibilities of such models : isn't it possible to simulate the behavior of a
real system of particles (like a fluid or a gas) as a cellular automata rule ?
After all, it is well known that the flows of a fluid, a gas or even a granular
medium are very similar at a macroscopic scale, in spite of their different
microscopic nature. A fully discrete and simplified molecular dynamics
could work too, provided the system is considered at an appropriate
observation scale.
Of course, the idea of using discrete systems as a model of real phenom
ena has already been considered for several problems. The Ising model
of classical spin is a famous example which will be discussed in more
detail in the next chapter. From the fluid side, already at the end of the
nineteenth century, Maxwell [29] , had proposed a discrete velocity system
of interacting particles as a model of a gas. In fact, such lattice gas,
discrete velocity models have been developed independently from cellular
automata theory [30,3 1] .
However, cellular automata provide a new conceptual framework, as
well as an effective numerical tool, which retains important aspects of the
microscopic laws of physics, such as simultaneity of the motion, locality
of the interactions and time reversibility.
Cellular automata rules are viewed as an alternative form of the
microscopic reality which bears the expected macroscopic behavior. From
a numerical point of view it was expected, at the end of the 1 980s, that
a wind tunnel could be replaced by a fully discrete computer model.
The first cellular automata model to give credit to this possibility is the
famous FHP model proposed in 1986 by U. Frisch, B. Hasslacher and
Y. Pomeau [32] , and almost simultaneously by S. Wolfram [33] . These
authors showed that their model, despite its fully discrete dynamics, fol
lows, in some appropriate limits, the behavior prescribed by the Navier
Stokes equation of hydrodynamics.
Note that models like FHP or HPP are often termed lattice gas auto
mata (LOA) to distinguish them from the less specific cellular automata
terminology. Clearly, from a mathematical point of view, a lattice gas
automata is a cellular automata, but the way one thinks, for instance,
of the game of life is quite different from the underlying philosophy
of the FHP model. This difference will become clear to the reader
as he or she becomes more familiar with the next chapter of this book.
Nevertheless, in this book, we will often use cellular automata to designate
a LGA.
Since the FHP rule was discovered, lattice gas automata or cellular
automata fluids as these kind of particle models are now often referred to,
have been developed intensively and several insufficiencies of the initial
model corrected. The Ecole Normale Superieure in Paris has been very
1 . 1 Brief history 7
( a) (b) (c )
Fig. 1 .2. The E!3 rule on a 256 x 256 periodic lattice : (a) initial configuration ; (b)
and (c) configurations after tb = 93 and tc = 1 10 iterations, respectively.
Each site of the lattice is a cell which is labeled by its position r = (i, j )
where i and j are the row and column indices. A function tpt (r ) is
associated to the lattice to describe the state of each cell at iteration t.
This quantity can be either 0 or 1 .
The cellular automata rule specifies how the states tpt+ l are to be
computed from the states at iteration t. We start from an initial condition
at time t = 0 with a given configuration of the values tpo (r ) on the lattice.
The state at time t = 1 will be obtained as follows
( 1 ) Each site r computes the sum of the values tpo (r ' ) on the four
nearest neighbor sites r ' at north, west, south and east. The system
is supposed to be periodic in both i and j directions (as on a torus)
so that this calculation is well defined for all sites.
(2) If this sum is even, the new state tp1 (r) is 0 (white), else, it is 1 (black).
The same rule (step 1 and 2) is repeated to find the states at time
t = 2, 3, 4, ... .
From a mathematical point of view, this cellular automata parity rule
can be expressed by the following relation
lpt+ l (i, j ) = tpt (i + 1, j ) Etl tpt (i - 1, j) Etl tpt (i, j + 1 ) Etl tpt (i, j - 1 ) ( 1. 1 )
where the symbol Etl stands for the exclusive O R logical operation. I t is
also the sum modulo 2 : 1 Etl 1 = 0 EB 0 = 0 and 1 Etl 0 = 0 EB 1 = 1 .
When this rule is iterated, very nice geometrical patterns are observed,
as shown in figure 1 . 2 . This property of generating complex patterns
starting from a simple rule is actually generic of many cellular automata
rules. Here, complexity results from some spatial organization which
builds up as the rule is iterated. The various contributions of successive
iterations combine together in a specific way. The spatial patterns that
are observed reflect how the terms are combined algebraically.
10 1 Introduction
Cellular
Automata
( a) (b) (c)
Fig. 1.3. The EB rule replicates any initial pattern when the number of iterations
is a power of two. Image (a) shows the initial pattern at time ta = 0. Images (b)
and (c) show successive iterations at times tb = 16 and tc = tb + 32.
1.3 Definitions
(ii) a set <l> (r, t) = { <1>1 (r, t ), <1>2(r, t), ... , <I>m (r, t ) } of Boolean variables
attached to each site r of the lattice and giving the local state
of each cell at the time t = 0, 1, 2, ... ;
(iii) a rule R = {Rt, R2, ... , Rm} which specifies the time evolution of the
states <l> (r, t ) in the following way
<I>j(r, t + 1 ) = Rj(<I>(r , t), <I>(r + J 1 . t), <I>(r + J2, t ), ... , <I>(r + bq, t )) ( 1 .8)
where r + Jk designate the cells belonging to a given neighborhood
of cell r.
The example discussed in the previous section is a particular case in
which the state of each cell consists of a single bit <1> 1 (r, t) = lpt (r) of
information and the rule is the addition modulo 2.
In the above definition, the rule R is identical for all sites and is applied
simultaneously to each of them, leading to a synchronous dynamics. It is
important to notice that the rule is homogeneous, that is it cannot depend
explicitly on the cell position r. However, spatial (or even temporal)
inhomogeneities can be introduced by having some <I>j(r ) systematically 1
in some given locations of the lattice to mark particular cells for which
a different rule applies. Boundary cells are a typical example of spatial
inhomogeneity. Similarly, it is easy to alternate between two rules by
having a bit which is 1 at even time steps and 0 at odd time steps.
In our definition, the new state at time t + 1 is only a function of the
previous state at time t. It is sometimes necessary to have a longer memory
and introduce a dependence on the states at time t - 1 , t - 2, ... ,t-k. Such
a situation is already included in the definition if one keeps a copy of the
previous states in the current state. Extra bits <l>t can be defined for this
purpose. As an example, the one-dimensional second-order rule
<l>t( r, t + 1 ) = R(<l>t(r - 1, t), <l>t( r + 1, t)) ® <l>t( r, t-1 ) ( 1.9)
where EEl is the sum modulo 2, can be expressed as a first-order rule by
introducing a new state <l>2(r, t ) as follows
<l>t( r, t + 1 ) = R(<l>t(r-1, t ), <l>t( r + 1, t)) EEl <l>2(r, t) ( 1 . 10)
<1>2( r, t + 1 ) = <l>t( r, t )
Note that in the first case, the initial condition requires one to specify
<l>t( r, t = 0 ) and <l>t( r, t = 1 ) . In the second case, there are still two initial
conditions to specify, but now they are <l>t(r, t = 0 ) and <l>2(r, t = 0 ) .
An expression such as rule 1 .9 provides a general way to build a
time-reversible dynamics. Indeed, since the operation EEl is commutative,
14 1 Introduction
1 .3.2 Neighborhood
A cellular automata rule is local, by definition. The updating of a given
cell requires one to know only the state of the cells in its vicinity. The
spatial region in which a cell needs to search is called the neighborhood.
In principle, there is no restriction on the size of the neighborhood, except
that it is the same for all cells. However, in practice, it is often made up
of adjacent cells only. If the neighborhood is too large, the complexity
of the rule may be unacceptable (complexity usually grows exponentially
fast with the number of cells in the neighborhood).
For two-dimensional cellular automata, two neighborhoods are often
considered : the von Neumann neighborhood, which consists of a central
cell (the one which is to be updated) and its four geographical neigh
bors north, west, south and east. The Moore neighborhood contains, in
addition, second nearest neighbors north-east, north-west, south-east and
south-west, that is a total of nine cells. Figure 1.4 illustrates these two
standard neighborhoods.
Another useful neighborhood is the so-called Margolus neighborhood
which allows a partitioning of space and a reduction of rule complexity.
The space is divided into adjacent blocks of two-by-two cells. The rule
is sensitive to the location within this so-called Margolus block, namely
upper-left, upper-right, lower-left and lower-right. The way the lattice is
partitioned changes as the rule is iterated. It alternates between an odd
and an even partition, as shown in figure 1.5. As a result, information can
propagate outside the boundaries of the blocks, as evolution takes place.
The key idea of the Margolus neighborhood is that when updating
occurs, the cells within a block evolve only according to the state of that
block and do not immediately depend on what is in the adjacent blocks.
To understand this property, consider the difference with the situation
of the Moore neighborhood shown in figure 1.4. After an iteration, the
cell located on the west of the shaded cell will evolve relatively to its
own Moore neighborhood. Therefore, the cells within a given Moore
neighborhood evolve according to the cells located in a wider region. The
purpose of space partitioning is to prevent these "long distance" effects
from occurring. The sand rule described in section 2.2.6 give an example
of the use of the Margolus neighborhood.
1 .3 Definitions 15
(a) (b)
Fig. 1 .4. (a) Von Neumann and (b) Moore neighborhoods. The shaded region
indicates the central cell which is updated according to the state of the cells
located within the domain marked with the bold line.
ul ur
11 lr
periodic fixed
adiabatic reflection
1.4 Problems
var i a b l e fo r each auto m ato n state and each l attice s ite . Then , the
ru l e is co m puted o n the fly, accord i n g to some arith metic o r p rog ram
i nstructions, ove r and over fo r each iteration of the ru n . ( i i ) Lookup
table: anoth e r tech n iq u e is to p re-co m p ute the o utputs of the ru l e fo r a l l
poss i b l e confi g u rati o ns of the n e i g h borhood a n d sto re them i n a lookup
tab l e . T h i s sol ution is fast s i nce the co m p utation red uces to cal c u l at i n g
an i nd ex i n t h e l o o k u p tab l e fo r each l attice s i te . But it uses q u ite a
l ot of m e m o ry if the n u m be r of b i ts n ecessary to specify the state of
each ce l l is l a rge (the tab l e g rows expo nenti a l l y fast) . ( i i i ) M u ltispin
coding: f i n a l l y a th i rd method is m u ltisp i n cod i ng . To save m e m o ry and
speed , o n e packs seve ral l attice s i tes in o n e co m p ute r wo rd (actua l ly,
32 s i tes in a 32-bit word ) . Then, b itwise l og ical operations ( l i ke AND ,
OR , . . . ) , can be used to i m p l ement the ru l e and co m p ute at once the
n ew state of 32 s i tes. Thus, even if one repeats the co m p utation fo r
each i terat i o n , there i s a co n s i d e ra b l e i m p rovement of perform ance
compared with m ethod ( i ) . In genera l , the b i ts a re packed in the words
in the fo l l owi n g way (fi nd out why this is advantageous) : the fi rst s ite is
stored in the fi rst b i t of the fi rst word ; the seco nd s i te goes to the first
b i t of the second wo rd and so o n . With n wo rds , s i tes n + 1 , n + 2 , . . . ,n + n
are stored i n the secon d b i t of a l l the words. T h u s , b i t b E {0, 1 , ... , 3 1 }
o f word k E { 1, . . , n} conta i ns o n e b i t of the state o f s ite l = b * n + k. Fo r
a ru l e req u i ri n g s b i ts per state, ooe n eeds s sets of n words. I n a 20
l attice, each l i ne is m u ltisp i n coded .
1 . 1 0. Snowfl a kes [56] : a s i m p l e CA ru l e can be d evised to model the
g rowth of an obj ect whose structu re s hows some s i m i larities with a
snowfl ake. C rysta l l ization can be vi ewed as a so l i d ificati o n p rocess i n
a sol i d-l i q u i d system . Altho u g h crysta l l i zati o n g rows from a n existi n g
seed , o n e o f the key featu re o f t h i s process is the growth inhibition
phenomenon whereby a crystal l i zed s i te p reve nts nearby s i tes from
so l i d ify i n g . T h i s effect can be i ncorpo rated in the s i m p l e fo l l owi n g
r u l e . Each s i te i s i n o n e o f two states: so l id o r l i q u i d ; a no n-so l i d
s i te that has exactly one so l i d n e i g hbor crystal l i zes . The oth e r s i tes
rem a i n u nchanged . S i m u l ate t h i s ru l e on a h exagonal l attice, sta rti n g
w i t h one i n itial sol i d seed i n the m id d l e . Mapp i n g a hexagonal g r i d
o n a squ are l attice c a n be performed b y co ns i d e r i n g two tru ncated
M oore n e i g h borhoods (see also section 7.2.4) : i n add ition to n o rth ,
south , east and west, odd l i n e s i tes of the l attice o n l y see no rth-west
and south-west n e i g h bors, whereas even l i ne s i tes see no rth-east and
south-east.
2
Cellular automata modeling
21
22 2 Cellular automata modeling
The simplest cellular automata rules are one-dimensional ones for which
each site has only two possible states and the rule involves only the nearest
neighbors sites. They are easily programmable on a personal computer
and offer a nice "toy model" to start the study of cellular automata.
A systematic study of these rules was undertaken by S. Wolfram in
1983 [24,25] . Each cell (labeled i) has at a given time, two possible
states S i = 0 or 1. The state si at time t + 1 depends only on the triplet
(si-l , si, si+ d at time t :
(2. 1 )
Thus t o each triplet o f sites one associates a value lik = 0 o r 1 according
to the following list :
( a) (b ) (c) ( d)
Fig. 2. 1 . Example of the four Wolfram rules with a random initial configuration.
Horizontal lines correspond to consecutive iterations. The initial state is the
uppermost line. (a) Rule 40 belonging to class 1 reaches very quickly a fixed
point (stable configuration). (b) Rule 56 of class 2 reaches a pattern composed
of stripes which move from left to right. (c) Rule 18 is in class 3 and exhibits a
self-similar pattern. (d) Rule 1 10 is an example of a class 4 cellular automaton.
Its behavior is not predictable and as a consequence, we observe a rupture in the
pattern, on the left part.
o ���--��-���
0 0.6
c(t)
Fig. 2.2. Non-trivial collective behavior for the simple 3D cellular automata rule
given by relation (2.4). The plot shows the behavior of the global density c(t) as
a function of its previous value.
s 1·(t + 1 ) = {0
1 if 2:-ne ig_hbors Sj( t)
otherw1se
E {0, 5}
(2.4)
Here the sum over the neighbors extends to the six nearest neighbors in
the cubic lattice, plus the central site i itself. Thus, the sum ranges between
0 and 7. The rule is that the new value si ( t + 1 ) is 1 if and only if this sum
equals 0 or 5.
As time goes on, one can record the density c ( t) defined as
c ( t) = L si ( t )
lattice
This quantity exhibits a non-trivial behavior in the sense that it does
not relax to a constant value at large time. Instead, it follows a low
dimensional dynamics which can be represented by the map shown in
figure 2.2 where c ( t + 1) is plotted as a function of c ( t ) .
Such non-trivial collective behaviors in spatially extended systems has
been observed on several totalistic cellular automata rules. A totalistic rule
means that the new state of the cell i is a function only of the sum of the
states Sj for cells j belonging to a given neighborhood of i. Rule (2.4) is an
example of a totalistic rule. Totalistic rules can be seen as mean-field-like
evolution since in the each site "sees" an average environment.
Non-trivial collective behaviors are not singular phenomena, requiring
the fine tuning of some conditions, but appear as general features, robust
26 2 Cellular automata modeling
examples, we shall develop our intuition and also identify some problems
related to the cellular automata approach. In particular, we shall see the
importance of developing techniques to assess the range of validity of a
model and its limitations.
t+l
t
X
Fig. 2.3. Deterministic growth from a rough surface parallel to the (1, 1) direction.
At each time step, all local surface minima are filled with new particles. The
corresponding lattice gas evolves according to the automata rule 1 84.
figure 2.4. Any lattice site may be present with a probability p or absent
with probability ( 1 - p). Moreover, any bond may be present with prob
ability q and absent with probability ( 1 - q). Assume that a single site is
"wet" at time t = 0. Present bonds conduct "water" only in the downward
(increasing time) direction, and when both sites connected by the bond
are present.
For q = 1, one speaks of site-directed percola tion, and for p = 1,
of bond-directed percola tion. Given the two probabilities p and q, one
can ask several questions : what is the probability of finding wet sites at
(time) level t and what is the typical size of the wet cluster ? In each
case, there are critical values (pc, qc), called the percolation thresholds,
above which the wet cluster spans the whole system. Directed perco
lation is a well-known problem of "geometrical phase transition" for
which scaling theory and the renormalization group approach have been
developed.
This problem can easily be formulated in term of probabilistic cellular
automaton. Consider a linear chain or ring of sites ; with each site one
associates a variable 1JJ i ( t) = 0 or 1 . At odd (even) times, odd (even)
indexed sites change their state according to a probabilistic rule, and even
(odd) indexed sites stay in the same state. The full space-time history
can be presented on a two-dimensional lattice. The rules are defined
by the conditional probabilities P ( 1pi(t + 1 ) 1 1JJ i- l ( t ), 1JJ i+ l (t)) . For directed
2.2 Modeling of simple systems : a sampler of rules 31
0
1
2
3
4
5
6
Fig. 2.4. Two-dimensional square lattice for directed percolation. The dots
represent the sites which can be present or absent. The heavy lines between the
sites represent the bonds. The vertical axis corresponds to time.
percolation, we have :
P ( 1 1 00 ) = 0, y = P ( 1 1 1 1 ) = pq , z
P ( l i01 ) = P ( 1 1 10) = p q(2 - q)
=
(2 .6 )
The situation in the parameter space (p , q) is represented in figure 2 .5.
At the transition point, the percolation cluster is a fractal object. More
over, near the transition point, many physical quantities have a power law
behavior in terms of p - Pc or q - qc, with universal critical exponents.
Here again, one sees that a very simple cellular automaton rule leads to a
complex structure.
Another interesting and simple example of a probabilistic cellular au
tomata rule is related to forest fire models [6 1,62] . This probabilistic
cellular automata model is defined on a d-dimensional hypercubic lattice.
Initially, each site is occupied by either a tree, a burning tree or is empty.
The state of the system is parallel updated according to the following rule :
1 . A burning tree becomes an empty site.
2. A green tree becomes a burning tree if at least one of its nearest
neighbors is burning.
1 .0 -�--:c�-;;;-
>..-.,
...
; ..
,----
-- ----�---
,.
,.
,.
,.
.·
z
,.
,.
,.
,.
,.
;
;
.·
;
;
.·
;
.·
0.5 ;
;
.·
;
.·
;
,
.·
,
.·
.·
,
I
•
•
•
.
0 0.5 y 1 .0
Fig. 2.5. Parameter space y = pq ; z = pq(2 - q ) defining the cellular automata
rules for generalized directed percolation. The dashed and dotted lines correspond,
respectively, to the bond and site percolation. The system percolates to the right
of the (solid) transition line.
(2.7)
and
(2.8)
Fig. 2.6. The forest fire cellular automata rule : grey sites correspond to a
grown tree, black pixels represent burned sites and the white color indicates a
burning tree. The snapshot given here represents the situation after a few hundred
iterations. The parameters of the rule are p = 0.3 and f = 6 X w-5•
flip if the number of its neighbors with spin up is the same as the number
of its neighbors with spin down. However, one has to remember that the
motion of all spins is simultaneous in a cellular automata. The decision
to flip is based on the assumption that the neighbors are not changing. If
they flip too (because they obey the same rule), then energy may not be
conserved. This is simply illustrated by a one-dimensional chain of spins.
Suppose one has the following spin configuration at time t
0 0 0 ( 2. 1 1 )
1 1 1 ...
... 1 0 1 0 1 0 ...
.. 0. 1 0 1 0 1 ...
The flipping rule described earlier is then regulated by the value of b. It
takes place only for those sites for which b = 1 . Of course, the value of b
is also updated at each iteration according to b ( t + 1 ) = 1 b ( t ), so that at -
si1' ( t + 1 ) =
{ S1ij(t)- Sij(t) if bij = 1 and S i-l,j + S i + l,j + Si,j- 1 + S i,j+ l
otherwise
= 2
(2. 1 3 )
and
(2. 14)
where the indices (i, j) label the Cartesian coordinates and sij( t = 0 ) is
either one or zero.
The question is now how well does this cellular automata rule perform
to describe an Ising model. Figure 2.7 shows a computer simulation of
the Q2R rule, starting from an initial configuration with approximately
1 1 % of spins S ij = 1 (figure 2.7 (a)). After a transient phase (figures
(b) and (c)), the system reaches a stationary state where domains with
"up" magnetization (white regions) are surrounded by domains of "down"
magnetization (black regions).
In this dynamics, energy is exactly conserved because that is the way
the rule is built. However, the number of spins down and up may vary.
Actually, in the present experiment, the fraction of spins up increases from
1 1 % in the initial state to about 40% in the stationary state. Since there
is an excess of spins down in this system, there is a resulting macroscopic
magnetization.
It is interesting to study this model with various initial fractions P s of
spins up. When starting with a random initial condition, similar to that of
figure 2.7 (a), it is observed that, for many values of p5, the system evolves
to a state where there is, on average, the same amount of spin down
and up, that is no macroscopic magnetization. However, if the initial
configuration presents a sufficiently large excess of one kind of spin, then
a macroscopic magnetization builds up as time goes on. This means there
is a phase transition between a situation of zero magnetization and a
situation of positive or negative magnetization.
It turns out that this transition occurs when the total energy E of the
system is low enough (low energy means that most of the spins are aligned
36 2 Cellular automata modeling
(a) (b)
(c) (d)
Fig. 2.7. Evolution of a system of spins with the Q2R rule. Black represents the
spins down Sij = 0 and white the spins up Sij = 1. The four images (a), (b), (c)
and {d) show the system at four different times ta = 0 < tb < tc < < td .
and that there is an excess of one species over the other), or more precisely
when E is smaller than a critical energy Ec.
Therefore, the Q2R rule captures an important aspect of a real magnetic
system, namely a non-zero magnetization at low energy (which can be
related to a low temperature situation) and a transition to a non-magnetic
phase at high energy. However, Q2R also exhibits an unexpected behavior
that is difficult to detect from a simple observation. There is a loss of
ergodicity : a given initial configuration of energy Eo evolves without
visiting completely the region of the phase space characterized by E = E0 .
This is illustrated by the following simple example, where a ring of four
spins is considered. For the sake of convenience, we shall represent it as
2.2 Modeling of simple systems : a sampler of rules 37
a linear chain with periodic boundary conditions (the first and last spins
are nearest neighbors).
t : 1001
t+1 : 1 100
t+2 : 0 1 10
t+3 : 001 1
t+4 : 100 1 (2. 15)
After four iterations, the system cycles back to its original state. The
configuration of this example has Eo = 0. As we observe, it never evolves
to 01 1 1, which is also a configuration of zero energy.
This non-ergodicity means that not only energy is conserved during
evolution of the automaton, but also other quantities which partition the
energy surface in independent regions. As we shall see, the study of the
dynamical invariants is very important when studying cellular automata
models.
( a) (b) (c)
Fig. 2.8. Evolution of the annealing rule. The inherent "surface tension" present
in the rule tends to separate the black phases s = 1 from the white phase s = 0.
The snapshots (a), (b) and (c) correspond to t = 0, t = 72 and t = 270 iterations,
respectively. The extra gray levels indicate how "capes" have been eroded and
"bays" filled : dark gray shows the black regions that have been eroded during
the last few iterations and light gray marks the white regions that have been
filled.
I
T-
- ·
,_ - I
I
( a)
(b)
�
(c)
�
Fig. 2.10. The HPP rule : (a) a single particle has a ballistic motion until
it experiences a collision. (b) and (c) The two non-trivial collisions of the
HPP model : two particles experiencing a head-on collision are deflected in the
perpendicular direction. In the other situations, the motion is ballistic, that is the
particles are transparent to each other when they cross the same site.
The aim of this rule is to reproduce some aspects of the real interaction
of particles, namely that momentum and particle number are conserved
during a collision. From figure 2. 10, it is easy checked that these properties
are obeyed : for the two situations (b) and (c) where the particles do
actually experience a collision, a pair of zero momentum particles along a
given direction is transformed into another pair of zero momentum along
the perpendicular axis.
The HPP rule captures another important ingredient of the microscopic
nature of a real interaction : invariance under time reversal. Figures 2. 10
(b) and (c) show that, if at some given time, the direction of motion of all
particles are reversed, the system will just trace back its own history. Since
the dynamics of a deterministic cellular automaton is exact, this allows
2.2 Modeling of simple systems : a sampler of rules 41
(a)
(b)
Fig. 2. 1 1 . Time evolution o f a HPP gas. (a) From the initial state t o equilibrium.
(b) Illustration of time reversal invariance : in the rightmost image of (a), the
velocity of each particle is reversed and the particles naturally return to their
initial position.
Fig. 2. 12. If a small error is introduced (e.g. one particle is added) in the situation
of the previous figure before the velocities are reversed, the system cannot return
to its initial state. The particles left in the right part of the system are those which
have been affected by the presence of the extra particle.
Fig. 2. 1 5. A conflict of motion results from the use of the Moore neighborhood.
The two gray particles are unstable and both want to move to the same new
location.
can stack on top of each other if this arrangement is stable. Of course, real
sand grains do not stand on a regular lattice and the stability criteria are
expected to depend on the shape of each grain. Despite this microscopic
complexity, the result is that sand piles that are too high topple.
In the framework of cellular automata modeling, the toppling mecha
nisms can be captured by the following rule : a grain is stable if there is
a grain underneath and two other grains preventing it falling to the left
or right, as depicted in figure 2. 14. In the Moore neighborhood, it means
that a central grain will be at rest if the south-west, south and south-east
neighbors are occupied. Otherwise the gain topples downwards to the
nearest empty cell.
For instance, suppose one has the situation shown in figure 2. 1 5 . Since
the updating of all cells is simultaneous, the two shaded grains will move
to the same empty cell : there is a conflict. To prevent conflict, each
particle should know about the presence of the other one before they
move, and take a common decision as to which one is going to fall. This
requires a larger neighborhood than the Moore neighborhood and a more
complicated rule.
The Margolus neighborhood gives a simple way to deal with the syn
chronous motion of all particles. Piling and toppling is determined by the
situation within 2 x 2 adjacent blocks. Of course (see section 1.3.2), for all
other steps, the space partitioning is shifted one cell down and one cell
left, so that, after two iterations, each cell sees a larger neighborhood.
Figure 2. 16 gives a possible implementation of a sand rule. It shows
how the configurations evolve due to toppling.
44 2 Cellular automata modeling
ffi EE � � � � [Sl�H�B
Fig. 2. 1 6. The sand rule expressed for the Margolus neighborhood. Only the
configuration that are modified are shown.
ffi
p � 1-p
ffi �
Fig. 2. 1 7. The sand rule allows for blocking (friction) with probability p when
two grains occupy the upper part of a Margolus block.
Fig. 2.19. Langton's ant rule. The motion of a single ant starts with a chaotic
phase of about 10 000 time steps, followed by the formation of a highway. The
figure shows the state of each lattice cell (gray or white) and the ant position
(marked by the black dot). In the initial condition all cells are white and the ant
is located in the middle of the image.
The proof is as follows : suppose the region the ant visits is bounded.
Then, it contains a finite number of cells. Since the number of iterations
is infinite, there is a domain of cells that are visited infinitely often.
Moreover, due to the rule of motion, a cell is either entered horizontally
(we call it an H cell) or vertically (we call it a V cell). Since the ant turns
by 90 degrees after each step, an H cell is surrounded by four V cells, and
conversely. As a consequence, the H and V cells tile the lattice in a fixed
checkerboard pattern. Now, we consider the upper rightmost cell of the
domain, that is a cell whose right and upper neighbor are not visited. This
cell exists if the trajectory is bounded. If this cell is an H cell (and is so for
ever), it has to be entered horizontally from the left and exited vertically
downward and, consequently must be gray. However, after the ant has
left, the cell is white and there is a contradiction. The same contradiction
appears if the cell is a V cell. Therefore, the ant trajectory is not bounded.
Following the same idea as in the HPP rule, we will introduce n i ('r, t)
as a Boolean variable representing the presence (n i = 1 ) or the absence
(n i = 0) of an ant entering site r at time t along lattice direction Ci, where
c 1 , c2 , c3 and c4 stand for direction right, up, left and down, respectively.
Up to four ants may enter the same site at the same time, from the four
sides. If the color Jl (r , t) of the site is gray (/1 = 0), all entering ants turn
90 degrees to the right. On the other hand, if the site is white (/1 = 1 ),
they all turn 90 degrees to the left.
This motion can be represented by the following expression
(2.27)
where the index i is wrapped around the values 1 to 4 (i.e i + 1 = 1 when
i = 4, and i - 1 = 4 when i = 1 ). The color of each cell is modified
after one or more ants have gone through. Here, we choose to make this
modification depend on the number of ants present
Jl(r, t + 1 ) = Jl(r, t ) EB n r ( r, t ) EB n2 ( r, t) EB n 3( r, t) EB n4 ( r, t) (2.28)
where EB is the sum modulo 2.
When the initial conditions are such that the cell contains one single ant,
rules 2.27 and 2.28 are equivalent to the original Langton's ant motion.
When several ants travel simultaneously on the lattice, both cooperative
and destructive behavior are observed. First, the erratic motion of several
ants favors the formation of a local arrangement of colors allowing the
creation of a highway. One has to wait much less time before the first
highway. Second, once a highway is created, other ants may use it to
travel very fast (they do not have to follow the complicated pattern of the
highway builder). In this way, the term "highway" is very appropriate.
Third, a destructive effect occurs as the second ant reaches the highway
builder. It breaks the pattern and several situations may be observed.
For instance, both ants may enter a new chaotic motion, or the highway
may be traveled in the other direction (note that the rule is time reversal
invariant) and is thus destroyed. Figure 2.20 illustrates the multi-ant
behavior.
The problem of an unbounded trajectory occurs again with this gener
alized motion. The assumption of Bunimovitch-Troubetzkoy's proof no
longer holds in this case because a cell may be both an H or a V cell.
Indeed, two different ants may enter the same cell, one vertically and the
other horizontally.
Actually, the theorem of unbounded motion is wrong in several cases
when two ants are present. Periodic motions may occur when the initial
positions are well chosen. Again, it comes as a surprise that, for many
initial conditions, the two ants follows periodic trajectories.
For instance, when the relative location of the second ant with respect to
2.2 Modeling of simple systems : a sampler of rules 49
Fig. 2.20. Motion of several Langton's ants. Gray and white indicate the colors
of the cell at the current time. Ant locations are marked by the black dots.
Initially, all cells are white and a few ants are randomly distributed in the central
region, with random directions of motion. The first highway appears much earlier
than when the ant is alone. In addition the highway can be used by other ants
to travel much faster. However, the "highway builder" is usually prevented from
continuing its construction as soon as it is reached by following ants. For instance,
the highway heading north-west after 4900 steps is destroyed. A new highway
emerges later, as we see from the snapshot at time t = 8564.
the first one is (�x, �y ) = (2, 3), the two ants returns to their initial position
after 478 iterations of the rule (provided they started in a uniformly black
substrate, with the same direction of motion). A very complicated periodic
behavior is observed when (�x. �y) = ( 1 , 24) : the two ants start a chaotic
like motion for several thousands of steps. Then, one ant builds a highway
and escapes from the central region. After a while, the second ant finds
the entrance of the highway and rapidly catches the first one. After the
two ants meet, they start undoing their previous path and return to their
original position. This complete cycle takes about 30 000 iterations.
More generally, it is found empirically that, when �x + �y is odd and
the ants enter the site with the same initial direction, the two-ant motion
is likely to be periodic. However, this is not a rule and the configuration
(�x, �y) = ( 1 , 0) yields an unbounded motion, a diamond pattern of
increasing diameter which is traveled in the same direction by the two
ants.
It turns out that the periodic behavior of a two-ant configuration is
not so surprising. The cellular automata rule 2.27 and 2.28 is reversible
50 2 Cellular automata modeling
in time, provided that there is never more than one ant at the same site.
Time reversal symmetry means that if the direction of motion of all ants
are reversed, they will move backwards through their own sequence of
steps, with the opposite direction of motion. Therefore, if at some point
of their motion the two ants cross each other (on a lattice link, not on
the same site), the first ant will go through the past of the second one,
and vice versa. They will return to the initial situation (the two ants being
exchanged) and build a new pattern, symmetrical to the first one, due to
the inversion of the directions of motion. The complete process then cycles
for ever. Periodic trajectories are therefore related to the probability that
the two ants will, at a some time, cross each other in a suitable way. The
conditions for this to happen are fulfilled when the ants sit on a different
sublattice (black or white sites on the checkerboard) and exit two adjacent
sites against each other. This explain why a periodic motion is likely to
occur when �x + �y is odd.
(2.29)
If the dynamics is time reversal invariant, one can change the direction of
motion of all ants and the system traces its own past, with the opposite
directions of motion. Let us introduce the operator R as the operator
which reverses the directions of motion but does not change the color
( Rs ) o = so
(Rs)i = S i-2 (2.30)
Time reversal invariance implies that U Rsou t = Rsin since if one reverses
the directions of motion and updates the rule, one expects to obtain the
initial state with reverse velocities. Since so ut = Usin , the condition for
time reversal invariance reads
URU = R or RU = u- 1 R (2.3 1 )
2.2 Modeling of simple systems : a sampler of rules 51
On the other hand, some two-ant configurations do not satisfy the condi
tion, because of the color rule 2.28
What about real ants ? So far we have not considered the behavior of
real ants and how they may be connected to this cellular automata. As
mentioned before, artificial life is a fascinating domain to which cellular
automata models may contribute. The initial purpose of Langton's ant
model was apparently to study large-scale regularities in complex systems.
However, this rule has some analogies with what happens in the real ant
world.
It is well known that real ants leave behind them a substance (the
pheromone) which is used by other ants as a stimulus. The idea of modi
fying the space texture (white or gray) captures an aspect this phenomena.
We have seen that the path created by an ant involve a spatial structure
that can be used by a companion to travel much more efficiently. It is
very tempting to relate this behavior with a famous experiment in which
real ants are able to find the shortest path to their food, as the result of a
collective effect.
Fig. 2.21 . Example of a traffic configuration near a junction. The four central
cells represent a rotary which is traveled counterclockwise. The gray levels indicate
the different traffic lanes : white is a northbound lane, light gray an eastbound
lane, gray a westbound lane and, finally, dark gray is a southbound lane. The
dots labeled a,b,c,d,e,f,g and h are cars which will move to the destination cell
indicated by the arrows. Cars without an arrow are forbidden to move.
road junction. This can be realized in a simple way if one assumes that a
rotary is located at each crossing. Road junctions are formed by central
circles around which the traffic moves always in the same direction. A
vehicle in a rotary has priority over any entering car.
The implementation we propose for this rule is illustrated in figure 2.21,
where a four-corner junction is shown. The four middle cells constitute the
rotary. A vehicle on the rotary (b or d) can either rotate counterclockwise
or exit. A local flag f is used to decide the motion of a car in a rotary.
If f = 0, the vehicle (d) exits in the direction allowed by the color of its
lane (see figure caption). If f = 1, the vehicle moves counterclockwise,
like b. The value of the local turn flag f can be updated according to the
modeling needs : it can be constant for some amount of time to impose a
particular motion at a given junction, completely random, random with
some bias to favor a direction of motion, or may change deterministically
according to any user-specified rule. For instance, figure 2.22 shows a
turn flag configuration which forces cars to move horizontally after the
junction.
As before, a vehicle moves only when its destination cell n ou t is empty.
Far from a rotary, the state of the destination cell is determined as the
occupation of the down-motion cell. This is also the case for a vehicle
turning in the rotary. On the other hand, a car wanting to enter the
rotary has to check two cells because it has no priority. For instance, car c
54 2 Cellular automata modeling
cannot enter the rotary because b is going to move to the white cell. Car e
cannot move either because it sees b (and cannot know whether or not b
will actually move). Car a, on the other hand can enter because it sees
that d is leaving the rotary and that the gray cell ahead is free.
Similarly, the incoming vehicle n in to a given cell is computed differently
inside and outside of the rotary. The light gray cell occupied by car b has
two possible inputs : in priority, it is the vehicle from the gray cell at west ;
if this cell is empty, the input will be the incoming lane, namely the car
labeled e.
Figure 2.23 shows two traffic configurations in a Manhattan-like city.
The difference between these two situations is the behavior at a rotary.
In the first case (a), on each rotary cell, a vehicle has a probability 1 /2
to exit. In (b), the turn flag f has an initial random distribution on the
rotary. This distribution is fixed for the first 20 iterations and then flips
to f = 1 f for the next 20 steps and so on. In this way, a junction acts
-
as a traffic light which, for some amount of time, allows only a given flow
pattern. We observed that the global traffic pattern is different in the two
cases : it is much more homogeneous in the random case.
A simple question that can be asked about this traffic model concerns
the so-called jamming transition. When the car density is low, all vehicles
2.2 Modeling of simple systems : a sampler of rules 55
( a) (b)
Fig. 2.23. Traffic configuration after 600 iterations, for a car density of 30%.
Situation (a) corresponds to an equally likely behavior at each rotary junction,
whereas image (b) mimics the presence of traffic lights. In the second case, car
queues are more likely to form and the global mobility is less than in the first
case.
can move because there is a high probability that the next site will be
empty. However, if the density of cars is made large enough, there will
be an increasing number of vehicles that are blocked by a car occupying
the next cell. Actually, it is expected that above a critical car density P c ,
the average velocity of the vehicles in the road network depends on the
total number of cars and decreases as the car density is increased. The
traffic congestion can be expressed by measuring how the average velocity
(v) and the global traffic flow p(v) depend on the average car density.
Precisely, p and (v) are defined as
p = N/L
where L is the total length of the road network and N = I: i n i the total
number of cars.
We observe in figure 2.24 that above a critical car density (which is
quite small when the number of crossings is large), the traffic congestion
increases. Below this critical value, the cars move freely. We also observe
that the strategy adopted at each crossing affects the global traffic flow.
Figure 2.24 (a) corresponds to a case of "free rotary," where the direction
of motion is chosen randomly at each rotary cell. Three curves are
represented, corresponding to the length of the road segments between
56 2 Cellular automata modeling
( a) ( b)
0. 35 c--,-----,--,
o ������
car density car density
0
Fig. 2.24. (a) Plot of the average car velocity versus car density in the road
network for various road spacings. (b) Plot of the traffic flow p (v) for the minimal
road spacing allowed in the model (four cells) and various strategies of motion.
two successive intersections. For the case of a road network spacing equal
to 256, we observe a clear second-order transition from a free moving
phase to a situation of slowing down (partial j amming).
In Figure 2.24 (b) the road spacing is fixed at 4 and three different
strategies of motion are considered at the rotary : (i)"traffic light" where
the behavior at each rotary alternates between two possible paths over a
period of 20 iterations ; (ii) "free rotary," as in figure (a) and (iii) "flip-flop"
which refers to a situation where each car flips the turn flag f as it moves
across a rotary cell. In other words, a vehicle imposes the behavior of the
next car to come. The initial distribution of f is random. The flip-flop
strategy is the most efficient (it permits a higher traffic flow than the
others) since it balances the traffic load equally throughout the network.
are allowed (these objects are not rigid), but the particles composing them
should not spread out in the entire space.
Here, we will restrict our attention to the motion of these strings in a
one-dimensional space and define collision rules between them.
A string can be thought of as a chain of masses linked by springs. More
precisely, it is composed of two kinds of particles, say the white ones
and the black ones, which alternate along a line. Two successive particles
along the chain are either nearest neighbors or separated by one empty
cell. Figure 2.25 shows such a chain composed of five particles.
The rule of motion is the following. The time evolution has two phases.
First, the black particles are held fixed and the white ones move according
to the rule below. Then, the white particles are held fixed and the black
ones move with the same prescription.
Let us consider the case where the white particles move. The rule of
motion is the following : the new configuration of the string is obtained
by interchanging the spacings that separate the white particles from their
two adjacent black neighbors at rest.
In other words, the motion of a white particle consists of a reflection
with respect to the center of mass of its two black neighbor particles. If
q(t) is the position of a white particle on the lattice, and q+ and q_ are
the positions of the right and left adjacent black particles (right and left
are defined according to the direction of the x-axis), then the new position
q(t + 1) of the white particle is given by
q(t + 1) = q+ + q_ - q(t) (2.40)
Of course, this rule is only valid for a particle inside the string which has
exactly two neighbors. At the extremity, an end particle moves so as to
get closer to the string center if its neighbor is two cells apart. On the
other hand if its neighbor is located on the next lattice cell, it moves away
from the middle of the string. This motion corresponds to a reflection
with respect to q+ + a, where a = 3/2 represents the length of the "spring"
linking the particles of the string together. To prevent the particles from
moving off the lattice, a has to be integer or half integer.
Thus, if the left end of the string is a white particle located at q(t), its
new position will be
q(t + 1) = 2(q+ - a) - q(t) (2.41)
58 2 Cellular automata modeling
� t=O
� t=l
� t=2
� t=3
f--{)e+(}+++O-- t=4
f--()-+-e()+++( t= 5
f----+--Oe+OeO- t=6
� t=7
� t=8
(2.43)
where the particles of a string have been labeled from 1 to N. Note that
here, the indices refer to the particles and not to the lattice cells.
The momentum of the string is given by
N
P (t) = L mi vi (t) (2.44)
i=l
2.2 Modeling of simple systems : a sampler of rules 59
where v;(t ) = q;(t + 1 ) - q;(t) is the distance particle i it will travel during
the next time step. Finally, the total string energy is
N- 1
1
E= -
2L
i= 1
[(x;+l - x; - a)2 ] (2.45)
0 time 512
Fig. 2.27. Elastic collisions between five strings o f different sizes. The system
is periodic with length 256 lattice sites (vertical axis). This figure shows the
trajectory of each string during 5 1 2 successive iterations. The total momentum is
zero.
and M /2, where M is the total mass. The smallest possible variation of
momentum for a given string is 11P = ± 1 .
I t i s possible t o define collisions between different strings s o that mass,
momentum and energy is conserved during the interaction. To this end,
we consider only strings which have the same kind of extremities (say,
white particles). Two situations are considered to deal with the conflicting
motion : (i) the left and right ends of two different strings are next to each
other and want to move past each other ; (ii) the left and right ends are
one site apart and want to move to the same site.
A way to solve these conflicts is to prevent the particles from moving
when they want to cross each other or to superpose. Momentum is
conserved in this way because, before collision, the two extremities have
opposite velocity (i.e their total momentum vanishes). Due to the collision,
the two particles are held motionless and, therefore, both have zero
momentum. With this mechanism, two colliding strings bounce back by
exchanging a quantum of their momentum. Note that such a collision may
last several time steps before each string has reversed its initial direction
of motion. These types of collisions may be referred to as elastic collisions.
They are illustrated in figure 2.27 in a situation of five interacting strings
moving in a periodic lattice 256 sites long. As we can observe, the motion
of a string is quite complex, due to the collision. In figure 2.28, we show
the position of the middle string over time. No regularity can be identify
during this interval of time which lasts more than 105 iterations.
Another type of collision can be defined. In case (ii) mentioned above,
it is also possible to superpose the two white extremities of mass 1 /2 and
2.3 Problems 61
256 .---.---,---.
Fig. 2.28. Erratic motion of the middle string of the previous figure, over a large
period of time.
time 512
Fig. 2.29. Trajectories o f five strings experiencing ballistic aggregation. The first
collision does not lead to aggregation because the appropriate conditions are not
fulfilled.
2.3 Problems
66
3.1 The one-dimensional diffusion automaton 67
N(t' r) = N N1 (t, r)
z (t, r)
) (3.5)
A 1 (t) and A z( t) are two functions to be determined. With the relation 3.4,
equation 3.3 reads
(
A(t + r) = pqexp(
exp(zk2) p exp(zkA)
)
-: ik 2) q exp( -: ikA ) A (t) (3.6)
3.1 The one-dimensional diffusion automaton 69
We then write
A(t) = exp(ico t ) ( ) AO (3.7)
(
and one is left with the condition
p exp(-ikA) - exp(icor)
q exp( ikA)
q exp( -ikA)
p exp( ikA) - exp( icor)
) A(O) = O (3.8)
A solution A(O)
=I= 0 exists provided that the determinant of this matrix
is zero, which gives a relation between co and k (the dispersion relation).
This relation is
exp(2icor) - 2p cos(U) exp(icor) + 2p - 1 = 0 (3.9)
and can be solved for exp(icor)
exp(ico ± r) = p cos(U) ± ( 1 - p) 1-
2
� sin2 (kA)
(p 1 ) 2
(3. 10)
A second step when taking the macroscopic limit is to make the lattice
spacing A and the time step r tend to zero. The quantity giving the time
evolution is the factor exp ico ± t = exp(ico ± r) tl r. Therefore, one has to
compute
- t / r) = lim
lim (exp(ico+r)
r--+0)--+0 r--+0)--+0
[
p cos(kA) ± ( 1 - p) 1 -
2
� sin2 (U
(p 1 ) 2
f'
(3. 1 1 )
This limit is easy to compute for co_, since, when A ---+ 0, the term in the
square bracket goes to 2p - 1 . As 0 < p < 1, (2p - 1 ) t/ r goes to zero when
r ---+ 0.
For co+, this limit is of the type "1 00 ". This indetermination can be
removed by taking the logarithm of the expression and differentiating
both the numerator and the denominator with respect to r.
1. ln(exp(ico+ r)) 1" a, ln(exp(ico+ r))
lm t t r--+0)--+0
lm
r--+0)--+0 r a, r
t r--+0)--+0
lim a, ln( exp( ico+ r)) (3. 12)
To carry out this calculation we need to specify how A and r are related
when they go to zero. The quantity ln( exp( ico+ r)) as given by equation 3. 10
is an expression containing A. Therefore, it is more convenient to write
t r--+0)--+0
lim a, ln(exp(ico+r)) = t r--+0)--+0
lim exp(-ico+r)aA. exp(ico+r)
dA
dr
(3. 13)
[ ]
in the calculation is
[
Ni ( t +r, r +ACi) - Ni ( r, t) = r at + +
(3.25)
where we have defined c 1 = - c2 = 1 and neglected third-order terms in
the expansion.
From the calculation we performed in section 3 . 1 .2, we know that there
is only one way to take the continuous limit A � 0 and r � 0, namely
that r "'A2 . As a consequence, A and r are not of the same order of
magnitude. For the sake of convenience, we shall write
(3.26)
We can now compare, order by order, the two sides of equation 3.23.
Using relations 3.22 and 3.25, we have, for the order O(c0 )
N �O) = N�O) (3.27)
because the left-hand side is of order c, at least. The equation has no
unique solution. The key to the Chapman-Enskog expansion is to impose
the condition that the density is completely expressed in terms of the
p
Nl(O) s
2
P = L Nfo) (3.28)
i= l
Consequently, we have
(3.29)
�l
The next order O(c) of the Boltzmann equation is iven by taking the
term Acia r N}0) of the Taylor expansion and the term Ni ) in the right-hand
side of equation 3.23. Since N}0l = /2, we obtain
p
A ( - 11 ) arp = (p - 1 ) ( -11 -1
1
)( � )
N( l )
( 3.3 1 )
2 N l)
3.1 The one-dimensional diffusion automaton 73
This equation has a solution because the vector ( 1, - 1) is an eigenvector
of the right-hand side matrix
( - 11 - 11 ) ( -11 ) - 2 ( - 11 ) (3.32)
Since this matrix is not invertible, it is not expected that an equation
such as equation 3.3 1 always has a solution. However, the magic of the
Chapman-Enskog expansion is that, precisely, the left-hand side of the
equation is in the image of the matrix. We will return to this problem in
more detail in section 3.2. For the present time, it important to notice that
the reason why the matrix is not invertible is particle conservation.
The solution of equation 3.31 is then straightforward
A
Ni(1 l = C iOrP (3.33)
4(p _
1)
The equation governing the evolution of is given by the continuity
p
equation 3.24. The order O(c 1 ) reads
2
L ACi Or N}0l = 0 (3.34)
i= 1
[
which is obviously satisfied by our solution. The next order is
Using the solution for N}0l and N}lJ, we obtain the equation
l
A1 2 ci2 ar2 Ni(0) - 0 _
(3.35)
a p+
,
A2 1
+
(1
2
a; p = o
) (3.36)
� 2(p - 1)
and, finally
at P = v a ; P (3.37)
which is the expected diffusion equation with the diffusion constant
A p2
D=� 2( 1 _
p)
(3.38)
The derivation here is particularly simple because the microdynamics of
the random walk cellular automaton is simple. For more complex rules,
such as the FHP rule we shall discuss in section 3.2, the principle of
the expansion will be identical but involves more complex steps that
were trivially verified here. For instance, the relation between A and T
is not known. In general, phenomena at various scales may be present.
In addition to diffusive behavior, convective phenomena may appear at
a scale where A "' T. In order to differentiate between these various
74 3 Statistical mechanics of lattice gas
orders of magnitude, a multiscale has to be used in conjunction with the
Chapman-Enskog method.
Fig. 3. 1 . Two-body collision in the FHP model. On the right part of the figure,
the two possible outcomes of the collision are shown in dark and light gray,
respectively. They both occur with probability 1 /2.
,-'- _y
-7\X,-/. -><
I
�
(3.47)
Fig. 3.4. Development of a sound wave in a FHP gas, due to particle overcon
centration in the middle of the system.
-Di + qD i- 1 + ( 1 - q)Di+ l
- Ti + Ti+3 (3.54)
where the right-hand side is computed at position r and time t. These
equations are easy to code in a computer and yield a fast and exact
implementation of the model. As an example, figure 3.4 illustrates a
sound wave in the FHP gas at rest. Although the initial concentration is
analogous to the situation of figure 2. 1 3, we observe here a much more
isotropic behavior.
Similarly, the particle current, which is the density p times the velocity
field u, is expressed as
6
p ( r , t ) u ( r , t) = 'L viNi ( r , t)
(3. 56)
i=l
Another quantity which plays an important role in the following derivation
is the momentum tensor II, which is defined as
6
IIap = L ViaVif3 Ni ( r , t) (3.57)
i=l
where the greek indices rx and f3 label the two spatial components of
the vectors. The quantity II represents the flux of the a-component of
momentum transported along the /3-axis. This term will contain the
pressure contribution and the effects of viscosity.
The macroscopic limit of the FHP dynamics will require the solution of
this equation. However, under the present form, there is little hope of
solving it. Several approximations will be needed.
82 3 Statistical mechanics of lattice gas
and, therefore
6 6
l:: Nfl = 0 � ....VI.
� N(t)
i - - 0' for t � 1 (3.63)
i=l i=l
These �relations will appear as a natural choice when we come to the
solution of equation 3.60.
In the case of FHP microdynamics, as opposed to the random walk,
we do not know how A and r are related in the continuous limit. On
physical grounds, we expect to observe phenomena on several time scales.
For instance, the ratio A/r is the velocity of the particles and should not
vanish in the continuous limit. In addition, dissipative phenomena (such
as viscosity) will be present too, and with a coefficient of the order A2 jr.
The difficulty is to capture, in equation 3.60, terms of the same magni
tude. We could formally introduce, as in section 3 . 1 .3, quantities A 1 , r1 and
r z such that A = d t and r c r1 + e2rz. It is, however, more appropriate
=
(3.68)
(3.70)
i=6 l i=6 l 6
L Di - L
i= l Di-1 L
i= l Di+l
= (3.7 1 )
i= l
The second equation in 3.69 is true because, due to their definition
(3.72)
and
(3.73)
Thus, after summation, all the terms cancel. Equation 3.69 is not true by
accident. It follows from the local conservation of particle numbers and
momentum during a collision.
Of course, equation 3.69 is also satisfied for the average (ni ).
Thus,
when equation 3.60 is summed over i, or multiplied by ci and then summed,
the right-hand side vanishes. This property provides the general form of
the equation of motion for p and pu.
Using equations 3.61, 3.67 and 3.68, we obtain for the order c of
84 3 Statistical mechanics of lattice gas
equation 3.60
6
and
6
L ['t" Ot1Cjer: N}0) + ACjer: C;p Otp N} )]
0
= 0 (3.75)
i= l
where we have adopted the convention that repeated Greek indices imply
a summation over the space coordinates
0(€) : (3.78 )
where n�� i s the zero-order Chapman-Enskog approximation o f the mo
mentum tensor defined in 3.57. The reader will recognize in equation 3.77
the usual conti nuity e quation, reflecting particle conservation. Equa
tion 3. 7 8 expresses momentum conservation and corresponds, as we shall
see, to the Euler equation of hydrodynamics, in which dissipative effects
are absent.
The same calculation can be repeated for the order 0 ( €'2). Remembering
relations 3.63, we find
(3.79)
and
1 1: 1: (o)
a t2 PUer: + a 1 p ller:( p) + l :'1 :'1 s<o) :'1 :'1
2
O t 1 PUa + l u 1 p u 1 y er:p y + 't"u t1 u tp ller: p
_
- 0 (3.80)
where S is a third-order tensor defined as
6
Sa.py = L V er: Vip V; N
j y i (3.8 1 )
i= l
These last two e qu ations can b e simplified using relations 3.77 and 3.78 .
Let u s first consider the case o f equation 3.79. One has
(3.82)
(3.83)
3.2 · The FHP model 85
(3.84)
which means that at the time scale T2, the density variations are negligible. ·
Similarly, since
(3.85)
[
Ot2PUrx + 01 p llrx(1)p + 2r Ut1 llrxp
(O) ( �
(O)
+ U ly Srxpy )
� ] = 0 ( 3.86)
�
OiPUrx + p llrxp +
O [ � ( EOt, n�W + a y s!jy � )l = 0 (3.87)
where we have used Ot = e011 + e20t2 and a�a = eO l rx· Similarly, equa
tions 3.77 and 3.84 yield
o , p + divpu = o (3.88)
which is the standard continuity equation. Equation 3.87 corresponds to
the Navier-Stokes equation. With the present form, it is not very useful
because the tensors n and S are not given in terms of the quantities
p and u. To go further, we will have to solve equation 3.60 and find
expressions for N}0l and N}ll as functions of p and u. However, for the
time being, it is important to notice that the derivation of the continuity
equation 3.88 and the Navier-Stokes equation 3.87 are based on very
general considerations, namely mass and momentum conservation which
ensure that I: Q i = I: i\Qi = 0. The specific collision rules of the FHP
model do not affect the structure of these balance equations. However,
the details of the collision rule will play a role for the explicit expression
of ll and S.
86 3 Statistical mechanics of lattice gas
(a
The Chapman-Enskog expansion revisited. Equation 3.90 can be solved
using the Chapman-Enskog method. Using equations 3.61, 3.67 and 3.68
for the expressions of Nj, O t and I o rrx ) in terms of e, one obtains an
equation for each order O(et ). The right-hand side of equation 3.91 can
n n ( e t ( an��-(o) )
be expanded as
i(N) =
i(N 0) ) + Nj l ) + O(e2 ) (3.92)
1= 1 1
The left-hand side of 3.9 1 is of order e 1 , at least, whereas the right-hand
side has a contribution of order 0 ( e0). The two lowest-order equations
are simply (see equations 3.74 and 3.75).
O ( e0 ) : Qi(N(O) ) 0 =
(3.93)
( ) .
and
(3.95)
and similarly
( 3.96 )
88 3 Statistical mechanics of lattice gas
( ;� ) T Eo = ( ;� ) T E 1 = ( ;� ) T E2 = 0 (3.97)
[ r
of the scalar product) to the kernel of its transpose
Im (;�) � Ker
( ;� r (3. 100)
Therefore, the solubility condition of equation 3.94 requires that + at1 N[0)
a l iXViiXN[0) Eo, E1 E2 .
N} l )
be orthogonal to and But that is precisely what
equations 3.77 and 3.78 express. Therefore, a solution is guaranteed
to exists provided that at1 N}0)
is suitably given in terms of the balance
equation. As we shall see, the appropriate choice will be to write
a N}-0) atJP + -a�-
atl Ni(0) -_ -� N[0) atJPUIX
up upu(X (3. 101)
and express at1 P and at1pu(X as given by equations 3.77 and 3.78.
Finally, note that when a solution to equation 3.94 exists, it is not
unique. As we said in section 3.2.3 we also impose the condition that
6
L iN ) 0, for t ;:::: 1
i=l i=l V V =
(3. 102)
reads
Ni(O) (O)
Ni+2 Ni+(O)4 _ Ni+(O)l Ni+(O)3 (O)
Ni+S
-
( 1 - N}0 l ) ( 1 - N�2) ( 1 - N�4 ) ( 1 - N�1 ) ( 1 - N�3 ) ( 1 - N�5 )
(3. 105)
We take the logarithm of this equation and introduce the notation
Mi
=
N �O)
1
(3. 106)
( 1 - N}0) )
so that our condition becomes
log(Mi ) + log(Mi+2 ) + log(Mi+4) - log(Mi+ l) - log(Mi+ 3) - log(Mi+S ) 0 =
(3. 107)
Note that this relation is the same for all i and we shall write it as
log M1 - log M2 + log M3 - log M4 + log Ms - log M6 0 (3. 108) =
for i = 1 and
log M2 - log M3 + log Ms - log M6 = 0 (3.111)
for i = 2. The other values of i yield relations that are a linear combination
of 3.110 and 3.111. Instead of equations 3.110 and 3.111, it is convenient
to replace them by their sum and difference
log M1 - log M3 + log M4 - log M6 = 0 (3.112)
and
log M1 - 2 log M2 + log M3 + log M4 - 2 log Ms + log M6 = 0 (3.113)
In order to combine equations 3.108, 3.112 and 3.113, it is convenient to
use the collisional invariants defined in equation 3.98
Eo (1, 1, 1, 1, 1, 1)
E1 (e ll , c2 1 , c31, C41, cs 1 , c6 1 )
E2 (3.114)
as well as the quantities
E3 (1, - 1, 1, - 1, 1, -1)
E4 (1, 0, - 1, 1, 0, - 1)
Es (1, -2, 1, 1, -2, 1) (3.115)
where Cirx denotes the spatial component of lattice direction6 vector Ci.
a
The six vectors Eo , ... , E5 form an orthogonal basis of IR because the
cis have the following properties
(3.116)
and
6
L Cirx Cif3 = 3 brxp (3.117)
i= l
The first two relations are obvious from figure 3.3. The last one can be
demonstrated as follows. First, we remark that it is equivalent to the
relation
(3.118)
where 1 is the 2 2 unity matrix and cici is the matrix whose components
x
are cirx Cif3 · Then, it is easy to check that if one multiplies the matrix I: i cic i
by any of the ck, one obtains 3ck because Ck · ck+ l = 1/2, ck · ck +2 - 1/2 =
and Ck + Ck +2 + Ck +4 = 0. With the vectors En, the local equilibrium
conditions 3.108, 3.112 and 3.113 can be written as
log M · E3 = 0 log M · E4 = 0 log M · E5 = 0 (3.119)
3.2 The FHP model 91
where log M is a notation for (log M1 , log M2 , log M3 , log M4 , log M5 , log M6)
and · also represents the scalar product in IR6 . Since the Ens form a basis
of the space, log M is any linear combination of the collisional invariants
Eo, E 1 and E2
(3.120)
or, by using the definition of Eo , E 1 , E2 and the notation b = (b 1 , b2 )
log Mi a + b · ci = (3.121)
We can obtain the solution for N(O) using 3.106 and 3.121
N(O) = 1 (3.122)
1 + exp( -a - b · ci )
�
=
1=
Ni(0) f(Xi ) - 1 + exp(-X (3.123)
i)
where the six-dimensional vector X (X1 , X2 , ... , X6) is defined as
=
=
X a(p, u)Eo + b 1 (p, u )E1 + b2 (p, u )E2 (3.124)
The expansion of N}0l around u 0 yields =
u=O
]
(3.129)
The unknowns in this relation are determined according to 3.99, which
states that
p = L Eoi · N�0) pu1 = v L E l i · N�o) pu2 v L E2i · N�0) (3.130)
=
because
Eo · Eo = 6 E1 · E1 = E2 · E2 = 3 (3.132)
and Ek · E1 = 0 if k =/= l. Similarly,
L Eoi axi axi aa aa + 3 ab 1 ab 1 + 3 ab ab2
6 aucx 2
(3.133)
i
=
aucx aup aup aucx aup aucx aup
(3.134)
and
(3.135)
because
{
L E01· EOl· Ekl· -- L. Ekl· - 06 ifotherwise
k=O (3.136)
.
I I
if k, l =I= 0 (3.137)
and
if k, l, m =/= 0 (3.138)
These results follow from the geometric properties of the lattice directions
ci and, in particular, from 3.117.
3.2 The FHP model 93
The condition p = Eo · N(0) now reads
p = 6f(x ) + 6urxarJ1 (x) + 3urxup arxpf'(x)
+ 21 urx up [6arx ap + 3 blrxblp + 3 b2rxb2p ] f" ( x ) (3.139)
( )
where we have introduced the notation
arx = (� ) a rxp iPa (3.140)
O Urx u=O =
O Urx O Up u=O
and similarly for b1 and b2 . Since 3.139 must be satisfied for all values of
p and u, one concludes that
6f(x) = p arx = 0 (3.141)
and
(3.142)
Now, since £ 1 · N(O) = p(udv), we also have
1
U l = 3urxblrxf1 (x ) + UrxUp [blrxfJ f1 ( x ) + (3arxblp + 3 blrx ap)f11 (x)] (3.143)
P -; 2
Consequently,
3buf' ( x ) f!_v=
b1 2 = 0 (3.144)
For symmetry reasons, the final condition E2 N(O) = p(u2 jv) gives
3 b22/1 ( x) = f!_v b2 1 = 0 b2rxp = 0
·
(3.145)
Finally, equation 3.1 42 reduces to
auf'(x ) + � brd"(x) = 0 a22f'( x ) + � b�2j" (x ) = 0 (3.146)
=
and the only non-zero quantities of our expansion are
f(x ) 6p bu b22 = 3vfp' (x)
= (3.147)
and
p 2 f" ( x )
au = a22 = - 18v (3.148)
2 (f' ( x )) 3
We can now substitute these results into the expression 3.125 for N}0)
p + p Ci . u - p2 u2 f"(x )
N'�O) 6 3v 36 v 2 (f1 (x)) 2
-+ -+
=
2p f"(x) up
+ 18 (jl (x)) 2 cirxCitJ -;Urx --;; (3.149)
94 3 Statistical mechanics of lattice gas
i=l
The above relation is satisfied for any particular orientation of the six
vectors Cj. The fourth-order tensor E?=l Cia Cip Ciy Cio is isotropic and this
is the reason why a hexagonal lattice is considered for a hydrodynamical
model, rather than a square lattice. Isotropic fourth-order tensors made
of the lattice directions play a crucial role in the lattice gas models
of hydrodynamics because of the convective nonlinear term u · Vu in
the Navier-Stokes equation. The isotropy requirement constraints the
possible lattices which can be used. This is a problem in three dimensions
where such an isotropic lattice does not exist. One has to consider a
four-dimensional lattice and then project in three-dimensional space. As
a result, one obtains two different speeds of particles and some directions
with two particles per links.
The result of equation 3.154 can be obtained by general symmetry
consideration [33] but a direct derivation is also possible. For this purpose,
we write
Ck = (ck l , Ck2 ) = (cos(k(J + cf>), sin(kO + cf>)) (3.155)
3.2 The FHP model 95
where
e 2n6 and ljJ [0, 2n]
= E (3. 1 56)
The quantities to be computed in relation 3. 1 54 are actually
L Ck io L Ck2• L c� l c�2 , L Ck l Ck2 , and L Ckl Ck2
k k k k k
(3.1 57)
Using �k=A rk (r n - 1 )/(r - 1) for f. 1 and exp(iO) cos e + i sin e,
= r =
one has
6 5
L (ckl + ick2 )4 L exp(4ik8) exp(4il/J)
k=l k=O
exp (4z.A.'+' ) exp(24i8) -1 =0
exp(4i8) - 1 (3. 1 58)
6 5
and
6 5
L ( Ckl + iq2 ) 2 (ck1 - ick2 ) 2 L 1 6
= = (3. 1 60)
k=l k=O
On the other hand, a directed calculation of the left-hand side of
relations 3. 1 5 8-3. 1 60 yields
6 6
L (ckl + iq2 ) 4 = L [(ck l + Ck2 - 6c� 1 c�2 ) + 2i(ck l Ck2 - Ck1Ck2 )] (3. 1 6 1 )
k=l k=l
6 6
L(ckl + iq2 ) 3 (ck1 - iq2 ) L [(ck l - Ck2 ) + 2i(ck 1 Ck2 + Ck1Ck2 )] (3. 1 62)
=
k=l k=l
and
6 6
k k k k k
(3. 1 64)
96 3 Statistical mechanics of lattice gas
(3. 1 66 )
where
3
g ( p ) = -G(p p-3
2
) = -
p-6 (3. 1 67)
ci,
- )u2) <5rxp
As we can see, this result is independent of the lattice directions because
of the isotropy of the tensor 3 . 1 54.
(
The quantity � p � g (p is called the pressure term and pg (p )urx up
the convective part of the momentum tensor. Our microscopic dynamics
(v2
p 2 2 (p)u
p -
p g 2)
gives an explicit expression for the pressure
= (3. 1 68)
Compared with the standard Euler equation, the above relation differs
because of the g(p) factor in front of the convective term ( u · V) u. The
fact that g(p) 1- 1 is of course a failure of the model to properly describe
· a fluid flow. This factor comes from the non-Galilean invariance of the
cellular automata dynamics and implies that a t urx + g(p )( u · V) u is no longer
the total derivative ( du/ dt ) .
However, at low Mach, g(p) can be assumed constant and, by renor
malizing the time t properly, one can transform a t , into g(p )a; and get rid
of the g(p) factor in the left-hand side of equation 3. 1 69.
where, t o first order i n u, IT�� = (v 2 /2)p� rxf3 (see equation 3 . 1 66). Since Peq
is constant, we have, to first order in fo:.p and fo:.u
(3. 172)
and
v2
Peqa t, ua + 2 a la !!P = 0 (3. 173)
[ ( )] -
the terms involved in the Navier-Stokes equation derived in 3.87, namely
atPUrx + ara IIrx(OJfJ + eiirx( lfJJ + 2r ea t1 IIrx(OJfJ + ara Srx(OfJJy -0 (3. 1 77)
p y
The new terms with respect to the Euler approximation of the pre
vious section are of two kinds : those involving the zeroth order of the
Chapman-Enskog expansion and II �� which implies computing N? J from
equation 3.94. For symmetry reasons, these new contributions are expected
to be odd functions of the velocity flow u. Therefore, we consider only the
lowest-order terms in the expansion of N}0l and II �� .
We first investigate the two contributions
r a
2 arp
ea t! II rx(OJfJ and (3.1 78)
(3. 1 79)
Since at1P = -divtpil (by equation 3.77), one has ea t1 P = -divpu and
r a (O J _ rv2 a
2 arp e a tj II rx fJ - 4 arp e a tjP() rxfJ
(3.1 80)
(3. 1 8 1 )
because the other terms contain an odd number of v;. Thus, we have,
using 3. 1 54,
v2p 3
3 4 (6rxp6y o + 6rxy 6p0 + 6rx06py )u0
v2
4 p( 6rxpUy + 6rxy UfJ + 6py Urx)
3.2 The FHP model 99
Therefore
T u� 2
� - Srxf3y
TV 2 TV 2 u� .
-
2- = - V2 P Urx + - -;- dtvpu
(0) �
(3. 1 83)
urp ry 8 4 urrx
Substituting the results 3. 1 80 and 3.183 into the Navier-Stokes equa
tion 3. 1 77 yields
(3. 1 84)
We observe that the contribution of the two terms we just computed has
the form of a ViSCOUS effect Vtattice V 2 pu, where
2
Vtattice = - TV (3. 1 85)
S
where Vtattice is a negative viscosity. As mentioned previously, the origin
of this contribution is due to the discreteness of the lattice (S��Y and
O t 1 IT�� comes from the Taylor expansion). For this reason, this term is
referred to as a lattice contribution to the viscosity. The fact that it is
negative is of no consequence because, the contribution - 8�13 EIT�d which
we still have to calculate will be positive and larger than the present
one.
and used equations 3.77 and 3.78 to express O t1 P and O t 1 P Urx . These
substitutions will ensure that the right-hand side of equation 3. 1 86 will be
in the image of (of!/ oN ).
100 3 Statistical mechanics of lattice gas
As we did for the lattice viscosity, we shall consider only the first order
in the velocity flow u. The omitted terms are expected to be of the order
O(u 3 ). From the expressions 3.152
and 3.166,
we have for the lowest order
in u
and (3.188)
Thus
oN �O) 1 (3.189)
_I_
ap 6 and
(3.190)
From this result, it is now clear that equation 3.186
will have a solution :
the six-dimensional vectors Qa.p of components Qa.p i given by (c ia.C ifJ - �c5a.p)
are orthogonal to the collisional invariant Eo, E1 and E2 defined in 3.98
( this is in particular due to the fact that L i C ia.C ifJ = 3c5a.p) . Since Eo, E1
and E2 are in the kernel of ( oO.joNf, then Qa.p is in the image space of
( oO.joN) ( see equation 3.100).
We can now consider the left-hand side of equation 3.186.
We compute
0
it for u = and introduce the notation
(3.191)
3.91 we obtain
From the explicit expression o f n i given b y equation
(anl ) -- -i (1 - s)3 - s3 (1 - s)2 - s (1 - s)4
oN1 u=O
_ !2 s2 (1 - s) 3 - !2 i (1 - s) 3
- -s (1 - s) 2 [s (1 - s) + s2 + (1 - sf + s (1 - s))
-s (1 - s)2
= (3.192)
3.2 The FHP model 101
Similarly, we compute the other terms and we obtain
( an )
-
1
=
- s (1 - s)
2
aN u=O 2
-2 s + 1 -3s + 1 4s - 2 -3s + 1 s+1
s+1 -2 s + 1 -3s + 1 4s - 2 -3s + 1
X
-3s + 1 s + 1 -2 s + 1 -3s + 1 4s - 2
4s - 2 -3s + 1 s+1 -2 s + 1 -3s + 1
-3s + 1 4s - 2 -3s + 1 s+ 1 -2 s+1
s + 1 -3s + 1 4s - 2 -3s + 1 s+1 -2
(3.193)
We have already mentioned that the vectors Q rxp are in Im ( We g�).
observe here that they also are eigenvectors. To see that, let us consider
the product of the first row of (an; aN) with Q rxfJ· Using
E81 {3 8p , We Obtain
=
EIIrxf3( l )
r: [ ]
3p(� : s) 3 � divpu<5rxp - i ( <5rxp <5ys + <5rxy <5ps + <5rxs<5py )8y pus
v 2 divpu<5rxp - ( 8rx pup + 8p purx )]
2p( ; _ s) 3 [ (3.198)
We can now rewrite explicitly, to first order in E and second order in
the velocity flow u , the Navier-Stokes equation obtained in 3.184. Using
expression 3.166 for II�t ' we have
[
8tPUrx + 8p (pg (p)urx up) - Vp - 8p 2p (�v� s) 3 (<5rxpdivpu - ( 8rxUf3 + 8p urx ))
= ]
(3. 1 99)
where the pressure p is given by relation 3.168.
Again, in the limit of low Mach number, the density can be assumed
to be a constant, except in the pressure term [80]. From the continuity
equation 3.88 , we then get divpu = 0 and JI( l ) contribution reduces to
1 TV2 arx8p pup + p 813u
2 rx]
-P 2p (1 - s) 3 [
TV2 V2 urx
2p(1 - s) 3
-vcon V 2 urx ( 3.200)
Finally, within this approximation equation 3. 1 99 can be cast into
8t u + g (p )(u · V )u _ !p Vp + vV2 u
= (3.201)
The quantity v is the kinematic viscosity of our discrete fluid, whose
expression is composed of the lattice and collisional viscosities
V =
(
TV 2 2p (1 � � ) 3 - �) ( 3.202 )
We observe that the viscosity of the FHP model depends strongly on the
density and may become arbitrarily large for the limiting values p = 0
and p = 6. Its minimal value is obtained for p = 3/2
Up to the factor g ( p ) = (3 - p ) / (6 - p ), equation 3.201 is the standard
Navier-Stokes equation. This term g ( p) f- 1 is an indication of the non
Galilean invariance of the model. However, as we explained previously,
3.2 The FHP model 103
Vcoll =
( �)3 )
'W 2 2p(1 �
and the mean free path.
The mean free path is the average distance a particle travels between
two collisions. If we denote by the probability of having a collision
Pk
after exactly k steps, the mean free path is defined as
00
-
The probability of having a collision after k steps is the probability of
having no collision during the first k 1 iterations and then a collision.
- Pcon)k-1Pcoll
Thus, we write
Pk = (1
where is the probability for a particle to undergo a collision.
Pcoll
Amfp APcoll k=�1 k( - Peon)k- 1 PcolA l
Thus, the mean free path becomes
= � 1 = -
104 3 Statistical mechanics of lattice gas
This quantity can be computed from the microdynamics
'r,
ni (t + r + Ai:\) ni(t, r) i( )
= +Q n
Amfp =
1
In the Boltzmann approximation and at equilibrium (u 0, Ni s
= = =
(p/6)), we obtain
Therefore,
Amfp - e ( 1 A- e ) 3
6 6
v Amf
and
Veon =
12 p
This result is important because it shows that in order to have a small
viscosity, one has to decrease the mean free path. This is achieved by
increasing the collision probability. In conclusion, the more collisions we
have, the smaller the viscosity is.
From the parabolic velocity profile (Poiseuille flow), one can deduce the
viscosity.
It turns out [84] that, as the system size gets larger, this viscosity
· increases logarithmically. Such a logarithmic divergence is expected for a
two-dimensional fluid, from general considerations of statistical mechanics.
This observation seems in contradiction with the result 3.202. The point
is that the transport coefficients, such as the viscosity given by 3.202,
are renormalized by hydrodynamic contributions (when going beyond the
Boltzmann approximation) whose effects are divergent with the system
size in two dimensions.
This problem is related to the existence of long time tails in the velocity
autocorrelation function in a fluid. It was first observed numerically
t-d/2 ,
by Alder and Wrainwright in 1968 that the velocity autocorrelation in
hydrodynamics decays as where d
is the dimensionality of the
t
system and the time. This long time tail is also well observed in lattice
gas models [85,86] .
The connection with the divergence of the viscosity is the following. It
is well known from statistical mechanics [87] that transport coefficients
(such as the viscosity) can be computed as an integral over time of
suitable autocorrelation functions. This is the Green-Kubo formalism. In
two dimensions, the velocity autocorrelation goes for long times as t- 1
and its integral diverges, as an indication that some pathology is present
in a two-dimensional fluid.
Clearly the whole collision process can be defined by giving the the full
collision matrix A.
In the example of the FHP model, there are 64 input
states and 64 possible output states and thus, is a 64 x 64 matrix. A
106 3 Statistical mechanics of lattice gas
In general, A is a sparse matrix : A(n, n') is zero for most pairs (n, n').
This is the case in particular when the states n and n' do not have the
same momtentum or do not contain the same number of particles. Since
A(n, n') is a probability one has obviously
L A(n, n') 1 = (3.203)
n'
for any n.
In some models, the collision matrix is symmetrix : = A(n, n') A(n', n)
and one says that detailed balance
holds. This is the case with the HPP
and plain FHP models. However, in general, this is not true. We may for
instance modify the two-body collision rule in the FHP model so that,
with probability 1, one has
n (0, 1, 0, 0, 1, 0) � ( 1 , 0, 0, 1, 0, 0) n'
= =
The lattice Boltzman equation. Using the collision matrix A, the asso
ciated lattice Boltzmann equation is easily written. In the Boltzmann
approximation (factorization assumption), if is the probability that
Ni
FHP model
3.2 The 107
P(n) II N? ( 1 - Nj ) ( l j )
= (3.204) -n
j
Hence
L P(n)A(n, n')
n
is the probability of output state n' and the average value N; (n�) is =
N; L: n; P(n)A(n, n')
=
'
n,n
'
n,n
Using expression 3.204 for P(n) and remembering that N; Ni(r + A.ci, t + =
n,n
' j
Fig. 3.5. The collision rule of the FHP-111 model. Before and after collision
configurations are shown. When two possible outputs are given, each of them is
selected with equal probability.
Figure 3.5 illustrates the collision rules. We observe that while mass and
momentum are exactly conserved in a collision, kinetic energy is not. A
collision between a rest particle and a moving one results in two moving
· particles and there is production of kinetic energy. However, the inverse
collision is expected to occur with the same probability and, therefore,
energy is conserved in a statistical way.
The viscosity of the FHP-III model is found to be 3 []
v � <v2 ( 4p( l _ � ; '(I _ �) - �) (3.209)
which is smaller than the viscosity of the FHP model. Therefore, increasing
the number of collisions results in decreasing viscosity. For this model,
the maximum value of R. [88]
is
R!?ax = 2.22 for p = 2 (3.2 10)
Although the FHP-III model has a better physical behavior, it is much
more complex than the FHP dynamics. From the point of view of
numerical efficiency, it is highly desirable to implement the collision rule
as a lookup table in which all the 2 7 possible collisions are computed in
advance. Otherwise the microdynamics requires an enormous number of
terms. As an example of the complexity of the collision term, we give
is smaller. Oscillations are seen along the streaklines, indicating the onset
of an instability.
Cellular automata fluids require quite large systems to exhibit non
trivial flow patterns. This is due to their relatively high viscosity which
limits the Reynolds number that can be attained. On the other hand,
we can see that the symmetry can be spontaneously broken due to the
intrinsic fluctuations in the fluid automaton.
1 12 3 Statistical mechanics of lattice gas
3.2. 7 Three-dimensional lattice gas models
Hydrodynamics is usually a three-dimensional problem and it is clearly im
portant to generalize the two-dimensional FHP fluid to three dimensions.
Although there is no difficulty in defining collision rules between particles
moving on a 3-d lattice, there are severe isotropy problems. It turns out
that no regular three-dimensional lattice exists with enough symmetry to
guarantee that fourth-order tensors such as (3. 1 54) are isotropic.
The solution to this problem has been proposed by d'Humieres, Lalle
mand and Frisch [36,88] . The idea is to consider a four-dimensional
lattice, the so-called the FCHC lattice (face-centered-hyper-cubic) which
is isotropic and then to project the motion to three-dimensional space.
In the 4-d FCHC model, there are 24 directions of motion, correspond
ing to the lines to connect a site to its nearest neighbors. These links are
of length .ji. The projection in three dimension results in six double links
of length 1 and 12 single links of length .ji. After projection, we obtain
a multispeed model with up to two particles along some directions.
The 24 lattice directions ci are the following [89] . First the double links
are
c1 = ( 1, 0, 0) c2 = (- 1, 0, 0) c 3 = (- 1, 0, 0)
c4 = ( 1, 0, 0) cs = ( 0, 1, 0) c6 = ( 0, - 1 , 0)
c 7 = ( 0, - 1, 0) cs = ( 0, 1, 0) c9 = ( 0, 0, 1 )
c1 0 = ( 0, 0, - 1 ) cu = ( 0, 0, - 1 ) c1 2 = ( 0, 0, 1 )
and, second, the single links
cu = ( 1, 1, 0) c1 4 = (- 1 , - 1 , 0) c 1 s = ( 1 , - 1 , 0)
c16 = (- 1 , 1, 0) c1 7 = ( 1, 0, 1 ) c l s = (- 1 , 0, - 1 )
c1 9 = ( 1, 0, - 1 ) c2o = (- 1 , 0 , 1 ) c 2 1 = ( 0, 1 , 1 )
c22 = ( 0, -1, -1) c 23 = ( 0 , 1 , - 1 ) c24 = ( 0, - 1 , 1 )
Figure 3 . 8 illustrates this projection in three-dimensional space. We refer
the reader to [88] for a detailed discussion of the collision rules.
Fig. 3.8. The possible particle velocities of the FCHC lattice gas model of
a fluid. Bold links represent the direction in which two particles may travel
simultaneously.
distibution we mean here that the modulus of the particle speed may vary
in addition to its direction.
With several velocity populations, kinetic energy can be defined in a
non-trivial way and a local temperature associated to the fluid through
the equipartition theorem.
Since we work on a lattice and would like to keep the dynamics to
its maximum simplicity, we will not consider a large range of different
velocities for the particles. Typically we shall have "hot" (or fast) particles
that travel to second nearest neighbor sites in one time step, normal,
speed-one particles and, finally "cold" particles at rest. These models are
usually termed multisp eed LGA.
The main difficulty with a multispeed model is to define appropriate
collision rules. We would like to conserve mass, momentum and enery
independently and we expect that collisions result in a change of the
velocity distribution. The constraint that particles are moving on a lattice
reduces very much the number of possible collisions and the velocity set
should be chosen carefully.
The first multispeed model of this type was proposed by Lallemand
and d'Humieres [37] in 1986. It is defined on a square lattice and com
prises particles of the same mass m = 1 but speed 0, 1 and J2. The
1 14 3 Statistical mechanics of lattice gas
'
Fig. 3.9. Mass, momentum and energy conserving collision in the three-speed
square lattice automaton. The reverse collision is also possible.
fast particles move along the diagonals of the lattice while the speed
one particles are identical to HPP particles. Collisions conserving mass,
momentum and energy can be defined between particles having the same
speed using the regular HPP rule. A non-trivially conserving energy
collision can be defined between particles of different speed. It is
illustrated in figure 3.9 and involves one fast, speed- J2 particle colliding
with one rest particles to produce two speed-one particles. Of course,
the reverse collision, in which two speed-one particles collide at right
angle and give one rest particle and one fast particle, is also taken into
account.
Energy is defined as pure kinetic energy
E = m-2v2
Rest particles have zero kinetic energy, speed-one particles have energy
1 /2 and fast particles have energy 2. With this definition, it is easy to
check that the collision described in figure 3.9 conserves mass, momentum
and energy. It is also interesting to note that this collision rule modifies
the velocity distribution : a pair of particles with speed 0 and J2 are
transformed into two particles of speed one. Thus rest particles couple
the population of speed-one and speed- J2 particles.
Of course, since this model is defined on a squre lattice we expect that
isotropy problems will show up with fourth-order tensors. Thus this model
is not appropriate to describe a fluid flow but, on the other hand, may
well simulate a fluid at rest, submitted to a temperature gradient. An
interesting application is the study of nonequilibrium fluctuations in such
a system [90] .
In section 3.3.3, we shall present other multispeed LGA that are defined
on a hexagonal lattice and can simulate isotropic flows.
3.3 Thermal lattice gas automata 115
The right-hand side term 0; is the collision term which now obeys an
extra conservation law for energy
Of course the usual mass and momentum conservation are still required
L m;O; L m;vi ni = = 0
i i
Clearly, in a one-speed model, energy and mass conservation are identical.
Q
The precise expression for depends on the model, the type of collision
and the lattice topology. But the three above properties must be satisfied
in order to describe a thermal cellular automata fluid.
The standard procedure can be repeated to obtained the macrosopic
equations related to the LGA microdynamics : ensemble average, Boltz
mann factorization and multiscale Chapman-Enskog expansion. The fluid
density p and velocity field u are defined as previously
p u :L m;v;N;
=
i
where N; = ( n i) is the averaged occupation number.
A new physical quantity is introduced : the local energy per particle e
l
2
1 16 3 Statis tical mechanics of lattice gas
A temperature T can be introduced in this model, assuming a local
version of the equipartition theorem [87] is valid. Thus we define, for a
two-dimensional system
e = kT
where k plays the role of the Boltzmann constant. Taking
Q=
J-+ '"""' ( Vj
-+
-+ )2
-u -+ -+
( Vj - U
)
i
N
.
L.....,;
I
2
It is a rather tedious task to obtain the explicit thermohydrodynamic
equations governing p, u, e, starting from a given model and we will not
develop further this derivation. In general, we expect a loss of Galilean
invariance but, in some models [8 1], tricks may be used to remove this
problem.
As an illustration of the properties of a thermal model, it is also
interesting to consider the simpler case of a fluid at rest, in a stationary
state and submitted to a temperature gradient. In this case we typically
expect the following result [78]
JQ = - KVT
where K = K(p, T ) is the thermal conductivity which is built-in in the
collision rule.
I •
� < f-.
J..L= 3, n=(O,-...J 3), e=9/2 J..L= 3, n=(O, ...J 3/2), e=3
�\ l
I L. •
in lattice gas models. We have already observed that the HPP model
described in section 2.2.5 conserved momentum along each lattice line and
each lattice column. We have mentioned that both two- and three-body
collisions are necessary in the FHP rule to avoid spurious conservations.
These kinds of undesirable conservations are related to the choice of
the collision rule and generally show up when performing the Chapman
Enskog expansion. However, in many cellular automata models, there are
other types of spurious conserved quantities, which are an artifact of the
discrete nature of the underlying lattice and have no physical counterpart.
They are the so-called staggered in v ariants, first discovered by MeN amara
and Zanetti [92]. They have received this name because they usually
depend on the parity of space and time.
For instance, we have already noticed in section 3.1.4 the existence of
a spurious invariant directly related to the checkerboard structure of the
square lattice
Neven(t) - Nodd( t) (-1) tf' [Neven(O ) - Nodd(O)]
= (3.212)
The checkerboard invariant generalizes to almost all lattice gas automata.
In the FHP model, this unphysical conserved quantity is called the stag
gered momentum.
Consider for instance figure 3.12, and more particularly the horizontal
line of sites labeled + The particles sitting on this line at time t will
" ".
either stay on this line if they travel horizontally or move to one of the
horizontal lines labeled "-", above or below. Now, suppose we compute
3.4 The staggered invariants 1 19
+ +
+ +
+ +
the total momentum along the y-axis (vertical direction) carried by each
particles. The particles moving horizontally have, of course, no vertical
component of momentum and do not contribute. Therefore, all the v ertical
momentum of the line will be distributed, at time t 'L, on the two
"+" +
horizontal "-" lines.
If we continue to label the horizontal lines by alternating and "-"
we obtain similarly that all the vertical momentum of a "-" line at t is
"+"
transferred to the surrounding "+"
lines at the next step. We can build a
where L+ and L_ denote the two sublattices obtained with our line
labeling and Py represent the y-component of momentum. If we add the
two equations, we obtain no more than the expected physical conservation
of momentum. But, by subtracting them, we obtain the typical form of a
staggered invariant, namely
Here, we see that the staggered momentum results from the detailed way
in which the momentum is conserved on the lattice.
Of course, the same construction we consider for the vertical component
of the momentum can be repeated for the lattice lines orientated 60 and
1 20 degrees with respect to the horizontal direction. As a result, we obtain
1 20 3 Statistical mechanics of lattice gas
a total of three staggered invariants, which are usually expressed as
G0 = L g0 ( t, r ) = L( -1) (t/Tl+ B· r L lJ · vi i r, t)
n( (3.214)
rEL rEL
where L is the set of lattice points. The quantity Li lJ vi i r t)
· n( , is the
momentum along lJ , which is defined as
lJ = ( 1 /Jc)(- sin(n( n /3)), cos(n( n /3))) n = 0, 1, 2 (3.2 1 5 )
For each value o f n, lJ i s orthogonal t o one o f the three families of
line orientations. The scalar product lJ · r specifies one sublattice division
because it take the same integer value (even or odd) on a line perpendicular
to lJ .
The local staggered invariants, such as g 0 often show a diffusive behavior
but sometimes may exhibit propagating modes, as shown by Ernst and
Brito.
Staggered invariants may play an important role in the physical behav
ior of a lattice gas model because they couple to the nonlinear terms in
the Navier-Stokes equations [84,92-94] . For instance the u · Vu term will
be accompanied by an analogous term containing the average staggered
momentum density g 0 . However, no direct evidence of pernicious effects
have been detected in fluid flow simulations, at least qualitatively.
The influence of a staggered invariant on the physical behavior of a
cellular automata fluid can be minimized by properly choosing the initial
condition of the fluid particles [84] . It is indeed always possible to prepare
an initial state in which the total staggered momentum G0 vanishes exactly.
In addition, it is possible to chose the initial positions of the particles so
that the staggered momentum coincides with the regular momentum. For
this purpose, they are only placed on the lattice of spacing 2Jc shown in
figure 3 . 1 3. The other sites are left empty. These positions correspond to
choosing the intersection of the L + lines for the three orientations. Along
the bold lines of figure 3 . 1 3, one has
(3.2 1 6)
and, at time t = 0, G0 reads
G0 = lJ i lJ · Ptot
· LL
rEL i v irn ( , 0) = (3.21 7)
where Ptot denotes the total momentum of the system. Since both G0
and i\ot
are invariants, the fact that they are equal at the initial time
guarantees that they are always equal.
Usually, there are many staggered invariants in a lattice gas automa
ton [95] and there is apparently no general way to find all of them. One
3.4 The s taggered invariants 121
Fig. 3 . 1 3 . The bold lines represents the sublattice for which the total staggered
momentum reduces to the usual momentum.
'
Fig. 3. 1 4. Collision between a fast (long arrow) and a slow (short arrow) particle
in the two-speed HPP model. Mass, momentum and energy are conserved during
this interaction process.
Of course, the fast particles also yield the usual checkerboard dynamical
invariant,
) _ 1 ) t/1:H · il nfast t
L L (r, ) (3.219)
rEL i=l,4
3.5.1 Introduction
Cellular automata fluids, such as those discussed so far in this chapter,
represent idealized N-body systems. Their time evolution can be performed
exactly on a computer, without many of the approximations usually done
when computing numerically the motion of a fluid. In particular, there is
no need, in a lattice gas simulation to assume some factorization of the
many-body correlation functions into a product of one-particle density
functions.
Of course, the cellular automata model may be inadequate to represent
a real situation but it includes naturally the intrinsic fluctuations present
in any system composed of many particles. This features is out of reach
of most tractable numerical techniques. In many physical situations,
spontaneous fluctuations and many-particle correlations can be safely
ignored. This is, however, not the case if one is interested in long time
tails present in the velocity autocorrelation (see section 3.2.3). We will
also consider in the next chapter other systems for which correlations or
3.5 Lattice Boltzmann models 123
local fluctuations cannot be neglected and where the N-body nature of
the cellular automata approach is necessary.
On the other hand, a cellular automata simulation is very noisy (because
it deals with Boolean quantities). In order to obtain the macroscopic
behavior of system (like streaklines in the flow past an obstacle), one
has to average the state of each cell over a rather large patch of cells
around it (for instance a 32 x 32 square) and over several consecutive time
steps. This slows down very much the simulation speed and requires large
systems. Therefore, the benefits of the cellular automata approach over
more traditional numerical techniques becomes blurred [3 8 ] .
In addition, due to its Boolean nature, cellular automata models offer
little flexibility to finely adjust external parameters. Tuning is done through
probabilities, which is not always the most efficient way.
When correlations can be neglected and the Boltzmann molecular chaos
hypothesis is valid, it may be much more effective to directly . simulate on
the computer the lattice Boltzmann equation
(3.220)
with Qi given, for instance, by (3.91) or (3.211). It is more advantageous
to average the microdynamics before simulating it rather than after. The
quantities of interestNi are no longer Boolean variables but probabilities
of presence which are continuous variables ranging in the interval [0, 1].
This approach reflects the fact that, at a macroscopic level of observation,
the discrete nature of the fluid particles vanishes.
A direct simulation of the lattice Boltzmann dynamics was first con
sidered by McNamara and Zanetti [51]. It considerably decreases the
statistical noise that plagues cellular automata models and considerably
reduces the computational requirements. As mentioned previously, the
main drawback is that this approach neglects fluctuations and many-body
correlation functions.
Strictly speaking, one can argue that a dynamics based on a set of
continuous variables is no longer a CA model because a real number
requires an infinite amount of information to be stored in a computer.
However, in practice, real numbers are approximated using a finite number
of bits (typically 32 or 64 in modern computers) and, therefore, the lattice
Boltzmann approach is also a cellular automata rule such that the state
of each cell is a large (but finite) number of bits. Of course, the use of a
lookup table is not possible and a direct calculation of the rule should be
repeated for each site and each iteration.
The lattice Boltzmann method or lattice Boltzmann equation (often
referred to as LBM and LBE) have been widely used for simulating various
fluid flows [96] and is believed to be a very serious candidate to overcome
traditional numerical techniques of computational fluid dynamics. Their
1 24 3 Statistical mechanics of lattice gas
flr Jo£\, t
+ + • ) - fi fr , t) = z (!}0l (r, t) - fi fr, t)) (3.22 1 )
or, equivalently as
(3.222)
where the local equilibrium solution f}0l is now a given function of the
density p = 1: fi
and velocity flow p u = 1: f/;i·
The function f}0l can
be chosen so as to produce a given behavior, first to guarantee mass and
momentum conservation but also to yield the appropriate local equilibrium
probability distribution. In particular, the lack of Galilean invariance that
plagues cellular automata fluid can be cured, as well the spurious velocity
contribution appearing in the expression 3 . 1 68 of the pressure term. In a
more general context, f}0l could include other physical features, such as a
3.5 Lattice Boltzmann models 125
local temperature. Equation 3.221 is often referred to as the Lattice BGK
(or LBGK) method [102] (the abbreviation BGK stands for Bhatnager,
Gross and Krook [103] who first considered a collision term with a single
· relaxation time, in 1954).
Equation 3.221 has been studied by several authors [104], due to its
ability to deal with high Reynolds number flows. However, one difficulty
of this approach is the numerical instabilities which may develop in the
simulation, under some circumstances.
Finally, let us mention, in a similar context, the multiparticle lattice
gas which is an intermediate approach (see section 7.3) between a pure
automata and a real-valued description and allows an arbitrary number
of particles in each direction [105]. This technique is not yet very well
explored and could possibly alleviate some of the instability problems and
restore the many-body correlations.
3.5.2 A simple two-dimensional lattice Boltzmann fluid
In this section, we present a detailed derivation of a lattice Boltzmann
dynamics of a two-dimensional fluid. A natural starting point would
be to consider again the hexagonal lattice used for the FHP model.
However, in most computers (and especially on parallel computers), the
data structure representing a hexagonal lattice is more difficult to work
with than a regular square lattice which is naturally implemented as an
array of data. Although simple coordinate transforms can be devised to
map the hexagonal lattice onto a square lattice (see section 7.2), boundary
conditions should be treated in a separate way.
The choice of a hexagonal lattice stems from isotropic requirements. Ac
tually, the FCHC model discussed in section 3.2.7 provides an alternative
since it can be projected onto a two-dimensional square lattice.
Figure 3.8 suggests that the resulting lattice is composed of eight ve
locities i\ with and collinear with the horizontal and vertical
v1v,2V, 3v,4,vVs 6 V7vs
directions, while and correspond to the diagonals of the lattice.
These eight velocities are shown in figure 3.15.
Clearly, one has, for odd i,
(3.223)
r
where, as usual, 2 and are the lattice spacing and the time step, respec
tively.
Since the square lattice results from the projection of the 24-particle
FCHC model, all velocities do not have the same weight. Actually, four
particles give the same projection along the main lattice directions. This
fact can be taken into account by assigning a mass mi 4 to the horizontal
=
126 3 Statistical mechanics of lattice gas
4
1
(v -2 ) [(x
· + (x · where v (..1./r) is the modulus of i , for i odd.
= v
Consequently one must have
(3.227)
and similarly for and This is also true for v i and v i+2 when i is even,
vs v7.
except that the coefficient v 2 is replaced by 2v 2 , because of equation 3.223.
3.5Lattice Boltzmann models 127
Therefore, with mi = 4 and mi = 1 for i odd and even, respectively, we
obtain
8
L miVirxVip = 12v 2 brxp ( 3.228 )
i=l
The fourth-order tensor I: i miVirxVifJViyVi.5 can be computed explicitly by the
method described by equations 3.155 through 3.164 . Since the summation
can be split into the odd and even values of i, we first compute
4
Trxpy .5 = L CjrxCipCiyCi.5 ( 3.229 )
i=l
where, for this calculation, c 1 , c2 , c3 and c4 are orthogonal unit vectors
defined as
ck = (cos(ke + lj>), sin(k8 + 4>)) ( 3.230)
E
with e = n /2 and 4> [0, 2n ]. Thus, we have
4
L (ckl + iq2 )4 = 4 exp(4i l/>) (3.231)
k=l
4
L (ckl + ick2 ) 3 (ckl - iq2 ) = 0 (3.232)
k=l
and
4
L ( Ckl + ick2 ) 2 (ckl - iq2 ) 2 = 4 ( 3.233 )
k=l
Therefore, we have
4
L [(ck l + ci2 - 6c� 1 c�2 )] = 4 cos 44> (3.234)
k=l
4
L [4(ck 1 Ck2 - ckl ck2 )] = 4 sin 44> ( 3.235 )
k=l
4 4
L [(ck l - Ck2 )] L [(ck 1 Ck2 + ck1 ck2 )] = 0
= (3.236)
k=l k=l
and
4
L [ck l Ck2 + 2c� 1 c�2] = 4
+ (3.237)
k=l
128 3 Statistical mechanics of lattice gas
obtain
8 8
2: mk Vk t Vk2 2: mk Vk1 Vk2 0
= = (3.242)
k=1 k=l
and
8 8
L mk vk1 L mk vk2 12v 4
= = (3.243)
k=1 k=1
All these results can be combined to give
8
L miVirx VifJViy ViJ 4v4 ( <5rxpt5yb t5rxyt5pb t5rxbt5py )
= + + (3.244)
i=l
3.5 Lattice Boltzmann models 129
The local equilibrium. The lattice Boltzmann dynamics is given by equa
tion 3.221
+ + r) - fi('f, t) = ni = � (!}0) (1\ t) - fi('r, t))
flr TVi, t (3.245)
We will find a solution to this equation using, as before, a Chapman
Enskog expansion
fi = f}0l + cf} 1l + ...
The main difference is that, here, f}0l is given, or more precisely, chosen,
in an appropriate way.
A natural choice for specifying f}0l as a function of the physical quan
tities p and pu is to adopt a similar expression to that obtained for the
FHP model, namely equation 3.152. Accordingly, we define
b � ·� u2 h VirxVif3UrxUf3
fi(O) = ap 2 v PVi
+ u p e 2
v P
+ + 4
v (3.246)
where a, b, e and h are coefficients which will now be determined. There
are several constraints to fulfill. First, mass and momentum conservation
1mpose
8 8
L: mi ni = O and L: mi vi ni = O (3.247)
i=l i= l
This implies that
8 8
L: md}0l = p and � � �
6 miVi�i(0) = pu (3.248)
i =l i= l
because p and pu are defined through relation 3.224. This choice is also
compatible with the Chapman-Enskog expansion which requires that the
local equilibrium solution completely defines the physical variables (see
equation 3.62). A second condition we will use to determine a, b, e and
h is the Galilean invariance. This is a freedom which is possible in the
framework of the lattice Boltzmann method.
Using the fact that I: mi = 20, I: vi = 0 and I: miVirxVif3 = 12v 2 brxp, we
obtain
u2
20ap (12h + 20e)p 2
+ v
i= l
12bpurx (3.249)
130 3 Statistical mechanics of lattice gas
(3.251)
In order to satisfy Galilean invariance, the coefficient in front of the
quadratic term purx up must be 1. Accordingly, we must have
h=!
8
Since 12h + 20e = 0, we obtain
e = --
3
40
and the expression for II�� reads
II�� = pbrxp + PUrxUp (3.252)
( ) (1 2 )
where the pressure is defined as
h u2 3 u2
p = 12v 2 a + ( e + 4) p = 5 v2 - 3v p (3.253)
v2 2
Using the fact that II�� can be written as (see 3.176)
II�� = c; p brxp + O(u2)
we conclude that
s
12av2 = �v2
c2 =
5
We notice that a contribution proportional to u2 is present in the
expression of p. This is an artifact of the model which has no physical
meaning. This contribution can be removed when rest particles are
introduced in the system [101].
3.5 Lattice Boltzmann models 131
The Navier-Stokes equation. In section 3.2.3, we obtained the following
result (see equation 3.87)
[ r
OtPUa + 0[3 IIaf3 + 2 (EOt1 ITa(O{3) + Oy Sa(O{J)y -_ 0 )] (3.254)
= !l
where t + � and r r tf takes into account the different time scales
= €
E E
of the problem (see section 3.2.3).
The derivation of 3.254 relies only on the fact that I: mi ni 0 and =
I: mi !J.i vi 0 (in section 3.2.3, mi 1, for all i). Therefore, this equation is
= =
)
where O l a denotes the derivative with respect to the oc-component of r 1 .
Since Qi ! (!�0) (r, t) - fi(r, t) , the above equation simply reads
=
- 1 (1)
r � fi Ot Ji(0) + O l aViafi(0)
=
up ::1
u p Ua
with div 1 L a O l a ·
=
Thus,
al� o)
-- = a
op
132 3 Statistical mechanics of lattice gas
and
( 3.256)
( - )
Thus
2
!? ) = -T� � viy Vio a� <5y o a ly PUo ( 3.25 7)
t
Using EO y ay , the order O(E) contribution to ll reads
=
8
Enrxf3( l ) E L md[1 l virx Vif3
i=l
[
T� adivpu L miVirxVif3 - � L miVirxVif3Viy Vio ay p usJ
Tv2 � [(12a - 4b) <5 apdivpu - 4b( ap purx + oa pup )] ( 3.258)
where we have used expressions 3.228 and 3.244. We then obtain
ap Ell�� -Tv 2 � [(8b - 12a) Oa divpu + 4b a�pua ]
= ( 3.259)
From this expression, we obtain two viscosity coefficients (shear and bulk
viscosity), as is usual in compressible fluids.
Our final step is the calculation of the lattice viscosity. The first term in
3.254 giving a contribution to the lattice viscosity is 8p ( u/2) ar1 ll��- With
ll�� = 12av2 p<5 ap + O ( u2 ), we have
TE o) .
2 a t1 n rx( /3 = 6 TV 2 aEa t1 pu rxp - 6 TV 2 au ap dIVp u
s: s:
�
=
where we have used ar1 p + div 1 pu = 0 (see equation 3.77) and the definition
of the length scale Ediv 1 div. =
Therefore, we have
TE . U
(o) = - 6TV 2 aa rx dlVp
2 af3 a t1 n rxf3 (3.260)
�
l
1�
8 t u (u · V)u !p Vp VJb V2 u
+ =
_
+ (3.264)
where VJb is the kinematic viscosity of our lattice Boltzmann fluid
VJb �-=
T� 2 ( � ) (3.265)
From this result, we observe that the lattice Boltzmann viscosity is in
dependent of the particle density p. In addition, � is a free parameter
which can be tuned to adjust the viscosity within some range. We can
see that when � is small, relaxation to f (O) is fast and viscosity small.
<
However, � cannot be made arbitrarily small since � 1/2 would imply a
negative viscosity. Practically, more restrictions are expected, because the
dissipation length scale should be much larger than the lattice spacing.
134 3
. . . . . . . . . .•
Statistical mechanics of lattice gas
. .
-·
.
.. ... .. .. . - - · ·
.. . . ...
-
· · · - - · - . . .. . ..
C
- · ·· · ·· · · ·· · · ·· ····
- -
n
Fig. 3 . 1 6. Laminar flow past a disk obtained from the 2-D lattice Boltzmann
model. System size is 5 1 2 x 1 28, � = 10 and u00 = 0. 1 .
·. ·. l·
..-- · .
. .
Fig. 3 . 1 7. Non-stationary flow past a plate obtained from the 2-D lattice
Boltzmann model. System size is 5 1 2 x 1 28, � = 1 . and u00 = 0.025. From lef� to ·
right and top to bottom, the figure shows the different stages of evolution.
3.6 Problems
4.1 Introduction
Po p p
Fig. 4. 1 . How the entering particles are deflected at a typical site, as a result of
the diffusion rule. The four possible outcomes occur with respective probabilities
Po, Pt . pz and P3 · The figure shows four particles, but the mechanism is data-blind
and any one of the arrows can be removed when fewer entering particles are
present.
there is a ballistic motion. On the other hand, when J.l2 = 1 the particles
bounce back to where they came from (rotation by 180° degrees). The
other two cases (/-l l = 1 or J.l 3 = 1) give clockwise and counterclockwise
90° rotations. We also observed that if J.l l = J.l3 = 0 everywhere on the
lattice, equation 4.6 reduces to two one-dimensional random walks along
the horizontal and vertical directions.
The J.lt are 1 with probability pc, otherwise they are zero. Consequently
their average value is
=
(J.lc ) Pt (4.7)
For symmetry reasons, it is natural to impose the condition that
(4.8)
because the particle is equally likely to be deflected clockwise as counter
clockwise. Thus, the probabilities Pt satisfy
Po + 2p + P2 1 = (4.9)
It is interesting to notice that a model similar to this one has been
4.2 The diffusion model 143
considered by Hauge [109] , for the case of continuous space and time.
We shall now derive the macroscopic limit of equation 4.6 following the
same procedure as developed in section 3.1.
The operation o f random rotation i s spatially uniform and data-blind,
i.e. it is produced by an external mechanism which is the same for all sites
and which is statistically independent of the actual occupation numbers
ni of particles at a given site. Consequently,
( J.1t ni+t l = Pt (ni+tl = Pt Ni+t (4. 10)
and equation 4.6 simply becomes
Ni('r + A.£\, t + -r) - Ni('r, t) = (po - 1)Ni + Pt Ni+l + P2 Ni+2 + P3 Ni+3 (4. 1 1)
This equation is the lattice Boltzmann equation associated with the dif
fusion rule. As opposed to the hydrodynamical situation discussed in
the previous chapter, it is exact and has not required any factorization
assumption. This is due to its linearity, of course.
We now write the left-hand side as a Taylor expansion
( l
P2 P3 Po - 1 Pl
Pl P2 P3 Po 1
-
and
N1(O)
(0)
N(O) - NN2(0) (4.15)
3O
N4( )
Since L Pt = 1, N (O ) oc Eo is clearly a solution, where Eo = (1, 1, 1, 1)
i s the collisional invariant which expresses particle conservation. Except
for pathological values of the Pt this is the only solution. Using the
Chapman-Enskog requirement
4
p = I: NfOl (4.16)
i= l
we obtain that
N (O) = !!__4
l (4.17)
The next order O(c) in equation 4.1 1 is
O(c 1 ) : -i (ci · 8r )N{0l = I: nij NY l (4.18)
j
In order to solve this equation, it is interesting to notice that
3
2: PtCi+t = (1 - 2p)c i ( 4.19)
t=O
Indeed, we have ci = -ci+ 2 and
=
po c i + pci+ l + P2 Ci+2 + pci+3 (po - P2 )ci
Therefore, the vectors E1 and E2 defined as
(4.20 )
4> =
-
A (4.22)
8(p + P2 )
This yields the solution
A
8( p + P2 ) CiaOa P (4.23)
Note that when p + P2 = 0, this derivation is not possible. But, then, we
are in the situation where p0 = 1 which corresponds to a ballistic motion
without any direction change.
The equation for p is then obtained in the usual way, by summing over
i equation 4.1 1. Of course, I: Qi = 0 because particle number is conserved.
Using the Taylor expansion 4.12, one has, at order O(e)
4
O(E) : L [A(ci · or)N}0l ]
0 (4.24) =
i= l
which is obviously satisfied since N}0 l p/4 and I: ci 0. More interest
= =
· · = 0 (4.25)
With 4.23, the equation reduces to
(4.26)
1 46 4 Diffusion phenomena
Since 2::: c iaCif3 = 2l5ap , as obtained in section 3.5, we have
O t P + div [ -D gradp] 0 = (4.27)
where D is the diffusion constant
D
-1
=�
2 ( 4(p +
1 1 ) =
-1 2
�
( 4[1 -p +(p Po+ Po)] ) (4.28)
P2 ) -4
and J = -Dgradp is the particle current. It is interesting to observe that
J does not coincide with the expression 2::: N/J i which is the standard
definition of the particle current. Indeed, we have
gradp (4.29)
where 8 n is the angle formed by the velocity vector at iteration n. One has
148 4 Diffusion phenomena
where dj E {0, ±n/2, n } is the random variable giving the rotation at time
j . With these definitions and the notation v( m) = (A./r)e - i8m , we can write
t- 1
(J2 = r2 L (v(n)v(m))
n, m = O
Therefore,
and we obtain
t- 1
(J 2 ( t) A2 t + 2A.2 L (Po - P2t-m
n>m
2 2 t- 1 k
A. t + 2A. L L (Po - P2 ) 1
k= 1 1 = 1
[k
A.2 t + 2A.2 (po - P2 ) � ((Po --Pl)) - 11 ] (4.31)
k= 1 Po P2 -
After summation over k, it is found that the mean square displacement is
(4.32)
provided that the lattice is infinite or the time short enough so that the
particle has not reached the limit of the system. This relation leads
4.2 The diffusion model 149
to a value of
L (v(j)i3(0))
j=1
00
the lattice are properly chosen. On a cubic lattice, we have six possible
directions of motion which can be mixed in several different ways, in
order to produce a random walk. Among the set of all the possibilities,
one of them gives equations very close to the two-dimensional case. For
this reason, we shall not repeat the details of the calculation. We shall
simply describe how to produce the random walk and write down the
corresponding macroscopic equation.
Let us first label the three main directions of the lattice c1, c2 and c 3 ,
with the property that c 3 = c1 x c2 . The three other directions are defined
according to the relation
i being wrapped onto { 1, 2, 3, 4, 5, 6 }
Now, we consider six operations �. k
0, .. , 5, that occur with probability
=
The results obtained in the previous section are exact in the limit of
an infinite lattice with Ji ---+ 0, r ---+ 0 and Ji 2 jr ---+ const. In practice,
however, one never performs numerical simulation in this limit and some
corrections to Pick's equation are expected. In this section, our purpose
is to solve the lattice Boltzmann equation for finite systems, with various
boundary conditions. As a result, we will be able to determine the length
and time scale for which a fair modeling of diffusion is obtained.
4.3 Finite systems 151
Since the cellular automata dynamics is linear, the lattice Boltzmann
equation is exact and also linear. For this reason, exact solutions can be
obtained and discrepancies with the continuous limit discussed rigorously.
(
and
where
b=
1 - aL
2 =! 1+
2
1
1 + t (4D. + 1 )
D . i s the dimensionless diffusion coefficient
)
D. = .!_2 D = P + Po = P + Po
2 4(p + P2 ) 4 ( 1 - P - Po )
-
Therefore, we obtain
1
NI(X' y ) - 1 - 1 + t(4D. X
+ 1) L
1 (L - x)
N3 (x , y ) _- 1 + t (4D.
+ 1) L
and finally, since b = (1 - aL ) /2 ,
N2 = N4 = 21 (N1 + N3 )
We observe that, in the limit A. � �
0, a -(1 / L) and b = 1. Thus, the
density
p = N1 + N2 + N3 + N4 = 4ax + 2 ( 1 - aL)
tends to 4 ( L - x) /L , as expected for the solution of ( d2 pjdx2 ) = 0 with
the condition p (x = 0) = 4 and p (x = L) = 0.
In contrast, when A. is finite, some corrections (of the order £ ) are
present. However, the density profile is always a straight line. But the
values at x = 0 and x = L differ from 4 and 0, respectively. At x = L/2 ,
the discrete and continuous solution intersect, for the value p = 2.
Finally, it is interesting to compute, in this simple case, the particle
current j = v[N1(x + A.) - N3 (x) ] , that is the net number of particles
that have crossed an imaginary vertical plan located at x + ( A./2 ) (note :
N1(x + A.) is the probability of having a particle move from x to x + A.
'L
during the time while N3 (x) is the probability of having a particle move
from x + A. to x during the same time interval ; this not equivalent to
N3 (x + A. ) - N1(x)) .
We have
A. 22
J = - (1 + aL + a.A.) = 4-D.
1
= 4Da
'L L + A.(4D. - 1) -
'L
Since
�
4.3 Finite systems 153
we find that the particle current is
j = -Dgradp
This relation, which is the usual one in a diffusion process is true even if
the lattice spacing does not tend to zero.
Q= ( po1 --Pol
1 - po
Po - 1
)
In two dimensions, T is diagonal and Tii is composed of either a transla
tion T±x or a translation T±y along the y-axis. The collision matrix Q is
given in equation 4. 14.
The Q matrix can be diagonalized easily. For the sake of simplicity, we
shall consider the one-dimensional case. We introduce the variable M
N = PM
where P is the transformation matrix
1
p = J2 1 1
(
1 -1 ) (4.38)
1 54 4 Diffusion phenomena
( )
Its inverse reads
p -1 = _1 1 1 (4.39)
J2 -1 1
and one has for the two components of the column vector M
M1 = J21 (N1 + N2 ) = Jl
1p
and
1 ( - N1 + N2 ) = - -pu
M2 = J2 1
-
Jlv
One immediately checks that
p - 1 (Q + 1)P = D =
(� o
2po - 1
)
The microdynamics of the diffusion rule reads
TN(t + ) = (Q + 1)N(t)r
M1 (t + r )
1 2po - 1
2 ( Tx + T_x)Ml (t) - 2 ( Tx + T_x)M2 (t)
1 1
M1 (t + r) 2 ( Tx + T_x) Ml (t) + 2 ( Tx + T_x)M2 (t) (4.42 )
and by simple substitution of ( Tx + T-x)M2 (t) from the second line into
the first, we obtain
M1 (t + r) = Po
2 ( Tx + T_x) Ml (t) - ( 2po - 1)Ml (t - r) (4.43 )
In order to put this equation in a more convenient form, we introduce the
discrete differential operators
1
8A t Ml (t) = (M1 (t + r) - M1 (t - r))
2r
and
1
VA 2 M 1 (t, x) = ,P (M1 (t, x + A) - 2M(x, t) + M1 (t, x - A))
These finite difference operators tend to their usual counterpart in the
limit A � 0 and r � 0. They also correspond to the standard way of
discretizing space and time derivatives.
Equation 4.43 can be written as
� (M1 (t + r ) - M1 (t - r)) + b(M1 (t + r) - 2Ml
+ M1 (t - r)) = Po (Tx + T_x - 2 )Ml
with
a = 2(1 - Po ) b = Po
Therefore, we obtain
8A t M1 + rPo aA t2 M1 = A2 Po_ VA 2 M1
2(1 - Po ) � 2( 1
Po )
We have p = J2M 1 , D = (A2 /r ) [po / ( 2 - 2p0 ) ] , the diffusion coefficient
we defined previously and c2 = (A2 /r 2 ), the speed at which information
travels in the system. With these definitions, we obtain that the density of
particles obeys a discrete telegraphist equation
A D A2 = A 2
8 t p + 2a
c t
p DV p (4.44)
156 4 Diffusion phenomena
. [ (
or equivalently
e1wr = exp - -c2· 1 +
2D.
where D. = (r/A.2 )D and c. = (r/A.)c are the dimensionless diffusion
coefficient and speed of sound, respectively. The above expression is the
dispersion relation for the telegraphist equation. It can be checked against
4.3 Finite systems 1 57
D,= 1 . 5 D,= 1 .5
�
� o r-
* I
/
/
� 1
0 kf...
___ __________,._,
1t 0 ---- ---
- 1 '::-
kf...
-
:-;;----
-------"
1t
Fig. 4.3. Comparison of the real and imaginary parts of the dispersion relations
of (i) the cellular automata model (solid line) ; (ii) the telegraphist equation
(dashed line) and (iii) Fick's law of diffusion (dotted line). Notice that some
curves may have two branches because of the ± sign in front of the square root.
r
equation 3.10. This is done in figure 4.3, where eiwr is given as a function
of kA. We also show the dispersion relation of the pure diffusion equation
eiw = e-D. (U.jl
We can observe that our cellular automata diffusion model is much closer
to the telegraphist equation than Fick's law. Of course, at large time, only
small k contributions will survive (i.e those for which eiwr is close to 1 )
and in this limit all three evolutions correspond. Note however that k ,....., n
also produces long-lived terms. As we previously mentioned, this is related
to the checkerboard invariant that plagues cellular automata dynamics.
4.3.3 The discrete Boltzmann equation in 2D
The same approach as described in the previous section can be applied
to a two-dimensional system. Of course, the algebra becomes more
complicated. Here, we would like to mention some results, without giving
a complete derivation. More details can be found in [1 1 2] .
In the 2D case, the discrete Boltzmann equation reads
r =
T N(r, t + ) (Q + 1 ) N(r, t)
where N is the column vector composed of the Ni, Q + 1 the four by four
( ) ( )
The matrix n + 1 can be diagonalized with the linear transformation
1 -1 1
1 � �
p = 2 1 -1 --11 -11 �
and p - 1 = 2 1 -1 -11 - 1
1 � _1 �
1 -1 1 1 -1 1 -1
so that
0 0
JJ
Po - P2 0
0 Po - P2
0 0
In terms of the new quantities
M = p -1 N (4.45 )
and, following the procedure of section 4.3.2, one obtains
D. 2 A 2 = A2 A
ro" t M1 + 2r o t M 1 D.VM 1 + A2
Po - P (-oA x2 + oA 2 ) M4 (4.46)
c. 4(P + P2 ) y
A2 Po 2+ P ( _ a 2 + ay2 ) M1
X
k,"A !l=(J.87
1!=0.93
J.L=0.99
(b)
,.------:--!
: (--'-
d) --=-----:::
:-:c :--1 1.5 ,-------:--!(1)-'---::-
-c -:
-:; -,
:-- 0.5
p=0. 1 8 D.=O.i p=0.3 D,=0.25 p=0.2 D,=l.O
Fig. 4.4. Contour lines of the diffusive part of the dispersion relation (J..L =
exp(iwr)) for various values of D. and p. The solid lines correspond to the CA
model, the dashed ones to the telegraphist equation and the dotted ones to Fick's
law of diffusion. For f.l = 0.99 the three dynamics are almost indistinguishable.
Note that the scales are not the same in all figures.
equations 4.46 and 4.47 can be Fourier transformed and the solution takes
the form
M1,4 - A 1,4 e i(wT+k·r) _
The geometry we want to study has the shape of a well whose vertical
walls are of infinite height and absorb particles. On the other hand, the
bottom line of the system acts as a source of particles. More precisely, we
shall impose the following conditions
=
p(x = O, y) = p(x L, y) = 0 for y � 0 (4.49)
p(x, y = 0) = cjJ = const for 0 x L < < (4.50)
and
ylim
-><XJ
p=0 (4.51)
The effect of a finite lattice spacing and finite time step are quite important
if L is not large enough. We shall also see that the choice of the Pi is
relevant even though the diffusion coefficient is not present in the Laplace
equation 4.48.
The exact solution of 4.48 with the above boundary conditions is
. ( )
known to be [113]
p=- 4 c/J " 1 nn y nn x 2 c/J _1 sin( y )
L.- - exp ( - - ) sm ( - ) = - tan (4.52)
n odd n
n
L L n sinh( i )
For a stationary state, the lattice Boltzmann dynamics 4.46 and 4.47
read
(4.53)
and
=
(-o" x2 + o" y2 )Ml + aV" 2 M4 f3 M4 (4.54)
--
where a and f3 are defined as
a = Po -+ pp f3 = 16pp(p++p P2 )
Po o
In order to solve 4.53 and 4.54 for the boundary conditions we defined
at the beginning of this section, we first assume that c 1 and c2 are,
respectively, parallel to the x- and y-axis. Next, we assume a solution of
the following form
M 1 ,4 = L A l ,4 (n) exp(z"y) sin(w"x) (4.55)
n
where
Q) n = nn (4.56)
L
z"
and has to be determined from 4.53 and 4.54.
4.3 Finite systems 161
This calculation turns out to be straightforward because the discrete
derivatives have simple eigenfunctions. Indeed, the functions exp(z"y) and
sin(cvn x) are eigenfunctions of the discrete operators a� and a�
a� sin(wn x) = �� sin(cvn x)
A b
ay2 exp(z"y) = 2z exp(z"y)
A
From the simple properties of the sin and exp functions, we find
bw = 2(cos( L n n A) (4.57)
- 1)
and
bz = exp(z"A) + exp( -z"A) - 2 (4.58)
After substitutions of 4.55 into 4.53 and 4.54, we obtain
(bw + bz )Al + cx (-bw + bz )A4 0 =
(-bw + bz )A l + [cx ( bw + bz ) - {J ] A4 0 =
{3 16 P(o -+PP2 )
p p
(4.60)
is a free parameter of the model. Its value plays a role when the quantity
A/ L does not tend to zero. Clearly, rx/ f3 can be made arbitrarily large by
choosing p small enough. On the other hand, its smallest value is
!X 1
f3 8
=
obtained for p 1/2 and Po P2 0.= =
=
It is then straightforward to calculate the density p 2Ml from 4.55
=
because, for y 0, our expansion is identical to 4.52. Therefore
p =
4</> L ! [exp(z"A)] (Y /"-l sin( nnx ) (4.61)
n n
n odd L
Strictly speaking, the density is only defined on the lattice sites but
equation 4.61 provides an interpolation of p everywhere. Its similarity
162 4 Diffusion phenomena
�=0.0
IX
�=0.1
IX
�=0.5
1 (a) L=20'A
1 n 20
Fig. 4.5. Comparison between the solution of the Laplace equation and the
solution of the CA dynamics for various values of rx/ p and for two system
sizes (a) L = 202 and (b) L = 1 002. The dashed line corresponds to the exact
solution exp -nn (in a log scale). This line turns out to coincide very well with
rx/P = - 1 /24 � -0.04 1 7.
with 4.52 makes it very appropriate for a comparison with the continuous
solution of Laplace equation. Since
y
.
. �
. . .
(a) ( b)
( c) (d )
Fig. 4.6. The diffusion front is the boundary between the source connected
cluster and the sink-connected cluster. The figures shows the four steps necessary
to extract the front : (a) the location of diffusive particles ; (b) the "land" cluster ;
(c) the "ocean" in gray ; and (d) the fractal front in black.
black site in its von Neumann neighborhood turns black, too. A similar
procedure is applied from the sink to determine the ocean.
Figure 4.6 illustrates these various steps, from the configuration of
diffusing particles to the construction of the front.
As we can see from figure 4.6, the diffusion front is a complex object.
Depending on the time at which it is obtained (i.e. the number of iterations
of the diffusion rule before rules (1) to (3) are applied), its width varies.
The longer the diffusion takes place, the wider the front is.
(J
Practically, the width is defined as follows. By symmetry, the front
properties are similar along the x-axis. Let us call P1(y, t) the average
166 4 Diffusion p henomena
density (over x ) of hull points at location y and time t. The average front
position YJ is defined by:
00
YJ ( t) = j0 yPPJj (y,(Y,tt)) ddyy
fow
and the width of the front a is given by
(J
2 - fow (y -00 YJ) 2 Pf (y, t) dy
'---"---
-- --:o=----'-----'c--.,---
-' -
j0 Pj (y, t ) dy
The relation between the width a and the time t determines the roughness
of the front. Typically, one expects that
oc
a ( t ) t'1 (4.62)
a
where is the roughness exponent.
In addition, the relation between the number
Nf ( t) = lo w Pf (y, t) dy
of particles in the front and its width a defines the fractal dimension DH
of the hull as
Nf = AaDH- l
where A is some amplitude. Here, we have subtracted -1 to DH because
we consider the dependence of Nf on a only. Clearly, Nf also depends
linearly on the front length Lx.
Roughness and fractal dimension have been measured according to the
above definitions and the results are given in figure 4.7. From the slope
of the log-log plot of Nf againts a and a against t, we obtain
a � 0.27
respectively. These values are in good agreement with those predicted by
independent theoretical arguments. The simulation which produced these
results has been obtained with the values p = 1/2 and Po = P2 = 0 in the
diffusion rule. Other values of the Pi yield the same fractal dimension and
roughness.
4.4.2 Diffusion-limited aggregation
Growth phenomena are frequent in many fields of science and technology.
A generic aspect of several growth processes is a far-from-equilibrium
formation situation and a fractal structure. Dendritic solidification in
an undercooled medium, electrodeposition of ions on an electrode and
viscous fingering are well-known examples of fractal growth [1 19].
4.4 App lications of the diffusion rule 167
Fractal dimension Roughness
Fig. 4.8. Two-dimensional cellular automata DLA cluster (black), obtained with
Ps = 1, an aggregation threshold of one particle and a density of diffusing particle
of 0.06 per lattice direction. The gray dots represent the diffusing particles not
yet aggregated.
1 1 .4
2 L......J..---l.-..l..-..._J...J.-..l-...l...-J 2 '---'----'--1---'---"'
0 4.2
ln(box-size) 0 ln(box-size) 4.2
(for instance four) and each time a diffusing particle reaches the site, the
counter is decremented by one. When it reaches 0, aggregation takes place.
Our approach has also some other differences compared with the origi
nal Witten and Sanders model. All particles move simultaneously and can
stick to different parts of the cluster and we do not launch them, one after
the other, from a region far away from the DLA cluster. For this reason,
we may expect some quantitative variation of the fractal dimension for
the clusters generated with our method. Figure 4.9 shows the result of a
box-counting algorithm in order to estimate the fractal dimension dt of
the DLA cluster obtained from our cellular automata rule. The principle
of this technique is to tile the cluster with a regular array of square boxes
of size t x t. One then counts the number Nb of these boxes which contain
a particle of the cluster. For a two-dimensional object, Nb is proportional
to the number of boxes, namely t-2 • For a fractal object, one expects that
Nb "' t-df
and, from figure 4.9 (a), one obtains
dt 1.78 (2-D) �
2.3 L..-.J...8-.l....-.l..-...L-..J.._...J-...J...J.-__j__J
3 9.8
ln(time)
Fig. 4. 1 1 . Formation rate o f cellular automata DLA clusters i n two and three
dimensions. The lattice has periodic boundary conditions.
where
oc � 2
in both two and three dimensions. Although these results are not sufficient
to conclude definitely that the 2-D and 3-D exponents are the same, an
explanation would be that in 3-D there is more surface to stick to than
in 2-D, but also more space to explore before diffusing particles can
aggregate. These two effects may just compensate.
A phenomena very similar to DLA is the so-called diffusion-limited
deposition [119] which corresponds to a deposition-aggregation process
on a surface. Our cellular automata model can be applied to model such
a process by having a nucleation surface instead of a single seed.
In the next section we consider a different deposition situation, in which
particles do not stick to each other but need an empty slot on the surface
to become adsorbed.
4.4.3 Diffu sion-limited surface adsorption
As a last example of the application of our cellular automata diffusion
rule, we show here how it can be used to model irreversible particle
adsorption on a surface. This problem is common in many areas of
science. For instance, adsorption of proteins and other particles at the
solid-liquid interface is a process of fundamental importance in nature.
The typical ingredients of irreversible adsorption are (i) the transport of
particles from the bulk to the surface and (ii) their subsequent adsorption
172 4 Diffusion phenomena
Fig. 4. 12. Excluded surface on the lattice, due to the adsorption of a particle.
We can imagine that the deposited particle is a diamond shaped object (dark
gray) which overlaps the surrounding light gray cells, making them unavailable
for further adsorption. In the diffusion plan, this picture is not the same, of
course, and the particles are point-like.
0.6 1.....-l--l--.J...._.J.._.l..<lilmiiO
1 1 66
Fig. 4. 14. (a) Kinetics of the cellular automata adsorption model in units of
iterations. (b) Available space as a function of the surface coverage.
Our analysis also shows that the jamming limit depends slightly on the
diffusion constant D and on the specific value of the Pi chosen for the
diffusion rule. This may be an indication that, for this non-Markovian
process, it is not sufficient to give D to uniquely specify the macroscopic
behavior.
Finally, we show in figure 4. 14 the kinetics of the diffusion-limited
adsorption process and the relation between the available space (the white
cells in figure 4. 13) and the coverage (the black cells). From the first plot,
we can conclude that the surface coverage obeys an exponential law
eCfJ - 8( t) exp( - tI Tc )
"'
where Tc is some characteristic time. The second plot describes how the
available space 4> decreases with the coverage fraction e. Clearly, for
e � �
0, we have ¢(8) - 5 8 since each adsorption reduces the number of
deposition sites by five cells. On the other hand, near the jamming limit
� �
8 800, only isolated empty sites remain and ¢(8) -8. In between, 4> is
a non-trivial function of e .
4.5 Problems
5.1 Introduction
t=5 t=llO
t=l l5 t=120
2d
and
_- .A. 1 1 eirx arx PA
A (i 1)
2d V
where V is the eigenvalue of the diffusion matrix n for the eigenvector
E rx = ( e 1 ,rx ; ez,rx ; ... ; e2d,x )
The equation for the density PA is now obtained by summing over i
equation 5.5, remembering that
Collecting all the terms up to O(c2 ), we see that the orders O(c0 ) and O(c)
5.4 Anomalous kinetics 187
vanish and we are left with
c2 'r2 8 t � A �0) + c2 A t 2.:)ei
l l
· V )A � t ) + c 2 � �(ei V)2A�o)
t
l
·
j
Using the definition of 'r2 , A t , R�j and performing the summations yield
8 t PA = DV2PA + �1 '"' PA PB P ca
� Rj ( ld ' ld ' ld ) (5. 13)
1
between the two effects. Actually, both initial condition and many-body
correlations may simultaneously affect the system. The slowest decay law
will be longer lived and mask the other behavior. A cd/4 will certainly
screen out a t-d/2 decay.
5.4.2 Cellular automata or lattice Boltzmann modeling
The A + B 0 reaction-diffusion process is therefore a good candidate for
�
(a) (b)
Fig. 5.4. Segregation in the A + B - 0 reaction-diffusion process, as obtained
from (a) the cellular automata model and (b) the lattice Boltzmann dynamics. The
situation corresponds to the evolution of a noncorrelated random, homogeneous
initial condition. In (a) the white dots represent lattice sites with a majority of A
particles and the gray ones a majority of B particles. The black dots correspond
to empty sites. In (b) the gray levels indicate the regions where A (or B) is
dominant : black when a(r, t)jb(r, t) > 1, dark gray when a(r, t)/b(r, t) > 6, light
gray when b(r, t)/ a(r, t) > 1, and white when b(r, t)/ a(r, t) > 6.
"
[]
D
o D
slope�-0.9 3
3 L-�---L--�--��-��
0 12
In(time)
Fig. 5.5. Time behavior of the total number of A particles in the A + B ---+ 0
process, in one dimension. The two plots correspond to the lattice Boltzmann
(dots) and the cellular automata (squares) dynamics, both for a correlated initial
condition. The lattice Boltzmann takes into account the fluctuations of the initial
state but not the fact that the dynamics itself is not mean-field. The observed
time behavior is t'1, with rx � 1 /2 for the cellular automata and rx � 0.9 for the
lattice Boltzmann dynamics.
with respect to the other, the predictions found by Lindenberg et al. [142]
(namely a c(d+2l/4 long time regime and a cd/ 4 short time regime)
are verified in the lattice Boltzmann dynamics. The cellular automata
dynamics, on the other hand, does not agree well with these theoretical
predictions which do not take the fluctuations fully into account.
5.5 Reaction front in the A + B � f/J process
Fig. 5.6. Reaction front in the A + B � (/) reaction diffusion process (according
to cellular automata dynamics).
difference Ao - Bo). Second, their scaling theory predicts that1 front width
W(t) follows, in the long time limit, the power law W ( t) t 16 ."'
Whereas the Jt law giving the front position holds in all generality,
the question of the width is more involved. It was shown [147], using
cellular automata models, that the upper critical dimension of the problem
is 2. That is, logarithmic corrections to the rate equation predictions
(i.e. the W t 1 16 behavior) are found for the space dimension d
"' =
=
2 and an important correction is measured in d 1. This reaction
diffusion process has now been intensively studied due to its departure
from the mean-field description. The reason for this departure is of
course that the average reaction term (R) is not the product of the A
density times the B density. Instead, it is the joint probability of having
one A and one B particle simultaneously present at the same position.
5.5 Reaction front in the A + B � 0 process 195
Unfortunately, this joint probability is not known in general. An equation
for it could be written down but it would involve other higher-order
correlation functions.
Thus the problem of finding the right behavior for W(t) in low
dimensional systems is clearly a very difficult analytical task. Of course
the reader may argue that, since the upper critical dimension is d 2, =
everything will be fine in the 3d physical space and the mean-field result
is enough. However, the problem is very interesting from a theoretical
point of view and sheds some light on what may happen even in very
simple nonequilibrium systems. In addition, some experiment have been
realized in very thin tubes [148] but they are not thin enough to probe
the one-dimensional regime.
5.5.1 The scaling solution
The cellular automata approach is a very successful tool in the analysis of
reaction diffusion fronts. Moreover, a scaling theory can be considered in
order to interpret, in a theoretical framework, the simulation results.
The idea is to write macroscopic rate equations (with dependence only
on one spatial coordinate for symmetry reasons) as :
(5.2 1 )
but without specifying an explicit form for the reaction term R. Here, as
opposed to the homogeneous A + B 0 case, the spatial fluctuations of
�
the initial configuration does not play a crucial role since we assume that,
= = =
at time t 0, ai 1, bi 0 on the left part of the system and conversely
on the right part. Thus, we may argue that the departure from mean-field
behavior is due to the fact that
R =f. Rmr = ab
where Rmr denotes the standard mean-field approximation. In general, R
is the joint probability of having one particles of type A and one particles
of type B simultaneously present at the same point. Of course, since R
is not known, we cannot solve equation 5.21. However, in the spirit of
a scaling approach, we may say something about the asymptotic time
dependency of a(x, t) and b(x, t). Thus, we shall assume that
a(x, t) t-Ya(xt - rx )
b(x, t) t-Ya(xcrx )
R(x, t) cP R(xt-rx ) (5.22)
cx,
where fJ and y are the exponents we are looking for. From these
exponents, we can obtained the behavior of the width front W(t) defined
196 5 Reaction-diffusion processes
as
W 2 (t)2:�:-�Llx ;- (x) )2 R(x, t) (5.23)
Lx;:-L R(x, t)
=
3.43
L72
o m "'-""""-��-'---''---'-_L_-'---"'---'--'-_L___,---'-_,__,__:]1__"---cw�
-'-
-257 -L50 -0.42 0_65 L72 2.79
x/t o-292
Fig. 5. 7. Scaling profile of the reaction term, as obtained with a data collapse
from the cellular automata simulations.
In this section we shall study in detail a more complex system in which sev
eral of the ingredients we have introduced so far are combined: reaction
diffusion will be accompanied by solidification and growth phenomena.
This gives rise to nice and complex structures that can be naturally mod
eled and analyzed in the framework of the cellular automata approach.
These structures are known as Liesegang patterns, from the german
chemist R.E. Liesegang who first discovered them at the end of the
nineteenth century [152] .
5.6. 1 What are Liesegang patterns
Liesegang patterns are produced by precipitation and aggregation in the
wake of a moving reaction front. Typically, they are observed in a test
tube containing a gel in which a chemical species B (for example AgN03)
reacts with another species A (for example HCl). At the beginning of
the experiment, B is uniformly distributed in the gel with concentration
b0. The other species A, with concentration a0 is allowed to diffuse into
the tube from its open extremity. Provided that the concentration a0 is
larger than bo , a reaction front propagates in the tube. As this A + B
reaction goes on, formation of consecutive bands of precipitate (AgCl in
our example) is observed in the tube, as shown in figure 5.8. Although
this figure is from a computer simulation, it is very close to the picture of
a real experiment.
The presence of bands is clearly related to the geometry of the system.
Other geometries lead to the formation of rings or spirals. The forma
tion mechanism producing Liesegang patterns seems quite general. It is
assumed to be responsible for the rings present in agate rocks [153, 154]
to yield the formation of rings in biological systems such as a colony of
bacterium [1 52] or to play a role the the process of cement drying.
Although the Liesegang patterns have been known for almost one
hundred years, they are still subject to many investigations. The reason
is that they are not yet fully understood and can present unexpected
structures depending on the experimental situation (gravity, shape of the
container, etc.).
On the other hand, it has been observed that, for many different
substances, generic formation laws can be identified. For instance, after a
5.6 Liesegang patterns 199
transient time, these bands appear at some positions xi and times ti and
have a width wi. It is first observed that the center position Xn of the nth
band is related to the time tn of its formation through the so-called time
law Xn Jtn.
"'"'
300
Xn 200
1 00
0 L_--�--�---L--�
0 1 00 200 300
Xn-1
Fig. 5.9. Verification of the spacing law for the situation with C particles. The
ratio Xn / Xn-1 tends to p = 1 .08.
No precipitation
I ,
� BANDS
\- - - - - - - .
Amorphous '. Dendrites'
solidification 1
I
:
I
Fig. 5. 10. Qualitative phase diagram showing the different possible patterns that
can be obtained with our cellular automata model, as a function of the values of
ksp and kp .
(a)
(b)
(c)
Fig. 5. 1 1 . Examples of patterns that are described in the phase diagram : (a)
corresponds to homogeneous clustering ; this is also the case of pattern (b) but
closer to the region of band formation. Pattern (c) shows an example of what
we called a dendrite structure. Amorphous solidification would correspond to a
completely uniform picture.
=
but with an aggregation threshold kp 0. Figure 5.12 shows the results
of such a modeling. The fractal dimension of these clusters is found to
be around 1.77, a value which is very close to that measured in a real
sample [163].
The patterns we have presented so far show axial symmetry, reflecting
the properties of the experimental setup. But the same simulations can be
repeated with different initial conditions. A case of interest is the situation
of radial symmetry responsible for the formation of rings or spirals. The
reactant A is injected in the central region of a two-dimensional gel initially
filled with B particles. The result of the cellular automata simulation is
shown in figure 5.13. In (a) concentric rings are formed, starting from
the middle of the system and appearing as the reaction front radially
moves away. In (b) a spiral-shaped structure is created. Although the two
situations are similar as far as the simulation parameter are concerned,
the appearance of a spiral stems from a spontaneous spatial fluctuation
which breaks the radial symmetry.
5.6 Liesegang patterns 205
7.0
"'
,.-_
"'
«l
s (b)
'-"
.s
3.2
0.65 In (size) 2.8
-0.4
�
u
§
0..
::s
u
(c)
u
0
'-"
.s
-4.2
(a) 0.65 In (size) 2.8
(a) (b)
Fig. 5 . 1 3 . Liesegang rings (a) and spiral (b), as obtained after 2000 iterations of
the cellular automata model with C particles (indicated in gray).
206 5 Reaction-diffusion processes
1 .9 ,-�-----,--,----�-�-,--r-....,.--;,---,
1.8
1 .7
i 1 .6
.s 1.5 ks/a0=0.0081 •
1 .4 ks/a0=0.0088
ks/a0=0.0089
1 .3 ks/a0=0.0090 x
ks/a0=0.009 1
•
1 .2 ksja0=0.0092
1.1 '----�-_.._-'-----'--�---''---'-_j
2 2. 1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3
log(xn )
-+ p
IX
(5.37)
where ()( is an exponent which depends on the initial concentration ao
and b0 . We find, for several simulations, that ()( is typically in the range
208 5 Reaction-diffusion processes
(a) (b)
Fig. 5. 1 6. Formation of (a) Liesegang rings and (b) spiral-shaped pattern, as
obtained after 2000 iterations of the lattice Boltzmann model with C particles.
0.5-0.6 (see figure 5.15). Note that, in this model, aggregation of several D
particles on top of each other is possible. If this superposition mechanism
is not permitted, it is found that the bands are wider and an exponent
rx � 1 is then obtained, in agreement with experimental observations [164].
5.6.5 Lattice Boltzmann rings and spirals
We can consider again the case of Liesegang rings and spirals, but now in
the framework of the lattice Boltzmann model.
For instance, figure 5.16(a) shows the situation where concentric rings
=
of precipitate are formed. The numerical parameters are : ao 1, bo / ao =
= = p =
0.013, Db/ Da 0.1, ksvl ao 0.0087, k / ao 0.0065. The nucleation
process takes place with a probability of 0.05 and aggregation with a
probability close to 1. This pattern turns out to be quite similar to real
Liesegang structures obtained in a similar experimental situation [152].
=
For the same set of parameters, but bo / ao 0.016, a different pattern is
observed in figure 5.16(b). There, a local defect produced by a fluctuation
develops and a spiral of precipitate appears instead of a set of rings. Such
a spiral pattern will never be obtained from a deterministic model without
a stochastic component.
From these data, we can check the validity of the spacing law for ring
formation. The relation
r /r - 1 � P
n n
holds, where r is the radius of the nth ring. In figure 5.17, the Jablczynski
n
.�
,...,
1 .2
(.)
s
<!.)
0
(.)
:Q
rfl
N l. l
�
(.)
�
......
1 .0 ---�--- - � -�-
Initial concentration o f B
=
bo (for ao 1) both for axial (bands) and radial (rings) symmetries. We
notice that p decreases when bo increases in agreement with experimental
data. Moreover, for the same set of parameters, the value of p is found to
be larger in the case of rings than it is for bands.
The vectors e 1 = - h e2 = -e4 are the four unit vectors along the main
directions of the lattice. The stochastic Boolean variable P it (r, t) is 1 with
probability Pi and selects whether or not particle t chooses to move to site
r + Jc e i. Since each particle has only one choice, we must have
Pot + Ptt + P2t + P 3t + P4t = 1
The macroscopic occupation number N ( r, t) = (n(r, t) ) is obtained by
averaging the above evolution rule over an ensemble of equivalent systems.
Clearly, one has
n(r,t)
=
( t=l
2::: Pit (r, t) ) PiN ( r, t)
The quantities IXz, /3j are the stoichiometric coefficients, and k is the reaction
constant.
In order to model this reaction scheme with a multiparticle dynamics,
one considers all the q-tuples that can be formed with IX l particles of A 1 ,
1X2 particles of A 2 , etc. These q-tuples are transformed into m-tuples of Bj
( � ) ( � ) (::q )
particles with probability k. At site r and time t, there are
n n 2
JV ( r, t) = 1 • • • ( r, t)
iX
ways to form these q-tuples, where nx(r, t) denotes the number of particles
of species X present at (r, t). If one of the nA; < IXi then obviously JV = 0.
This techniques offers a natural way to consider all possible reaction
scenarios. For instance, in the case of the annihilation reaction 2A 0, �
k=0. 8
6 �������
0 ln(time) 10
Fig. 5. 1 8. Time decay of NA, the total number of A particles in the A + A ---+ f/J
reaction-diffusion process, with the multiparticle method. A non-mean-field power
law t-d/ 2 is observed.
When k is very small, we may assume that all the % q-tuples need to
be considered and the above reaction rule can be expressed as
%(r, t)
nA/r, t + T) = nAJr, t) - CX[ L Kh
h=l
%(r, t )
nBir, t + T) = nBj (r, t) + /3j L Kh (5.42)
h=l
where Kh is 1 with probability k.
This algorithm may become quite slow in terms of computer time if
the nx are large and k � 1. In this case, the Gaussian approximation
described in the previous section can be used to speed up the numerical
simulations: the number of accepted reactions can be computed from a
local Gaussian distribution of mean k..!V(r, t) and variance k(1 - k)JV(r, t).
In order to check that our multiparticle reaction rule captures the
true nature of fluctuation and correlation, we simulated the A + A 0 -+
The first important result is that, if one assumes that each configuration
with a total of M particles over L sites appears with the same probability,
( )) ( ) ,....,
the quantities (n:,l ) can be computed
/ nA 1 ___!___ MA 1
__!__ (5.43)
\ rxz =
Lrx1 rxz rxz ! Ni1
where NA1 is the average particle number per site of species Az and the
number of lattice sites L is supposed to be large enough.
This result can be derived from the following combinatory arguments :
let us suppose that a number M of particles is homogeneously distributed
among L sites of the lattice. The average number N of particles on these
=
sites is then N MIL .
The probability for a particle to pick a given site r to stand on is 11 L
(and 1 - 1 I L to choose any other site). Thus the probability that a given
() ()
site contains n particles is
P (n) - _
M -1 n
1--
( )
1 M-n
-
M (L l) M - n
_
_
n L L n LM
The average we want to compute is
1 M M! (L - l) M -n
rx! �
(M - n) !(n - rx) ! LM
1 M' (M' + rx) ! (L - 1) M' -n'
rx! E(M' - n') !n' ! LM' +rx
5. 7 Multiparticle models 217
( + a)
M' L
-rx M' M' ! (L - 1) M' -n'
(M'a+ a)
11. E (M' - n') !n' ! L M'
-
L rx (1 + 1) M' �� (5.44)
where M' = M - a and n' = n - a.Thus
1:
In the limit � 0, we obtain the usual form of the rate equations for the
reaction process considered, namely
a t NA- ( t ) - k ' Nrx! Nrx2 Nrxq
'f:
I
- -- A 1 A2 · · · Aq
k Nrx2 N rxq
- -1:' Nrx1
a t NB,. ( t ) - A 1 A2 · · · Aq
imposed macroscopic length scales (like the size of the container). Turing
patterns exhibit regular structure with an intrinsic wavelength depending
on the diffusion constants and reaction rates. Typical examples of inho
mogeneous stationary states observed in experiments have a hexagonal or
a striped structure [168].
For the sake of simplicity, we consider here only one of the simplest
models showing Turing patterns : the Schnackenberg reaction-diffusion
model [169] in two dimensions. It describes the following autocatalytic
reaction :
A�X
2X + Y � 3X (5.45)
where the densities of the species A and B are kept fixed (for instance
by external feeding of the system). This situation of having a fixed
concentration of some chemical is quite common in reaction-diffusion
processes. As a result, there is no need to include all the dynamics of
such reagents in cellular automata or multiparticle models. It is usually
enough to create randomly a local population of these particles at each
lattice sites.
Here we consider a two-dimensional multispecies multiparticle lattice
gas model with alternating reaction and diffusion steps, as explained in
sections 5.7.1 and 5.7.3, namely
k
• Reaction step : for the reaction nS1 +mS2 � qS3 all possible families
m
composed of n S1 and S2 particles have a probability kj to create
q particles s3 .
• Diffusion step : the particles jump to a nearest-neighbor site with
an equal probability for each direction. This gives a bare diffusion
coefficient of ..12 /4r. A simple way to tune it without having a
probability for a particle to be at rest is to vary the number t
of consecutive diffusion steps for a given species : this amounts to
=
introducing a different time step Trn rjt for this species and yields
D = tA-2 j4r.
The instability of the homogeneous state leading to Turing struc
tures can be understood using the corresponding macroscopic rate equa
tions [130] for the local average densities x and y
8t x k 1 a - k2 x + k 3 x2 y + Dx V2 x
8ty = k4 b - k3x2 y + Dy V2 y ( 5.46)
where a and b represent the densities of particles A and B, respectively.
A conventional analysis shows that for some values of the parameters a
5.8 From cellular automata to field theory 219
• • •!.
(a) (b)
vary continuously in space and time. They describe the system at the
macroscopic scale, while keeping track of all the fluctuations related to
the disregarded microscopic degrees of freedom. Such a description is
called a continuous field theory. These field theories can then be studied
in their own right, using standard methods such as perturbation techniques
or renormalization group.
The first step to perform this program is to notice that the probabilistic
cellular automaton rules used for reaction-diffusion can be expressed as a
master equation (in discrete time) for generalized birth-death processes.
The way to associate a continuous field to a master equation for
a birth-death process (in continuous time) has been studied by many
people [172-174].
For the sake of simplicity, and to avoid as much as possible technical
difficulties, we shall start with a very simple one-cell problem (zero
dimensional system). Moreover, we shall assume that only one type of
particle is present. The state of the cell is characterized by </J n (t), the
probability of having n particles at time t. We consider a multiparticle
model, with no exclusion principle, so that any value of n is allowed.
One considers a dynamical process described by the following master
equation
=
8 t </J n (t) L [w(n' � n )</J n' (t) - w(n � n' )</J n (t)] (5.47)
'
n =/= n
where w(n' � n) is the transition rate from the state with n' particles to
the state with n particles. One recognizes the degenerate form (only one
site) of a probabilistic cellular automaton rule. This stochastic process is
Markovian: there is no memory effect.
A natural way, in statistical physics, to deal with a problem in which
the particle number is not conserved is to introduce a so-called Fock space
structure. One works in a phase space which is the direct sum of all the
phase spaces corresponding to an arbitrary number of particles. Creation
and destruction of particles is described by creation and annihilation
operators acting on this Fock space. To each state { <P } one can associate
n
a l z) = z * l z) (5.61)
In this approach the master equation takes the form
8 t i<I> ) = 2? I <D) (5.62)
the Liouvillian 2? is a polynomial in the creation and annihilation
operators and can always be written in normal form ( all the creation
operators are on the left of the annihilation operators)
(5.63)
The exact form of 2? i.e. the values that the constants CiJ take on, will
depend on the form of the transition rates w(n' � n) for the model under
consideration. Thus one can take either the w or the Liouvillian 2? to
define the model.
Before developing further this general formalism, let us consider a very
simple example, namely the simple decay A � 0 with a reaction rate w.
In this simple case, the different steps can be made explicitly. The master
equation is
D t cP n (t) -wnc/J n (t) + w(n + 1)¢ n+l (t)
= (5.64)
From equation 5.64 it follows that the generating function G(z, t) obeys
the equation
O t G(z, t) = w(l - z)oz G(z, t) (5.65)
whose solution is
G(z, t) = G((z - l) exp(-wt)) (5.66)
Thus the factorial moment of order one is
n 1 (¢) = (n) (t) =
a
= G(z, t) l z =l (n) (O) exp (-wt) (5.67)
oz
as expected from physical grounds. We can now construct the Liouvillian
2·'? . We have
O t l <l>)
= - L wnc/J n l n) + L w(n + 1)¢ n+l ln) -wnal<l>) + wa l <l>) (5.68)
=
n n
and the Liouvillan takes the simple form
2? = w(a - n a) (5.69)
Let us now return to the general case in zero dimension. To obtain a field
continuous field theory, one constructs from 5.62 and 5.63 a path-integral
5.8 From cellular automata to field theory 223
representation for
P(z, tlzo, 0) = (zl exp t2� lzo ) (5.70)
by inserting 5.59 into 5.70 (N - 1) times, in the usual way [175]
N- 1 dz d k -
(
P(z, tlzo, 0) = IT j ; � e zkzZ ) (zN I exp c-2� 1 zN- 1 )
•
k= 1 m
where the dot signifies differentiation with respect to time. One recovers
an object familiar in field theory. The long time properties are extracted
by studying the action
S(t, O) = lot dt' [zz·· - I.= Cij z i (z * )i ] (5.74)
ij
(5 .87 )
where we have used 5.60. Now using 5.58 and 5.61 one has
a t G(z, t) L. . Cij Z i ( j_
=
)i (z J <D)
lj
az
!l'C§ G(z, t) ( 5 .88 )
where
(5 89 )
.
One sees the close relationship between !l':F and !l'C§ . To each creation
n
operator present in !l'g; one associates a factor z in !l'C§ and to each
226 5 Reaction-diffusion processes
annihilation operator a in seJF one associates the operator a I az in !l?cg .
It is now possible to proceed from 5.85 as described above to find the
evolution equation for f(a, t). One finds
atf(a, t) = se&f(a, t) (5.90)
where
!l? & = �
a . .
"""' Cij(1 - -)1a1 (5.91)
lj
.. aa
is the Liouvillian describing the evolution of the quasi-probability density
f(a, t). This Fokker-Planck-like equation is exact, no approximation or
truncation has been made in deriving it from the master equation 5.47. It
can be used as the basis for the study of processes described by 5.47, or as
the starting point for deriving a path-integral representation in terms of
the a variables. To carry out this latter program, one introduces operators
& and p [181] satisfying
[a,A pA ] _ = z [&, &]_ = 0 [p, p]_ = 0
•
(5.92)
with
p = -z
A
a•
(5.93)
aa
in the a-representation. Then just as the master equation 5.62 is the Pock
space analogue of 5.85, one can write an operator analogue of 5.91
(5.94)
where
se m = l: Cj(1 + ip) i ( &)j (5.95)
ij
Introducing Ia) and lp), which are eigenkets of & and p with eigenvalues
a and p, respectively, one can derive a path-integral representation for
(5.96)
in exactly the same way as described in 5.70. One finds the analogous
expression to 5. 73 to be
j d[p, a] exp -{lot dt[ipa - I,: Cij(1 + ip) iaj] } (5.97)
lj
The "action" in the path-integral 5.73 and that in 5.97 are identical as
long as the identifications
a � z * , ip � (z - 1 ) (5.98)
5.8 From cellular automata to field theory 227
are made. Although it was not obvious a priori that the two actions would
be identical, if 5.98 is expressed in the form
a a
a +-+ ' ( z - 1) (5.99)
az aa +-+
it becomes very much more plausible, given the conjugate nature of the
variables a and (z - 1) indicated in 5.84.
To show the equivalence of the two formalisms in the general d
dimensional case, one can proceed in exactly the same way as for the
one-site problem. First of all, an analogous procedure to that carried out
in 5.87�5.89 shows that the Liouvillian describing the evolution of G(�, t)
is !f!q; = !f!(z 1 , 0�1 , , Zj, 0�j , ) and hence that governing the evolution of
••• •••
f(q,_, t) is !J!fl' = !1!( 1 - 0�1 , a 1 , ··· ' 1 - �j , aj , ... ). Secondly, the operator ana
a
logue of !J!fl' is immediately seen to be !!! (!) = !1!(1 + ip 1 , & 1 , ... , 1 + ipJ , &1, . ).
. .
Finally, path-integral representations for the solution of the Fokker�
Planck equations can be constructed from !f!:F and !!!(!) and shown to be
the same, if the identification
(5. 100)
is made.
Instead of considering the generalized Fokker�Planck equation obtained
in the Poisson formalism, one can consider the associated generalized
Langevin equation. There is a one to one correspondence between the
two [179]. Let us consider again the simple case A + A 0 [182]. The �
( )
generalized Fokker�Planck equation is
N a2 a
a d ({ a(n) } n , t) = - k ; 2
aa2 (m)
- 2 aa(m) (a (m)f({a(n) } n , t))
N a
-d ; aa(m) ((a(m - 1) - 2a(m) + a(m + 1))f({a(n)} n , t)) (5. 101)
(D
The corresponding generalized Langevin equation is
2 )
da(r, t) = : 2 a(r, t) - 2ka(r, t) 2 dt + J-2k a(r, t) 2 dW(r, t) (5. 102)
r
This equation should be considered as a pair of coupled stochastic equa
tions for the real variables aR(r, t) and a1 (r, t), respectively the real and
imaginary parts of a(r , t)
drx1 (r, t) (: 2
)
D 2 rx1 (r, t) - 4krxR(r, t)rxJ (r, t) dt
r
+ JikrxR(r, t)d W(r, t) (5.104)
As one sees from the above equations, the noise terms arising from the
microscopic fluctuations are multiplicative with non-trivial correlation.
In particular, the noise-noise correlation is negative. This is an unusual
feature related to the nonequilibrium aspect of such reaction-diffusion
problems.
A Boltzmann-like approximation will overlook such complications and
leads to erroneous results. These problems show clearly the advantage
of working with cellular automata algorithms able to model important
fluctuations in reaction diffusion problems.
5.9 Problems
5. 1 . Toffo l i and M argo l us [26] p ropose a cel l u l a r auto m ato n (tu be
worms ru l e) to model the Belousov-Zhaboti nsky react i o n . The state of
each s i te is either 0 ( refracto ry) or 1 (exc ited) and a l oca l t i m e r (wh ose
val ue is 3, 2, 1 or 0) contro ls the refracto ry period. Each iteration of
the r u l e can be exp ressed by the fo l l owi n g seq uence of operations: ( i )
where t h e t i m e r i s zero, t h e state i s exci ted ; ( i i ) t h e t i m e r i s decreased
by 1 u n l ess it is 0; ( i i i ) a s i te beco mes refracto ry whe n ever the t i m e r
is eq ual t o 2; ( i v) the t i m e r is reset t o 3 fo r t h e exc ited s i tes w h i c h
h ave two , o r m o re than fo u r, exc ited s i tes i n the i r Moo re n e i g h borhood .
S i m u l ate t h i s auto m ato n , start i n g from a random i n itial conf i g u ration
of the t i m e rs and the exci ted states. Observe the fo rmation of s p i ra l
pai rs o f exc i tations. C heck that th i s ru l e is ve ry sensitive t o smal l
mod ifications ( i n particu l a r to the o rder of operations ( i ) to ( i v) ) . What
are the d iffe rences betwee n th is ru l e and the G reenberg-Hast i n g m od e l
d i scussed i n section 5.2?
5.2. Mod ify the G reenberg-Hasti n g ru l e (or the Toffol i and M a rgol us
tu be-wo rms ru l e of the p revious exce rc ise) in order to m od e l a one- or
th ree-d i m e n s i o n a l exc i tab l e m ed i a .
5 . 3 . Contag i o n i n e p i d e m i c mod e l : a s i m i l a r dynam i cs t o that used for
d escri b i n g an exc i ta b l e c h e m i cal med i a can be con s i d e red to model
contag i o n in an e p i d e m i c d i sease [56] . Each s i te is occ u p i ed by an
i nd iv i d u a l who can be either suceptib/e, infectious or immune. A
healthy, s ucepti b l e i nd iv i d u a l i s i nfected o n l y if there is at l east one
5.9 Problems 229
i nfectious person i n i ts von N e u m a n n n e i g h borhood . An i nfectious i nd i
v i d u a l rem a i ns so d u ri n g m steps a n d t h e n beco mes i m m u ne. I m m u n ity
stays fo r m steps: afte r that the i m m u n e i nd i v i d u a l s beco m e s ucepti b l e
agai n . Co n s i d e r a CA s i m u l ation o f th i s contag i o n p rocess , start i n g
w i t h a random conf i g u ration o f s ucepti b l e , i nfectious o r i m m u n e i n d i
v i d u a l s . What is the d i ffe rence betwee n th is model and the G reenberg
H asti n g ru l e? Modify the e p i d e m i c ru l e to m ake a p robab i l istic CA.
What is then the d ifference from the forest fi re mod e l s d iscussed i n
chapte r 2?
5.4. P ropose a CA model of a prey-p red ato r syste m where two spec i es
( p reys and p red ato rs) d iffuse on a 20 l attice and i nte ract accord i ng to
the fo l l ow i ng ru l es : n ew p reys come to l ife at a rate proportional to
the cu rrent n u m be r of p rey (e . g . one parent p rod uces one offspri ng) .
A p rey is eate n w i th probab i l ity k when it encou nte rs a predator.
P red ato rs die with a g i ve n constant p robab i l ity and co m e to l ife when
a parent predato r eats a p rey. Ass u m i n g that p reys and p redato rs are
homogeneously d i stri b uted , show that th i s syste m is ro u g h l y descri bed
by the mean-f i e l d equations p = ap - kpP and P = -bP + kpP where
p stands fo r the total n u m ber of preys and P the tota l n u m be r of
p redato rs . The q u antities a, b and k are paramete rs of the mode l .
T h i s syste m o f equations is known a s t h e Lotka-Vo lterra model and
yields an osc i l l atory behavior. S i m u l ate the auto m ato n mode l . Are
the speci es homogeneously d i stri buted in space o r do you observe
an osc i l l ato ry space-t i m e patte rn? Mod ify the r u l e as fo l l ows ( i ) eac h
p redato r has an i nte rnal " starv i n g c l ock" w h i c h is d ec reased by 1 at
each t i m e step. When a p rey is eate n , the c l ock is reset to i ts maxi m u m
val ue; when the c l ock ru ns down to zero , the p redator d i es . ( i i ) Add a
" p regnancy c l ock" w h i c h determ i n es the ti m e n ecessary for a p redato r
t o g i ve b i rth t o a baby. H o w do these m od if i cati o n affect t h e g l obal
behav i o r? I m ag i n e othe r mod ifications that are sound from a b i o l og ical
po i nt of v i ew. Can you eas i ly write down the correspond i ng d iffe renti a l
eq u ation?
6.1 Introduction
Tc)/ Tc. This nonanalyticity usually takes the form of a power law:
(6.2)
and y is called the critical exponent for the variable Y . Several critical
exponents can be defined. In the equilibrium case, hypothesis concerning
the homogeneity of the thermodynamic potentials leads to relations among
these exponents, called scaling relations.
To illustrate the problem, let us consider again a liquid-gas system.
At a given critical density, pressure and temperature, the system changes
continuously (without latent heat) from the gaseous to the liquid phase.
Two quantities that can be experimentally determined are the isothermal
compressibility and the specific heat at fixed density. It is found that
both quantities diverge (with different critical exponents) as the critical
point is approached. The universal character of the phenomenon is
shown by the fact that the values of these exponents are the same for
different liquids. More strikingly, these exponents turn out to have the
same values for an order-disorder transition in a binary alloy. This
universal behavior is nowadays well understood for equilibrium phase
transitions. The so-called renormalization group theory [171,186] gives
a clear explanation and classifies the different systems into universality
classes. The different universality classes are characterized only by a small
number of parameters, namely, for systems with short-range interactions,
234 6 Nonequilibrium phase transitions
the dimensionality of the system d and the number of components n of
the order parameter describing the change of symmetry at the transition.
The situation is not so clear in the framework of nonequilibrium systems.
Nonequilibrium statistical mechanics is still in its infancy and there is no
general theory available to describe such systems. Most of the known
results are based on numerical simulations. However, here again the
concept of universality classes appears to be relevant although we do not
completely understand how the universality classes are characterized.
Accordingly, we shall discuss some simple examples to illustrate the
basis ideas and possible approaches to study such nonequilibrium systems.
First we shall sketch the theoretical approaches and their difficulties. Then
we shall discuss the modeling in terms of cellular automata and review
the problems and difficulties associated with this approach.
6.2 Simple interacting particle systems
take into account the correlations between cells. For example, one can
write the evolution equation for the joint probabilities Pij that two nearest
neighbor sites are respectively in the states i, j. In one dimension, these
236 6 Nonequilibrium phase transitions
evolution equations are
PAA (t + 1) = [
PAA (t)D(t) P�A (t) + � P6A (t) + PPOA (t)PAA (t)]
+P2 PoA (t) + P2 Poo(t) (6.6)
where D(t) (PAA (t) + iPBA (t)). By symmetry, POA PAO · Moreover, the
= =
1 .00
0.80
0.60
><"'
0.40
0.20
Fig. 6. 1 . Phase diagram for the one-dimensional A model. The full line
correspond to the one-site mean-field approximation and the dotted line to the
results of numerical simulations of the CA rules described later.
where the sum is over all the configurations and p is the probability dis
tribution. Physical states 1'1' ) must satisfy the positivity and normalization
conditions :
( {h , h , .. . , }k , ... } 1'1' ) � 0, V{j1 , h , . .. , A, ... } (6. 1 8 )
and
( ·1 '1' ) 1= (6.19)
where
=
(I = IJ ( (<I>o,A I + (<I>1,A I ) L( {h , h , ... , )k, . .. } l (6.20)
{ik} {ik}
S 0 are the states l i t , i2 , ... , im) for which all the sites i t , i2 , ... , im are in the
excited state llfJ u ), while all other sites are in their ground state l lfJo,i ) . The
corresponding eigenvalues are Am - m(1 + p) with m 1, 2, ... .
= =
Let Pm be the projection onto the subspace of states with exactly m
excited sites. Then:
(jJ (jJ
[S 0r t L Am - t Pm -(1 - v) L m- t Pm
= = (6.3 1)
m=t m=t
It is necessary to introduce new creation and annihilation operators bj
and bi defined as
bJ llfJo,i) llfJt ,i) (6.32)
hi llfJt,i) 0 (6.33)
bil lfJ u ) llfJo,i) (6.34)
bi l tpo, i) 0 (6.35)
They are related to the creation and annihilation operators a i , ai as follows :
a i b i (1 + (1 - v)bi ) - (1 - v)bj((1 - v) + bi)
= (6.36)
240 6 Nonequilibrium phase transitions
a; = b i (vb j - 1) - vbJ (bi - v) (6.37)
The steady-state occupation fraction can be written as
x= ( 'P0 1(vb;b j - b; )I'P) (6.38)
The interaction terms Vi can be expressed in terms of the bs and b t s.
[S 0] - 1 V is then a six-order polynomial in v, while p appears to all order
in this operator. One therefore adopts v as the expansion parameter. If
the lowest-order calculations are easy to perform, the higher orders soon
become very tedious. Moreover, it is not clear how the series converges, if
it does.
This shows that, even more than for the equilibrium case, numerical
simulations are very necessary to study such systems.
Cellular automata approaches. The cellular automata version of this
model is straightforward. One considers a d-dimensional lattice. Each cell
of the lattice j has two possibles states : 1 '¥1 ) = 1 0) or lA). The cellular
.
automata probabilistic rules are
=
If I 'I'J )( t) 1 0) then :
{
!
I 'I'J )( t + 1) = 10lA)) ', w1t
with probability (1 - p),
h probab·1·1 1ty p . ( 6. 3 9 )
If I 'I'J )( t) = lA) then :
lA), with probability p if the site j has one nearest
neighbor empty,
I'PJ )( t + 1) = lA), with probability 1, if all the neighbors are occupied,
10), with probability (1 - p) if the site j has at least one
nearest neighbor empty.
(6.40)
Simulation of the above cellular automata rule has been performed for
chains of lengths L between 30 and 65 ' 536 [191]. As can be seen on
figure 6.1, a second-order phase transition is obtained at Pc = 0.465. The
order parameter critical exponent f3 extracted by fitting the data with the
relation 6.3 over the range 0.001 � (Pc - p) � 0.02 is f3 = 0.280 ± 0.010.
This shows that a mean-field-like approximation gives a qualitatively
reasonable phase diagram, but overestimates the critical probability Pc
and gives a poor value for the exponent {3.
For the two-dimensional case, one-site and pair mean-field approxima
tions predict a second-order phase transition for, respectively, Pc = 0.800
and 0. 785. Here again the value of the order parameter critical exponent is
f3 = 1. The results of numerical simulations of the CA rules lead to a phase
diagram topologically similar to the one obtained for the one-dimensional
case. The critical probability is Pc = 0.747. The order parameter critical
6.2Simp le interac ting partic le systems 241
exponent f3 extracted by fitting the data with the relation 6.3 over the
=
range 0.001 ::::;; (Pc - p) ::::;; 0.02 is f3 0.52 ± 0.01.
Fig. 6.2. Typical microscopic configuration in the stationary state of the CA Ziff
model, where there is coexistence of the two species. This simulation corresponds
to the generalized model described in equation 6.53. The gray and black dots
represent, respectively, the A and B particles, while the empty sites are white. The
control parameter XA is larger in the right image than it is in the left.
1 0)
step of the automaton, I C) will become I B) or depending upon the
fact that a nearest neighbor cell is empty and ready to receive the second
B atom of the molecule B2 . This conditional state is necessary to describe
the dissociation of B2 molecules on the surface.
The time evolution of the CA is given by the following set of rules,
fixing the state of the cell j at time t + 1, llJlj)( t + 1), as a function of the
state of the cell j and its nearest neighbors (von Neumann neighborhood)
at time t:
= 10)
{
R1 : If l lJlj)( t) then
l A ) with probability XA
lJll j)( t + 1)= (6.4 5 )
I C) with probability (1 - XA)
=
{ 10)
R2 : If llJlj)( t) l A ) then
if at least one of the nearest neighbor cells
=
l lJlj)( t + 1) of j was in the state IE) at time t ( 6.46)
l A ) otherwise
=
{ 10)
R3 : If l lJlj)( t) IE) then
if at least one of the nearest neighbor cells
=
llJlj)( t + 1) of j was in the state l A ) at time t ( 6.47 )
I B) otherwise
6.3 Simple models of catalytic surfaces 245
0.40
0.20
Fig. 6.3. Stationary state phase diagram in the simple CA model of the Ziff
problem. XA is the concentration of the A species injected on the surface. XA. and
XB are respectively the coverage fractions of A and B.
{
R4 : If l lfJj ) ( t ) IC) then
1 0) if none of the nearest neighbors was
=
l lfJj ) ( t + 1) in the state IC) at time t (6.48)
I B) otherwise
This last rule expresses the fact that the atoms of Bz can be adsorbed
only if they had been dissociated on two adjacent cells, that is if at least
two adjacent cells were empty at time t 1. Rules R1, R4 describe
-
=
R2 : If ltpj ) ( t ) lA ) then
=
ltpj ) ( t + 1) { IO) if the MNN
of j was in the state I B ) at time t
lA ) otherwise
(6.50)
=
R3 : If ltpj ) ( t ) I B) then
=
ltpj ) ( t + 1) { IO) if the MNN
of j was in the state lA ) at time t
I B) otherwise
(6.51)
6.3 Simple models of catalytic surfaces 247
Fig. 6.4. Illustration of the modified rules R2 and R3. The arrows select which
neigbor is considered for a reaction. Dark and white particles represent the A
and B species, respectively. The shaded region correspond to cells that are not
relevant to the present discussion such as, for instance, cells occupied by the
intermediate C species.
Fig. 6.5. Illustration of the modified rule R4. Light gray particles represent
the intermediate C species and the arrows indicate how the matching pair is
done. The shaded region corresponds to cells that are not relevant to the present
discussion such as, for instance, cells occupied by A or B particles. White cells
are empty cells.
R4 : If l1pJ ) ( t ) = I C ) then
11pj ) ( t + 1) =
{ I1 0)B )
if MNN is in the state I C ) at time t
otherwise (6.52)
The above rules R2 and R3 are illustrated in figures 6.4 and 6.5.
The simulations made on this model [194] reproduce the correct orders
of the transitions (second and first order), but the limits of the reactive
window are respectively 0.0495 and 0.0546, illustrating once more the
nonuniversal character of these quantities.
248 6 Nonequilibrium phase transitions
{
For example rule 6.49 can be modified as follows :
If l lJJj ) ( t) 1 0) then
=
l A ) with probability XA
llJJj ) ( t + 1)
= I C ) with probability XB2 (6.53)
1 0) with probability 1 - XA - XB2
In addition, since empty sites may now be present on the lattice during the
adsorption B2 2B a generalization of rule 6.52 can be considered: a cell
�
in the intermediate state C will give two adjacent B atoms if its matching
arrow points to an empty site which is not pointed to by another C state.
This new rule is illustrated in figure 6.6.
The phase diagram obtained for this generalized CA Ziff model is given
in figure 6.7, with the value XB2 0.1. Here again, the phase diagram is
=
topologically similar to the case XB2 1 - XA, but the locations of the
=
(a) (b)
Fig. 6.6. A less constrained adsorption rule B2 - 2B when some cells may be
empty. Here the B 2 molecule (or C state) is represented as two disks on top
of each other. Dissociation is possible if the upper disk can move to the site
indicated by the arrow without conflict with other moves. Note that we still
accept dissociation when a pair of matching Cs sit next to each other.
o B coverage
• A coverage
o �����--����
0.04 0.06
Fig. 6.7. Stationary state phase diagram corresponding to the CA Ziff model
given by relation 6.53 and the generalized dissociation rule.
R2 ( t ) "' tz (6.56)
If one is not precisely at criticality, these quantities will have an upward
(downward) curvature in a log-log plot. Thus this method leads to a
precise determination of Pc·
Another approach is based on the concept of damage spreading [201] .
Here again, we discuss this method in the context of the A model.
Let us consider a system with N sites and two different configurations
A = { tp f } and B = { tpf } of this system. The Hamming distance between
these two configurations is defined as
DAB = 1 "' A B
N � llp i - lpi I
I
(6.57)
distance) as
(6.68)
252 6 Nonequilibrium phase transitions
1 X
DAB = - --'-'--- (6.70)
( )
N XA 1 - XA
Thus the Hamming distance is exactly related to the nonequilibrium
one-point and two-point correlation functions, XA and Coi · In particular,
when p approaches Pc , the Hamming distance diverges allowing a precise
determination of Pc ·
II
Fig. 6.8. Behavior of probabilistic rule 94 for p=O.O, 0.05, 0.25 and 1 .0, respec
tively. A kink corresponds to two nearest neighbors black sites while an anti-kink
corresponds to two nearest-neighbor white sites.
[(01 1 ) ; ( 1 10)] � { [0 ; 0]
[1 ; 1]
with probability p
with probability ( 1 - p )
(6.72)
For very small p, the system orders itself spontaneously : there are two
symmetric adsorbing states. The corresponding patterns are, respectively,
vertical stripes with lp i = 0, lp i+ l = 1 . Starting from a random initial state,
after some short transients one finds small domains separated by doubly
occupied neighbors called kinks. For p = 0, the kinks are frozen and
between two kinks one observes periodic patterns with very short periods.
For small p the kinks are no longer frozen. They move by annihilating
random walks and the kink density decays in time as (pt ) 1 1 2 . If p grows
further, in addition to enhancing the random walk it leads to a splitting
of the kinks according to : kink� kink+ [kink+antikink] . For small p, the
annihilation procedure dominates but if p is larger than a critical value
Pc = 0. 1 3 ± 0.02, a single kink in the initial state is enough to create a
completely disordered state. Note that in the limit p = 1 , the rule is simply
Wolfram's rule 22 which has a chaotic behavior. These different situations
are illustrated on figure 6.8.
At p = Pc, the density of kinks decreases as
nkink ,...., crx, rx = 0.27 ± 0.08 (6.73)
and for p � Pc from above, the density of kinks in the stationary state
tends to zero as
nkink "' (p - pc)f1, f3 = 0.6 ± 0.2 (6.74)
These exponents are not those of directed percolation.
254 6 Nonequilibrium phase transitions
This cellular automata model is in fact closely related to the more gen
eral problem of the branching annihila ting random walk (BAR W) [204] .
In the BARW, particles on a lattice are subject to the following sequen
tial dynamics. First, one chooses one particle randomly. With probability
q this particle jumps to a nearest-neighbor site. If the new site is already
occupied by one particle, both particles annihilate. Secondly, with proba
bility ( 1 q) the particle gives birth to n offspring distributed randomly
-
6.5 Problems
7. 1 . 1 One-dimensional waves
In this book, we have already encountered one-dimensional wave propa
gation. The chains of particles, or strings, discussed in section 2.2.9 move
because of their internal shrinking and stretching. From a more physical
point of view, this internal motion is mediated by longitudinal backward
and forward deformation waves. This is easy to see from equation 2.46
q l ( t - 1 ) - 2q l (t) + q l ( t + 1 ) 2(q2 ( t) - q 1 ( t) - a )
qi( t - 1 ) - 2qi( t) + qi( t + 1 ) q i- l ( t) - 2qi (t) + q i+l( t)
qN (t - 1 ) - 2qN ( t) + qN (t + 1 ) 2(qN-l ( t) - qN ( t) + a )
256
7.1 Wave propagation 257
_ .,.__ .,.__ _
Fig. 7. 1 . The string motion i s due to forward and backward traveling waves f t
and h
where q i ( t ) is the location along the x-axis of particle i in the chain and
a the equilibrium separation between consecutive particles. Now if we
define Ll i as the deformation from the equilibrium distance
Ll i = q i+ l - q i - a
we obtain the following equations of motion
Ll t ( t + 1 ) - 2c5 t ( t ) + Ll t ( t - 1 )
L'l 2 ( t ) - 3 L'l t ( t ) ( 7. 1 )
Lli ( t + 1 ) - 2L'l i ( t ) + Lli ( t - 1 )
Lli -t ( t ) - 2 L'li ( t ) + Lli + l (t) (7.2)
Ll N-l ( t + 1 ) - 2Ll N-! ( t ) + LlN-! ( t - 1 )
Ll N_2 ( t ) - 3 LlN-l ( t ) ( 7.3 )
Disregarding momentarily equations 7. 1 and 7.3, it is interesting to note
that equation 7.2 admits a very simple set of solutions
Ll i ( t ) = f ( i - t) + g ( i + t ) ( 7.4 )
as can be seen by direct substitution. The functions f and g can be chosen
arbitrarily and correspond to a backward and forward traveling wave.
As we shall see, equations 7. 1 and 7.3 impose reflection and transmission
boundary conditions on these longitudinal deformation waves.
Now, let us return to the interpretation of the motion of our string
particles, as presented in section 2.2.9. Black and white particles alternate
along a chain and each type of particle moves at every other time step
as if they were linked by a spring. Suppose we define ft ( i, t ) as the
deformation on the left of a moving particle (say a black one) and h ( i, t )
as the deformation on its right
ft ( i, t ) qi - qi-1 - a
h ( i, t ) q i+ l - q i - a
for i odd or even depending on which kind of particle is going to move.
According to the rule of motion, our black particle moves so as to
exchange ft and h, as shown in figure 7. 1. Then, at the next iteration, ft
and h are no longer defined with respect to the black particle that just
moved, but with respect to the white particles around it.
This choice of defining ft and h as the left and right deformation,
respectively, is less general than the solution 7.4. This is a particular
258 7 Other models and applications
case where f1 is zero when h is not and vice versa. But this choice
corresponds to the actual motion of the string and leads to the same kind
of formulation as found in lattice gas automata
( h ((ii +- 11,' tt ++ 11 )) ) = ( 1 1 ) ( h ((i,i, tt)) )
h
0
0 h
7. 1 .2 Two-dimensional waves
The problem of modeling two-dimensional waves is more difficult because
waves spread as they propagate in space. Thus, a representation of
the deformation in terms of a field of integer values is in general not
possible. However, the construction proposed in the previous section can
be generalized provided that the fi s are real-valued.
Let us assume we have a two-dimensional lattice of black and white
particles organized in a checkerboard fashion. Figure 7.2 shows a basic
7. 1 Wave propagation 259
element of this lattice, made of a central black particle and its four white
connecting neighbors.
As in the one-dimensional case, the black particles move at even time
steps and the white ones at odd time steps. In figure 7.2, the motion is
relative to the positions of the four connected white neighbors at rest.
The rule of motion for the black particles is to j ump to the symmetrical
position with respect to the center of mass (the cross in the figure) of the
four white surrounding particles.
Let us denote the location of the black particle by
ri,J = ( x i,J , Yi,J )
The surrounding white particles will be at positions r i-l,J , r i+l,J , ri,J-1 and
ri,J+l · We define their x-axis separation to the central black particle as
f l ( i, j, t )
X i-l,J ( t ) - X i,j ( t )
= /3 ( i, j, t ) X i+l,j ( t ) - X i,j ( t )
=
X i,j-l ( t ) - X i,j ( t
f2 ( i, j, t ) = ) j4 ( i, j, t ) Xi,J+l ( t ) - X i,j ( t ) (7.6)
=
( ) ( )
particle in figure 7.2. Thus we obtain the following evolution law for the
x-separation between black and white particles
fi ( i + 1, j, t + 1) f i ( i, j, t )
h ( i, j + 1, t + 1 ) w h ( i, j, t )
h ( i - 1, j, t + 1 ) = h( i, j, t ) (7.9)
( )
f4 ( i, j + 1, t + 1 ) f4 ( i, j, t)
where the propagation matrix W reads
1 1 -1 1
1 1 1 1 -1
Wjree = 2 (7. 10)
-1 1 1 1
1 -1 1 1
where the subscript "free" indicates propagation in a medium with neither
reflection nor absorption.
It is important to notice that the above dynamics requires to work with
real-valued quantities /;. In other words, by following our motion rule 7.7,
the white and black particles will eventually jump off lattice sites. Thus,
equation 7.9 is no longer a pure cellular automata rule as it was in the
one-dimensional case but, rather, a lattice Boltzmann dynamics. Figure 7.3
illustrates this dynamics.
Equation 7.9 with W given by 7.10 leads to the propagation of two
dimensional waves. This will be proven below but, at this point, it should
be clear from the physical interpretation we gave : masses are vibrating
around their local equilibrium positions, and the deformation propagates
like a wave. Note that equation 7.9 concerns only the deformations along
the x-axis. This is enough to simulate two-dimensional wave propagation
since an x-deformation also propagates along the y-axis.
7. 1 Wave propagation 26 1
fl - �
, - -�·� · -
----- -<---
2
_l
Fig. 7.3. Scattering of an incoming flux h at a node. The value of each outgoing
flux is indicated on the arrows.
t= IO t=20 t=30
t=40 t=50
It differs for the horizontal and vertical coupling (it is zero in the latter
case, for the x-motion).
A bond breaks if one of the f; (or g; ) becomes too large. Each time a
bond breaks a virtual particle is added. Figure 7.5 illustrates the various
stages of this process. Of course, it is possible to distribute the breaking
thresholds randomly among the bonds, so as to produce a disordered
media.
Some modifications to the model can be considered. For instance,
instead of comparing the distance f; directly with the breaking threshold,
7.1 Wave propagation 263
Fig. 7.6. Different patterns of fracture obtained with our model. Dissipation is
an important parameter to obtain a smooth fracture, as in (d) rather than many
micro-fractures and branching as in (b).
0
'-'
-..
,.-.,
"0
0
0
0..
"'
�
()
«!
....
()
'-'
0
0 1 000
time
Fig. 7.7. Propagation speed of the factures shown in figures 7.6 (b) and (d)
(upper and lower curve, respectively).
( )
be implemented by bouncing back the incoming flux with the same or the
opposite sign. Thus we modify the free propagation matrix 7. 10 as follows
0 0 ±1 0
+ 0 0 0 ±1
W;;f = ±1 0 0 0
(7. 1 1 )
0 ±1 0 0
where the sign ± 1 depends whether or not there is a change of phase
during the reflection step. A positive value corresponds to a free end while
a negative one corresponds to an absorbing boundary.
With this approach, we can simulate free wave propagation and total
reflection at obstacles. The lattice sites that represent obstacles are charac
terized by a "scattering" matrix wrt ' while free sites are evolved according
to the scattering matrix Wf ree· In the next section, we will return to the
general problem of modeling a change of medium and different speeds
of propagation in each of them. For the time being let us explain how
attenuated or damped waves can be modeled.
In the interpretation of black and white moving particles, attenuated
waves may be seen as resulting from a particle movement with friction :
the displacement is amortized and the moving particles jump closer to
their initial position than they would otherwise do. Therefore, similarly to
what we did in equation 7.7 we shall now write
4
7.1 Wave propagation 265
( )
Repeating the same calculation for the other f;s yields the attenuation
propagation matrix
Watt(oc) =
!4 ; �;1 ��1 ( 7. 1 3 )
ex ex ex ex
4 -1 4 4 4
� �-1 � �
Clearly, when oc = 0, Watt = Wr--;f and when oc = 2, Watt = Wtree · In the
general case, Watt can be written as
0( 0(
Watt = Wf ree + ( 1- 2 ) Wr--;f
2
Damped waves are very useful to mimic an infinite system with a finite
lattice : by smoothly absorbing the waves at the boundary of the lattice, it
is possible to prevent them from coming back into the system as if they
had freely continued their propagation beyond the limits of the simulation.
For the sake of completeness, we also give here the expression for
Watt for a one-dimensional system, which can be obtained with a similar
derivation.
1D : Watt(oc) = � .:_ 1 2 �
� (
�- 1 )
( 7. 1 4 )
Fig. 7.8. Lattice Bolztmann wave model and focusing of a plane wave by a
parabolic mirror.
The dispersion relation. By using A and r as the lattice spacing and time
step respectively, equation 7.9 can be rewritten as
4
fi ('r + TC j, t + r) = L Wi,Jfj('r, t )
J =l
where W is given by equation 7.10. The dispersion relation of this lattice
Bolztmann wave model can be obtained by considering a solution of the
form
fi ( r , t ) = A i e i(w t+kr )
After substitution, it is found that the above expressiOn 1s a solution
provided that w and k satisfy
e iwr = �( K ± i �4 - K2 ) ( 7. 1 5 )
where K = cos(k l A)+cos(kzA) and k1 and kz are the two spatial components
of k. A Taylor expansion of 7. 1 5 gives, with k 2 = k y + k�
w 2 r2
2
A2k 2 Ak
iwr - -- + O(r 3 ) = - 4 + 0 (A4 ) ± i J2 + 0 (A 3 )
[ ]
( 7. 16 )
or, equivalently
�
fi(r + -rvj , t + -r) - fi(r, t ) = (!�0) (r, t ) - fi(r, t) ) ( 7. 1 8 )
where, here, v i designates the velocity vector along the ith lattice direction.
The quantities fi are interpreted as the average number of particles
traveling with velocity V i at site r and time t. At this point the relaxation
parameter � is a free quantity. We shall, later, see that only one specific
value is acceptable to get clean wave propagation.
We now introduce the conserved quantities
'¥ = L mdi j= L mdi v i ( 7.19 )
where mi are free parameters corresponding to the mass of the particles
traveling in lattice direction i.
( 7.20 )
4
summed over i. From equation 7.19, this amounts to asking that
'¥ = 2: md�0)
i =O
4� ( 4 )
Using the expression 7.20 for f�0) , we have
1 ...
'¥ = 'l' � mI· a·I + -2 �
� m1· b·v·
1 1 ·J
1=0 2V 1=0
tz=O (t1=0 )
and
2� 2 mi b i ( v i ] ) v i
J = '¥ mi a i i\ + ·
But we know from chapter 3 that 2:� 1 V i = 0 and 2:� 1 VirxVip = 2v 2 �rxp,
since the v i s point along the main directions of a square lattice. Thus, if
(�4 )
mi b i = 1 we simply have (remember that v0 = 0 )
1
mi b i ( Vi · J ) v i = J
� -+ ...
-+
2v 2
Therefore mass and momentum conservation can be enforced if
mi = m ai = a bi = b for i = 1, 2 , 3, 4
and
The propagation matrix. We can now rewrite the evolution law 7. 1 8 using
the above expression for f� 0) . We obtain
( l ( l
+ (am - � ) fi+2 + amfi+3 + amofo] (7.23)
(
f4(r + t"V4, t + t") f4(r, t )
)
where
aomo + � - 1
W = -;1 ; amo am + �
amo am + �
�1 am
�i
am
�1 �i
<:. amo am - � am am + � - � am
amo am am - � am am + � - �
(7.25)
By choosing the parameters in matrix W in an appropriate way, we find
the matrix Wfree given in equation 7. 10. Indeed, in the case that there is
no rest particle (i.e. 0,fo = ao =
0), we have
a = -1m 4
Now, if � is chosen as
� = -1 2
W reduces to
-mo1 1
0
1
0
1
0
1
0
2m 2 2 -2 2
!!!!!. 1 1 1 1
W= 2m 2 2 2 -2
mo 1
-2
1 1 1
2m 2 2 2
mo 1 1 1 1
2m 2 - 2 2 2
7. 1 Wave propagation 27 1
which, except for the first row and column, is exactly the same as 7. 10.
Thus, with fo = 0 the lattice BGK formulation 7. 1 8 is equivalent to the
wave model developed in section 7. 1.2.
Note also that the quantities fi are not guaranteed to be positive. For
some choice of 1p and J , it is certainly possible to make ff0l negative as can
be seen from the definition 7.20. Actually, it is clear that fi will be negative
since they represent an oscillating wave. This is not very satisfactory if the
fis are interpreted as an average number of particles. However, since the
dynamics is linear, we can always think that fi describes the fluctuations
around some large positive constant value.
Adjusting the propagation speed. Equation 7.25 is a five by five matrix and
the presence of rest particles allows us to change the wave propagation
speed. In fact, each lattice site can have its own refraction index and
it is then possible to model propagation through different media or in a
disordered medium.
To obtain the resuting wave equation, we first compute the equation
governing the evolution of the conserved quantities 'I' and J . To this end,
we shall use the same formalism (Chapman-Enskog expansion) as used in
section 3.5.2. Since we have mass and momentum conservation we know
that we have (see equations 3.87 and 3.88 )
at 'I' + opJp o ( 7.26 )
[
otlrx + op IIrxp + � (eot1 II�t + oySrxpy ) - 0
] ( 7.27 )
where Ot = E Ot1 + e2 ot2 is expressed in terms of the two relevant time scales
and e is the expansion parameter.
Using the usual definitions (see chapter 3 ) , the momentum tensor is
II = rr (O) + eii ( l)
and its the zeroth order is given by
rr�o; - L mdf0l VjrxVip
= am'I' L VirxVip + ::2 L VjrxVipViy
2amv 2'Pb rxp
Hence
( 7.28 )
and
( 7.29 )
where we have used the continuity equation 3.77 for the last equality.
272 7 Other models and applications
where iha is the order c of the spatial derivative with respect to coordinate
a, i.e. Oa = E O!a· Thus,
Errap( l)
Remember that, since we are here working on a square lattice, without the
diagonals, Tapyo is not isotropic (it depends on the particular orientation
of the vectors v i) . This is of course not desirable and we shall see below
how this problem is solved. For the time being we have
(7.3 1 )
Finally, we also have to compute
s�jy L miViaVipViy /}0)
= � I: ViaVipViy Vio) fo
2 2 (
The term involving 'I' vanishes since it contains the product of three v i s.
Therefore
T
apoyS (0) _- T 2 Tapyoopoylo
apy (7.32)
2 4v
We can now collect relations 7.28, 7.29, 7.3 1 and 7.32, and plug them
into equation 7.27. After grouping the terms, we have
Ot la + 2amv 2 oa'l' + (2( - 1 ) amrv 2 o a divJ - (2( - 1 ) 4Tv 2 Tapyoopoylo
�
0 =
(7.33)
As we can see, by choosing ( = 1 /2, we cancel the anisotropic term
containg Tapyo as well as the dissipative term (2( - 1 ) amr v 2 8a divl
In the fluid model of section 3.5.2, ( plays a similar role in the sense
that ( = 1 /2 corresponds to the limit of zero viscosity. But there, this
7. 1 Wave propagation 273
1 1 - 2n2
mo
1 - 2n2
1 mo
W(n) =
2n2 m
1 - 2n2
mo
1 - 2n2 1 1
mo
m
(7.36)
274 7 O ther models and applications
Fig. 7.9. Lattice Bolztmann wave model and focusing of a plane wave by a
convex lens with a refraction index of 1 .7. The s ource is on the leftmost side of
the system. Energy is shown in this simulation. Light gray denotes high energy.
where n is the refraction index defined as the ratio of the free propagation
speed v / J2 to the propagation speed c = 2am v 2
v2
n2 = -
2c2
Figure 7.9 illustrates the propagation of a plane wave through a lens
shaped medium. The lattice points outside the lens are characterized by
a matrix W ( l ) and fo Ci\ t) = 0 while the internal sites have W( n ) with
n ?.: 1. Due to the change of speed between the two media and the shape
of the interface, we observe a focusing of the incident plane wave.
Time reversal invariance. In the previous section, we saw that the relax
ation time � should be set to 1 /2 in order to obtain a good wave model.
There is another intere sting way to see this, using time reversal symmetry.
Time reversal is a fundamental symmetry of microscopic physics and is
clearly satisfied by the wave equation which has only second-order time
derivative. We shall see here that time reversal implies
� 1
s = -
2
For this purpose, let us write equation 7. 1 8 as
( 7. 3 7 )
where .f stands for the five-element column vector whose ith component
is h Thus, in order for equation 7 . 3 7 to be invariant under time reversal,
one should have ( see section 2.2.7)
W R.f ' =
Rf
Using that ( see equation 7.20)
b· �
a z· 'P(.f) + _l2 iJ '. · J(.f)
f
�)
. t
=
. 2v
_ __
.
we have
( WR.f' )t =
(1 - z) (R.f ' )t + z[ a t 'I' (Rf ' ) + �2 vt · l (R.f ' ) ] (7.39)
'P(R.f ' ) =
L mJ (Rf ' )J =
mo (R.f ' ) o + L m (Rf ' )J
j=O j f'O
4
Similarly,
4
I (Rf ' ) L 11lj Uj (Rf ' ); =
L u_; m (Rf);
} =0 j fO
4
- L fnVJ+2i"J+2 =
- L m_;v_;f}
J=!=O j=O
4
L.. m}-ii..;f";
� (because of momentum conservation )
} =0
- l (f)
Substituting these expressions for 'P( Rf' ) and I ( Rf ' ) into 7.39 yields ( for
i j 0)
( W Rf ' )t =
( 1 - l ) f'i+2 +
l fa n�
l rC )
· r· b �
- 2v2 Vi .
J� ( f) l
. J
Z Z
276 7 Other models and applications
1 1 1 (0) 1 (0)
1 - 1 -
( z ) 2 fi+2 + z ( z ) fi+2
1 -
1 1
2 -
1 (0)
whose solution is
( = 1 /2
Similarly, the condition ( WRf ' ) o = ( Rf) o = fo also imposes this value for
(.
I n conclusion, with ( = 1 /2, the LBGK formulation o f the wave model
becomes
(7.40)
This expression gives a natural way to extend the model to a three
dimensional cubic lattice (see also [21 2] ) since our derivation did not use
any specific properties of the two-dimensional space.
In the previous section, we saw that ( = 1 /2 makes the viscosity
vanish, that is requiring time reversal invariance is equivalent to removing
dissipation (actually breaking of time reversal symmetry is the standard
definition of dissipation). This observation shows that the LBGK fluid
models of section 3.5.2 is not reversible in time. Thus, this symmetry is
lost when compared with the cellular automata fluids such as FHP. In the
cellular automata approach, the microdynamics is time reversal invariant.
This symmetry breaks only when the average behavior is considered to
obtain the Navier-Stokes equation.
deformation state when one looks at them, we may also argue that kinetic
energy vanishes at the observation time steps.
In this section we want to obtain the condition that the evolution matrix
W must satisfy in order to conserve the energy
4
E (r, t ) = L il = (!If)
i=O
where (f l f) denotes the scalar product.
( OJ
It is easy to check that the matrix W given by 7.25 obeys
RWR = W
where R is the flux reversal operator introduced in 7.38
1 0 0 0
0 0 0 1 0
R= 0 0 0 0 1
0 1 0 0 0
0 0 1 0 0
The above property means that W and R commute (note that R 2 = 1 ) and
reflect a natural spatial symmetry of the dynamics : if an incoming flux fi
enters a site it produces outgoing flux fj = Wji/i . Now the incoming flux
is fi+2 = ( Rf) i we shall get the same outgoing flux, but pointing in the
reverse direction, namely ( Rf' )j .
Energy conservation requires that the energy E ' after scattering is equal
to the energy before scattering
E ' = (f ' lf' ) = ( Wf i Wf) = ( W j_ Wf l f)
where W j_ is the transpose of matrix W. Thus energy will be conserved if
w j_ w = 1
In the case that the dynamics is invariant under time reversal (and that
is really what we want here), we learned in the previous section that
WR W = R
Since we have also shown that R W R = W we obtain that
W2 = 1
Therefore the condition of energy conservation demands
W j_ = W
or, in other words that W be symmetric. We can return to equation 7.36
and use the extra freedom in the choice of � to make W symmetric. The
278 7 O ther m odels and applications
condition is
- = 2 v�
mo m . mo
- - - ( 4n2 - 4) or n2 - 1
m mo m
Thus, with energy conservation and time reversal invariance, the wave
propagation matrix in a medium of refraction index n � 1 is
Fig. 7. 10. Wave propagation in a region of the city of Bern, as predicted by the
lattice Boltzmann model. The gray levels show, from white to black, the intensity
of the wave. The source antenna is located at the dot mark. The white straight
line indicates the path along which intensity measurements have been compared
with real measurements.
personal mobile communication systems ( 900 and 1 800 MHz). The reason
is that the discreteness of the lattice is not fine enough to work at such a
high frequency. The renormalization proposed is the following [215] .
We suppose that the measured (real) amplitude A 3 v of the wave can be
approximated by the outcome azv of our simulation using the following
relation :
( 7.42 )
where A is the actual wavelength used in the measurement and A. the
simulated one. The quantity (> represents the distance around the buildings
separating the transmitter from the receiver.
The rationale for this expression is that it is true for wave propagation
in vacuum. It is then assumed that the effects of multiple reflections
and _ scattering on buildings is already captured in azv : these are the
geometrical contributions which are expected to be the same in two
and three dimensions. Finally, since the distance traveled by the wave
is not accounted for in the same way in 2D and 3D, the rest of the
renormalization is assumed to be included in the function f>Ol Here we
have used for f>(r) the so-called Manhattan distance which is the shortest
distance following the lattice edges representing the discrete city map.
Figure 7. 1 1 gives a comparison between the predictions of our model
and real measurements. The renormalization given by equation 7.42
has been used to compute the wave intensity along the path shown in
figure 7.10. We observe good qualitative agreement between the two curves
and also acceptable quantitative agreement knowing that our simulation
neglects all small-scale details of the real city and that each building is
assumed to produce the same absorption and reflection. Finally, it should
be noted that, in the simulations, some streets are too narrow to permit
wave penetration which may explain why some intensity peaks are not
correctly reproduced by the simulations.
�
- 1 05 r
� -115 iI
"' t\
tl..�
� - 1 25 �-�
.. I \
I \
. \
- 1 35 '
i \/
- 1 45 L--�--�----�j
0 1 00 200 300 400 0 1 00 200 300 400
distance (in meters) along the street
vapor
rapidly recondense into a thin layer over the substrate. On the other hand
a non-volatile liquid covers the substrate much more slowly than a volatile
substance, because it has to flow.
It has long been observed that the spreading of non-volatile liquids
are often accompanied by the formation of a so-called precursor film
at the interface between the substrate and the liquid [220] . This film
spreads faster than the bulk of the droplet and is typically a monolayer
of molecular thickness [22 1] . Sophisticated experimental techniques [222]
have also shown that the fluid spreads as a series of superposed molecular
layers.
As a result of these features, the spreading of liquid is a very interesting
phenomena for which much is still to be understood from first principles.
The Navier-Stokes equation, commonly used to described fluid motion, is
not the appropriate approach in this case. The usual boundary condition
in the Navier-Stokes equation is a "no slip" condition at the liquid-solid
interface. A zero velocity field at the boundary is clearly incompatible with
droplet spreading [223] and the existence of a precursor film. Furthermore,
the Navier-Stokes equation assumes that each fluid element is composed
of many molecules. This hypothesis is not true in our case since the
layering which is observed occurs at a molecular scale.
Several theoretical models have been proposed to describe the spreading
of a droplet [224, 225] . They assume the existence of two-dimensional
layers of fluid, moving on top of each other with some viscous force.
These models give predictions of the speed of the precursor film, the
droplet profile and the partial/complete wetting transition.
Wetting and spreading phenomena are of great importance in industrial
processes used in the manufacture of lubricant and paints. From a nu
merical point of view, spreading can be investigated in terms of molecular
dynamics simulations, since it results from inter-molecular attractive or
repelling forces. However, molecular dynamics requires massive compu
tational resources and we are interested here in the long time regime for
which this approach may not be appropriate.
284 7 Other models and applications
Surface tension. Let us now discuss the way to compute the force F in
our model. Surface tension is produced by short-range mutual attraction.
Keeping only interactions with nearest-neighbor particles, we define
F('r, t) = f L Cj L nj ( r + Cj, t) (7.46)
i,=1,6 j=1,6
where, as usual, Cj denotes the unit vector joining two nearest-neighbor
sites (j is labeling the direction) and f is some constant describing the
magnitude of the elementary attraction.
Figure 7. 1 3 illustrates the result of such an interaction. The force F
given in equation 7.46 (with f = 1 ) causes the particles at the middle site
to change their velocities according to equation 7.44. This example shows
a situation where equation 7.44 can be fulfilled exactly. This is usually not
the case and one has to find the "best" approximation to equation 7.44
by inspection of all possible distributions of the nis. Of course, a lattice
Boltzmann version of this model would give much more flexibility in
the implementation of equation 7.44 because the n i would no longer be
restricted to the values 0 or 1 .
I n order to adjust the surface tension, a threshold quantity can be
introduced in the rule : the mutual interaction F between fluid particles
is retained only when it is larger than this threshold value. In particular,
this prevents a single particle from capturing another one and forming
with it a stable oscillating pair,
In the same spirit, we can also introduced in the rule a "temperature
effect," that is some probability that surface tension does not operate at a
site. When this happen, the particles are submitted only to the action of
the external force. The purpose of this probability is to cause evaporation,
i.e to let particles escape from the attraction of the liquid bulk.
286 7 Other models and applications
't,
-
',
',
, ,
'
',
Fig. 7. 14. The various boundary conditions that have been used when a particle
meets a solid wall. (a) Specular reflection, (b) bounce back condition and (c)
trapping wall condition. Condition (c) is necessary to produce a precursor film.
y =
where .A is the length of the lattice spacing.
In the second mapping technique (see figure 7. 1 6), odd and even hor
izontal lines are treated differently. Odd lines are shifted half a lattice
spacing with respect to even lines. In addition to north, east, west and
south neighbors, lattice sites on an even line are also connected to their
north-east and south-east neighbors. On the other hand, lattice sites re
siding on an odd line are connected to their north-west and south-west
288 7 Other models and applications
Even j : X = Ai
Odd j :
The advantage of the first technique over the second one is that the
transformation is the same for all j and that the connectivity of each
lattice site is identical. On the other hand, boundaries are more easily dealt
with in the second method. Periodic boundary conditions in the vertical
direction are preserved in the mapping of figure 7. 1 6 while they are not
in figure 7. 1 5. Also, a vertical wall is inclined in the computer topology
of the first mapping, while, in the second case it can be implemented as a
straight vertical line. Thus, despite its more difficult implementation, the
second method is probably safer.
7.2- Wetting, spreading and two-phase fluids 289
Fig. 7. 1 7. Free surface of a liquid in a capillary tube : (a) a wetting situation ; (b)
inert walls ; and (c) a non-wetting situation. Gravity is acting downwards.
With both specular reflection and bouncing back conditions, the droplet
is seen to spread away. However, no precursor film are observed. The
fluid may be considered as being at the transition between partial and
complete wetting. We can measure the position x0 (t ) of the first fluid layer
as a function of time (the existence of layers of molecular size is obvious
here, due to space discreteness). We observe, in figure 7. 19(a), a t 1 12
behavior
xo(t ) "" t 1 12
in agreement with the results of [226,227] . Upper layers are found to
advance also with power law, but with different exponents.
The droplet profile can be measured. It turns out to fit relatively well
a parabola, at least for the lower part of the droplet, as indicated on
figure 7. 1 9(b ) This result agrees with the prediction of [227] for the
.
(a) (b)
3 .6 IL-L......J--l----L---L---l..---1...--l.--'--' l --'---'---..L.--'-----'---L---'
0 2.8 1 10 36020
ln(t)
Fig. 7. 19. (a) Measurement of the time behavior of the first fluid layer at the
liquid-solid interface (horizontal advance versus time in a log-log scale) ; the
slope gives x '"" t0·5. (b) Measurement of the droplet profile at a given time, with
an appropriate change of scale ; only the first 3 1 layers are taken into account
(one-third of the total height) ; this plot indicates a parabolic profile, at least for
the foot of the droplet.
Fig. 7.20. Snapshots of a spreading droplet, as given by our model with the
"trapping" boundary condition. A rapidly advancing precursor film is present at
the liquid-solid interface.
292 7 Other models and applications
down from the bulk of the droplet. Figure 7.20 illustrates the formation
of the precursor film.
Although the way we introduced the interaction between the particle
is not exact from the point of view of momentum conservation, we see
that this microscopic approach exhibits several key features of the wetting
and spreading phenomena : the formation of a meniscus in capillary tubes
and the existence of a precursor film during the spreading of a droplet.
The trapping boundary condition provides a very intuitive explanation of
the formation of the precursor film. More complex situations, such as
spreading over a rough surface, could be studied easily with the present
approach.
The collision rule. Before we define more precisely the spin interaction,
let us return to the particle motion. A collision rule which conserves
mass, momentum and spin can be defined in analogy with the FHP rule
described in section 3.2.
We denote by s i(' r , t ) E { -1, 0, 1 } the state of the automaton along lattice
direction i at site r and time t ( s i = 0 means the absence of a particle).
Clearly the presence of a particle is characterized by sr = 1, regardless
of its spin. Thus, the collision term can be obtained by using sf as an
occupation number.
When a collision takes place, the particles are redistributed among the
lattice directions but the same number of spin + 1 and - 1 particles should
be present in the output state as there were in the input state. A way
to guarantee this spin conservation is to assume that the particles are
distinguishable, at least as far as their spin is concerned.
In figure 7.22 we show the case of a collision between two spin 1
particles and one spin - 1 particle. We assume that all particles bounce
back to where they came from (this is not the only possibility to conserve
momentum and any permutation of the three particles could be considered
294 7 Other models and applications
Fig. 7.22. Three-particle collision in a FHP Ising fluid. The black arrow indicates
a particle with spin + 1 while the gray arrows show particles with spin - 1 .
too). Thus, for such a three-body collision, the state after collision is
given by
si('f + 2c i, t + -r ) =
Si - SiSf+2 sf+4 (1 - sf+ 1 )(1 - Sf+ 3 )(1 - sf+ s )
+si+JSf+l Sf+ s (1 - Sf )(1 - Sf+2 )(1 - Sf+4 )
The identity of the particle appearing or disappearing in direction i is
expressed by the variables s i and s i+3, whereas the presence or absence of
other particles in direction j is indicated by s].
Similarly, by including the two-body collisions, we can express the full
collision term as
si("f + ACi, t + -r ) =
si
-SiSf+2 Sf+4 (1 - Sf+ 1 )(1 - Sf+ 3 )(1 - Sf+ S )
+si+JSf+1 st+ 5 (1 - st )(1 - sf+2 )(1 - st+4 )
-si sf+ 3 (1 - sf+ 1 )(1 - sf+2 )(1 - sf+4 )(1 - sf+ s )
+pqsi+1Sf+4 (1 - st )(1 - sf+2 )(1 - Sf+ 3 )(1 - sf+ s )
+p(1 - q)si+4 st+ 1 (1 - st )(1 - st+2 )(1 - Sf+ 3 )(1 - sf+ s )
+( 1 - p)(1 - q)si+2 st+5 (1 - st )(1 - Sf+t )(1 - sf+ 3 )(1 - st+4 )
+( 1 - p)qsi+ s st+2 (1 - st )(1 - sf+ 1 )(1 - sf+ 3 )(1 - st+4 )
( 7.47 )
where p and q are random Boolean variables that are 1 with probability
1/2, independently at each site and time step. These quantities are
necessary because, with two-body collisions, several collision outputs are
possible and one has to select randomly one of them.
2 0 1 2 0 1 2 0 1 2 0 1
0 1 2 0 1 2 0 1 2 0 1 2
2 0 1 2 0 1 2 0 1 2 0 1
0 1 2 0 1 2 0 1 2 0 1 2
Fig. 7.23. The three sub-lattices on the hexagonal lattice used for the syn
chrounous spin update. The values 0, 1, 2 label the sites according to the
sub-lattice to which they belong.
any two spins that are neighbors on the lattice. This is for the same reason
as explained in section 2.2.3 when we discussed the Q2R rule.
In a hexagonal lattice, it is easy to see that the space can be partitioned
into three sub-lattices so that all the neighbors of one sub-lattice always
belong to the other two (see figure 7.23).
Therefore, the spin interaction rule described above cycles over these
three sub-lattices and alternates with the FHP particle motion discussed
in the previous section.
It is of course possible to vary the relative frequency of the two rules
(Glauber and FHP). For instance we can perform n successive FHP steps
followed by m successive steps of the Ising rule in order to give more or
less importance to the particle motion with respect to the spin flip. When
n = 0 we have a pure Ising model on a hexagonal lattice but with possibly
a different number of spins per site.
If the temperature is large enough and a periodic boundary imposed,
the system evolves to a configuration where, on average, there are the same
number of particles with spins up and down. Of course, the situation is
not frozen and the particles keep moving and spins continuously flip. As
in a regular Ising systems, there is a critical temperature below which we
can observe a collective effect like a global magnetization (an excess of
particles of a given spin) and the growth of domains containing one type
of spin. This situation is illustrated in figure 7.24 and corresponds to the
case n = m = 1, namely one spin update cycle followed by one step of
FHP motion. It is observed that the critical temperature depends on the
frequency updating n and m.
Another interesting situation corresponds to the simulation of a Raleigh
Taylor instability. Two immiscible fluids are on top of each other and
the heavier is above the lighter. Due to gravity, the upper fluid wants to
penetrate through the lower one. Since the two fluids are immiscible, the
interface between them becomes unstable and, as time goes on, gives rise
to a mushroom-like pattern.
Gravity can be added to our model in the same way as explained in
7.2 Wetting, spreading and two-phase fluids 297
Fig. 7.24. Three snapshots of the evolution of the Ising FHP model below the
critical temperature. Particles with spin + 1 are shown in black while gray points
show particles with - 1 . White cells indicate empty sites.
t=50 t=l50
t=300 t=350
1 3
.\2
� - - -
' --I-
\._2 - ==> -
Fig. 7.26. Two examples of multiparticle collisions which conserve mass and
momentum. The values indicated on the arrows give the number of particles
traveling on the corresponding link. It is interesting to note that these collisions
have a non-zero global momentum, as opposed to the usual FHP model.
(note that these models are also referred to as Integer Lattice Gases). Here
we present a multiparticle model for which the algorithmic complexity
grows as the the square root of the number of particles present at a given
site. For this reason, it is expected to eventually win over a basic cellular
automata model as far as the CPU time required to obtained a given
accuracy is concerned. In addition, the rule we propose is rather simple
and allows us to adjust the viscosity with an external parameter.
z f�O) + ( 1 - z) fi < 0
decreases rapidly as the relaxation time � deviates from the zero viscosity
limit � = 1/2 (see section 3.5 ) . On the other hand, it increases when the
number of particles decreases.
To compute the collision output, we run through each of the N particles
and place them in direction i with probability P i · This will give us a
temporary particle distribution /i which then must be corrected to obtain
f� , in order to ensure exact momentum conservation.
7.3 Multiparticle fluids 301
Practically, Ji is obtained by choosing a random number s E [0, 1 [ and
comparing it with the quantity
i
so = 0 Si = L Pi
}=1
If si- 1 ::;; s < s i , the particle is placed in direction i (by construction, the
probability that this happens is indeed Pi = si - Si-d· Thus we write
N
Ji = L (si- 1 ::;; sh < si ) ( 7.49 )
h=1
where (si- 1 ::;; sh < s i ) is to be taken as a Boolean value which is 1 when
the condition is true and zero otherwise.
Since sh is a random number uniformly distributed in the interval [0, 1 [,
Ji is also a random quantity. If none of the Pi is zero, its expectation value
lS
!!}0) + ( 1 - !)� fi
N
(Ji ) = L Pi = ( 7.50 )
h=1 �
Note that when N is large enough, equation 7.49 can be computed using
the same Gaussian approximation as used in section 5.7. In this way, the
algorithmic complexity of the operation is independent of the number of
particles N.
While the distribution Ji of output particles obviously conserves the
number of particle, equation 7.50 shows that it only conserves momentum
on the average. Some particles must be pushed to other directions.
Momentum tuning is performed iteratively, according to the following
steps
• At each site where momentum is not correctly gtven by L_1 f/v1,
choose at random one lattice direction i.
• If f1 > 0 move one particle randomly to an adjacent direction.
• Accept the change if it does not worsen the momentum.
• Iterate this procedure until a particle distribution f� is obtained
which satisfies momentum conservation L-1 fj v1 = L- 1 fi v1.
From the way the particles are distributed with the probabilities Pi, we
expect that roughly .j(N ) of them are misplaced. This gives an estimate
of the number of iterations necessary to re-adjust the particles direction.
According to the above discussion it is reasonable to argue that, on
average, the quantity f').ji defined in equation 7.48 vanishes. This fact is
302 7 Other models and applications
If the average distribution (fi (r, t )) is smooth in space and time, it can be
differentiated. Then, the same derivation as that given in section 5.7 can be
repeated to show that a correct hydrodynamical behavior emerges from
our multiparticle model : equation 7.52 is equivalent to the Navier�Stokes
equation with viscosity
where c; is the sound speed whose value is model dependent (different for
hexagonal, square or cubic lattices) [233 ] .
The present muliparticle scheme is intrinsically stable because it is
based on integer numbers and probabilities. No small fluctuation will be
amplified unphysically to make the arithmetic to blow up as is the case
with the LGBK model. In principle, any value of the relaxation parameter
� can be considered. Of course if � :::; 1 /2 it is not clear whether or not a
hydrodynamical behavior is reproduced.
{
LBGK expression for such a situation [233] we can write the appropriate
local equilibrium function as f}0l
f(O) = p [ 1 + 2 Vi . U + 8 rx Uf3 (Virx VifJ - 4 U rxf3 ) ]
4 u v2 s:
7
� �
i
/0) = 12[ 1 _ 2 Vurx Uf32 ]
o
!!__
( 7.53 )
2 v
7.3 Multiparticle fluids 303
Fig. 7.27. Flow instability past a plate. The system size is 80 x 500 (although
only half of it is shown here). The parameters of the simulations are : u00 = 0. 1 5,
� = 0.55 and p = 3500 on average.
The first simulation is shown in figure 7.27. It pictures the flow instability
which develops in a fluid flow past an obstacle (von Karman street).
A second experiment (figure 7.28) shows the velocity time autocorrela
tion in an equilibrium multiparticle fluid. We consider a periodic system of
size 400 x 400, with a random initial configuration of zero average velocity
and 500 particles per site. The quantity of interest is the velocity-velocity
time correlation
C ( t ) = (v( i\ t)v(r, o ) ) - (v(r, o ) ) 2
where v(r, t ) is defined as
v(r, t) = l: vdi(r, t )
i
This quantity is known to exhibit the so-called long time tails, that
is it obeys a power law [87,234,86] instead of the fast decay exponential
predicted by a linearized Boltzmann equation [235] . The plot in figure 7.28
shows the behavior of
- 1 3 L__L __ L__L__L_-L�L_��--��
1 .5 ln (t) 6
Fig. 7.28. Long time tail in the multiparticle fluid model. Simulations are
performed on a 400 x 400 lattice, with an average density of 500 and a relaxation
time � = 1 . The slope of the straight line is - 1 .
x-velocity 0.25
Fig. 7.29. Velocity profile in a multiparticle Poiseuille flow. The plot shows
the horizontal average velocity (ux(Y )) as a function of y the vertical position
between the upper and lower boundaries. The solid line corresponds to the best
parabola fitting the data.
7.4 Modeling snow transport by wind 305
As mentioned several times in this book, the lattice gas approach makes
it quite natural to introduce complex boundary conditions. It is, for
instance, easy to deal with a dynamically changing boundary. This kind of
problem appears in particular in sedimentation processes where particles
in suspension in a fluid are transported by the flow and, under the effect
of gravity, deposit on the ground. As a result, the shape of the ground
continually changes according to its own dynamics, like for instance the
piling and toppling of a granular material. The fluid flow has of course
to self-adjust to this moving boundary. This change of flow will, in turn,
change the way deposition takes place.
Thus, we see that the problem can be quite intricate and difficult to
formulate in terms of differential equations. Moreover, in addition to
deposition, the whole process also includes erosion phenomena : the fluid
may pick up some of the particles located at the top of the deposit and
transport them downstream again according to various possible mecha
nisms such as creeping, saltation or suspension.
Such processes take place, for instance, in sand dune formation, sedi
mentation by water and snowdrift formation. Although complete formu
lation of the sedimentation problem is difficult, the lattice gas approach
give a very natural framework to propose intuitive rules accounting for the
flow motion, particle transport, erosion and deposition. One of the major
difficulties is to deal (on the same lattice) with the very different orders of
magnitude that exist between the characteristic scales governing the fluid
dynamics and the suspension. But, as usual, the hope is that, provided
a scale separation exist, the correct phenomena can be captured even
though the real scale separation is much larger than the one in the model.
In this section we present such a lattice gas model to describe snow
transport by wind and the mechanisms of deposition and erosion.
104
one discussed in section 3.5 (that is a LBGK model on a square lattice
106
with diagonals) is the natural choice to simulate wind flow. Indeed, our
aim is to simulate high Reynolds number flows such as R > for
small-scale modeling and up to for large-scale situations such as a
mountain area.
0 0
The main difference from the situation described in section 3.5 is that
now we shall include rest particles. As a consequence, the lattice directions
will be labelled from i = to i = 8, where i = refers to the rest particle
population (in particular the associated velocity vo is zero). This model is
often referred to as the D2Q9 model [233] (2 dimensions and 9 particle
populations).
The physical quantities of interest are
• the densityp = 2:::�=0 h
• and the velocity field u = � I:�=l fivi.
The lattice Boltzmann dynamics is expressed by the usual equation
flr + TVi, t + r)
fi(r, t) + z (f7q(u(r, t), p(r, t)) - f; (r, t))
=
where
1
cs2 = 3
is the speed of sound and the quantity ti are
4
to = -
9
for the rest particles,
7.4 Modeling snow transport by wind 307
Sap = 21 ( 8p ua + 8a up )
Instability appears where the magnitude I SF I = J2SapSap of this tensor is
large.
One common subgrid model is the Smagorinski model. It amounts to
correcting the relaxation time as
t = � + 3 C]magoA.2 I SF I
where Csmago > 0 is the so-called Smagorinsky constant (in practice, a
parameter of the model set, for instance, to 0 . 3 ) .
It turns out that the magnitude of the tensor Sap can be computed
locally, without taking extra derivatives, just by considering the nonequi
librium momentum tensor
rr;}l = Li V ia V ip(fi - f�q )
Then, the quantity I SF I is obtained directly as
-� + �2 + 18A.2C}mago rr;}Jrr;}J
ISF I = -- -'----
-
6 A_ 2 C}mago
--=-----=------
Unlike other numerical models [236,237] , the aim of our lattice model is
not to treat these processes (mainly saltation and suspension) separately.
By setting reasonable microscopic rules we expect to naturally recover the
macroscopic reality as the emergence of a complex behavior.
These rules governing the snow particles and their interaction with
wind particles have several components : transport mechanisms, including
gravity, erosion and, finally deposition rules. They are now described in
more detail.
7.4 Modeling snow transport by wind 309
Px = max 0,
( v 1 · (u - ug )
2 v
)
P-x = max 0,
vs (u - u )
2
( ·
v g
)
Py = max
(0, v3 · (u - ug )
v2
) P-y = max
(0, v1 · (u - ug )
v2
)
Since v 1 = -vs and v 3 = -v7 one cannot have Px and P-x simultaneously
non-zero (and similarly for Py and P-y ). On the other hand, we may have
any combination of x- and y-motion.
To select the direction of motion of a particle with this approach, two
random numbers qx and qy, uniformly distributed in the interval [0, 1],
are chosen. If qx < Px (or qx < P-x) the particle will move along the
positive x-direction (or the negative x-direction). Otherwise, if qx is larger
than both Px and P-x• no motion along the horizontal axis is permitted.
The same algorithm is applied to qy for the y-axis motion. When both
the x- and y-axis are selected, the particle moves with velocity v2 , V4, V6 or
vs, according to the sense of motion.
The average velocity (v } of the particle is then
(v } = Px Vt + P-x Vs + Py V3 + P-y V7
3 10 7 Other models and applications
Fig. 7.30. Example of particle transport in a fixed, circular velocity field. The
lattice contains 128 x 128 sites and a particle source is set in the middle of
the small circle. The particles follow their · path with a dispersion due to the
probabilistic algorithm.
V3 -v7,
Since Px and P- x are incompatible (and so are py, P-y) and v1 = -vs,
V = (vl · (uv - ug)) V...l (v3 · (uv - ug)) V3 = U -Ug
= we may as well write
( ... ) ... ...
2 2
-+
v1 v3
since and are orthogonal. This last relation shows that, although the
Fig. 7.3 1 . Four sequential stages in a snow deposit simulation. The wind is
blowing from left to right and the presence of a fence (with a ground clearance)
is shown by the vertical line segment. Light gray regions indicate solid snow (the
deposit) and white dots show the flying snow (particles in suspension). In order
to feed the system with snow, a source of snow is placed at the lower left corner
of the system.
Fig. 7.32. Three sequential stages of snow deposit in a trench. A source of snow
is placed on the left side of the system.
multiparticle snow model, a site will become a solid site only after it has
received more than a given threshold number of particles. This technique,
which takes into account some cohesion between snow particles, makes
the use of complex piling-toppling rules (such as described in section 2.2.6)
unnecessary.
Erosion. The erosion process is very complex and not yet fully understood.
Deposited particles can be picked up again by the wind, but the physical
conditions that are necessary to produce this and the interactions that are
involved are not easy to identify [238] . Some empirical laws exist that
express the erosion rate in terms of the local properties of the system, such
as the wind speed at the erosion surface. Several rules can be considered
in the lattice gas approach to reproduce qualitatively the experimental
observations. The basic erosion mechanism we use is that a solid snow
particle is ejected vertically with a random speed when the appropriate
erosion conditions are met.
312 7 Other models and applications
313
3 14 References
[14] A. Reggia, S.L. Armentrout, H.-H. Chou, and Y. Peng. Simple systems
that exhibit self-directed replication. Science, 259 : 1 282, 1 993.
[1 5] E.F. Codd. Cellular Automata. Academic Press, 1 968.
[1 6] C.G. Langton. Self-reproduction in cellular automata. Physica D,
10 : 1 35�144, 1 984.
[1 7] J. Byl. Self-reproduction in small cellular automata. Physica D,
34 :259�299, 1 989.
[1 8] C.G. Langton, C. Taylor, J.D. Farmer, and S. Rasmussen, editors.
Artificial Life II. Addison-Wesley, 1 992.
[1 9] C. G. Langton. Editor's Introduction. Artificial Life, 1 :v�viii, 1 994.
[20] M. Gardner. The fantastic combinations of John Conway's new solitaire
game life. Scientific American, 220(4) : 1 20, 1 970.
[2 1] E.R. Berlekamp, J.O. Conway, and R.K. Guy. Winning Ways for your
Mathematical Plays, volume 2. Academic Press, 1 982. Chapter 25.
[22] M. Gardner. Wheels, Life and Other Mathematical Amusements.
Freeman, 1983.
[23] K. Preston and M. Duff. Modern Cellular Automata : Theory and
Applications. Plenum Press, 1984.
[24] S. Wolfram. Theory and Application of Cellular Automata. World
Scientific, 1986.
[25] S. Wolfram. Cellular Automata and Complexity. Addison-Wesley,
Reading MA, 1 994.
[26] T. Toffoli and N. Margolus. Cellular Automata Machines : a New
Environment for Modeling. The MIT Press, 1 987.
[27] Information Mechanics Group. CAM8 : a parallel, uniform, scalable
architecture for CA Experimentation. http :/ jwww.im.lcs.mit.edujcam8.
[28] J. Hardy, Y. Pomeau, and 0. de Pazzis. Molecular Dynamics of a
classical lattice gas : Transport properties and time correlation functions.
Phys. Rev. A, 13 : 1 949�60, 1976.
[29] J.C. Maxwell. Scientific Paper II. Cambridge University Press, 1 890.
[30] H. Cornille. Exact (2+ 1 )-dimensional solutions for two discrete velocity
Boltzmann models with four independent densities. J. Phys. A,
20 :L1063�67, 1 987.
[3 1 ] J.E. Broadwell. Shock structure i n a simple discrete velocity gas. Phys.
Fluids, 7 : 1 243, 1 964.
[32] U. Frisch, B. Hasslacher, and Y. Pomeau. Lattice-gas automata for the
Navier�Stokes equation. Phys. Rev. Lett., 56 : 1 505, 1 986.
[33] S. Wolfram. Cellular automaton fluid : basic theory. J. Stat. Phys.,
45 :47 1 , 1 986.
References 315
1-
those observed on a catalytic surface. A-particles can be adsorbed on
vacant sites with some rate p. An adsorbed particle having at least one
nearest neighbor empty is desorbed with a probability p. As a
function of the value of p the stationary state is poisoned or not.
advection Transport of some quantity p due to an underlying flow
with speed u, or due to an external drift. An advection term is usually
written as u · V p. Convection is often used as a synonym of advection.
annealed disorder One possible type of disorder in statistical systems.
The annealed degrees of freedom responsible for the disorder are in
thermal equilibrium with the other degrees of freedom of the system. In
some CA models, an annealed disorder can be produced by an
independant random process at each time step.
BGK models Lattice Boltzmann models where the collision term Q
is expressed as a deviation from a local equilibrium distribution f ( 0l ,
namely Q = (f (O) - f)/�, where f is the unknown particle distribution
and � a relaxation time (which is a parameter of the model). BGK
stands for Bhatnager, Gross and Krook who first considered such a
collision term, but not specifically in the context of lattice systems.
Boltzmann equation A balance equation which expresses how the
average number of particles with a given velocity changes between
(t + dt, r + dr) and (t, r), due to inter-particle interactions and ballistic
motion ; t and r are the time and space coordinates, respectively.
CA Abbreviation for cellular automata or cellular automaton.
cellular automaton System composed of adjacent cells or sites
(usually organized as a regular lattice) which evolves in discrete time
327
328 Glossary
steps. Each cell is characterized by an internal state whose value belongs
to a finite set. The updating of these states is made in parallel according
to a local rule involving only a neighborhood of each cell.
chaos Way to describe unpredictable and apparently random
structures.
Chapman-Enskog method Expansion around the local equilibrium
distribution function f (O) , often used to solve a Boltzmann equation. One
expresses the solution as f = f (O) + ej U) + e 2 j ( 2 ) + ... where e is the
expansion parameter. By identifying the same order in e, the Boltzmann
equation yields f ( i) in terms of j U - 1 ) . However, the relation is not
invertible unless some assumptions (that are the essence of the
Chapman-Enskog scheme) are made.
collision term The right-hand side of a Boltzmann equation or of the
microdynamics associated to an LGA model. Describes the balance of
particles entering and leaving a lattice site due to collisions.
conservation law A property of a physical system in which some
quantity (such as mass, momentum or energy) is locally conserved during
the time evolution. These conservation laws should be included in the
microdynamics of a CA model because they are essential ingredients
0
governing the macroscopic behavior of any physical system.
continuity equation An equation of the form O t P + divpu =
expressing the mass (or particle number) conservation law. The quantity
p is the local density of particles and u the local velocity field.
coupled map lattice A dynamical system defined on a spatial lattice.
Each lattice site is characterized by an internal state taking a continuous
range of possible values. The dynamics of each site is defined by a
function (map) giving the evolution in terms of the local state and those
of the nearest neighbors. For descrete time, it can be viewed as a CA
with continuous states.
critical phenomena The phenomena which occur in the vicinity of a
continuous phase transition, and are characterized by very long
correlation length.
diffusion A physical process described by the equation O t P = DV2 p,
where p is the density of a diffusing substance. Microscopically, diffusion
can be viewed as a random motion of particles.
directed percolation Percolation phenomenon in which the
connectivity between the lattice sites (bonds or sites) has a preferred
direction.
Glossary 329
DLA Abbreviation of Diffusion Limited Aggregation. Model of a
physical growth process in which diffusing particles stick on an existing
cluster when they hit it. Initially, the cluster is reduced to a single seed
particle and grows as more and more particles arrive. A DLA cluster is a
fractal object whose dimension is typically 1.72 if the experiment is
conducted in a two-dimensional space.
driven diffusion Diffusion with a drift, possibly caused by an external
field, in some preferred direction.
dynamical system A sytem of equations (differential equations or
discretized equations) modeling the dynamical behavior of a physical
system.
equilibrium states States characterizing a closed system or a system
in thermal equilibrium with a heat bath.
equipartition Postulate according to which all the microstates of an
isolated system formed by a large number of particles are realized with
the same probability.
ergodicity Property of a system or process for which the
time-averages of the observables converge, in a probabilistic sense, to
their ensemble averages.
Euler equation Equation of fluid mechanics describing the evolution
of the velocity field u of a perfect fluid (without viscosity). It reads
1
a tu� + (�u · V )�u = - - VP
p
where p is the density and P the pressure.
exclusion principle A restriction which is imposed on LGA or CA
models to limit the number of particles per site and/ or lattice directions.
This ensures that the dynamics can be described with a cellular automata
rule with a given maximum number of bits. The consequence of this
exclusion principle is that the equilibrium distribution of the particle
numbers follows a Fermi-Dirac-like distribution in LGA dynamics.
FCHC model Abbreviation for Face-Centered-Hyper-Cubic model.
This is a four-dimensional lattice gas model (generalizing the FHP rule)
whose projection onto the 3D space provides a model of hydrodynamic
flows.
FHP model Abbreviation for the Frisch, Hasslacher and Pomeau
lattice gas model which was the first serious candidate to simulate
two-dimensional hydrodynamics on a hexagonal lattice.
330 Glossary
FHP-111 An extension of the FHP lattice gas model, involving rest
particles and more complicated collision rules in order to achieve lower
viscosity and higher Reynolds number flows.
field theory The theory that describes the dynamics of fields defined
on space-time and obeying partial differential equations.
Fock space A Hilbert space useful to study processes in which the
number of particles is not conserved. The creation and annihilation of
particles are described in terms of creation or annihilation operators
acting on the vectors of this Hilbert space.
fractal Mathematical object usually having a geometrical
representation and whose spatial dimension is not an integer. The
relation between the size of the object and its "mass" does not obey that
of usual geometrical objects. A DLA cluster is an example of a fractal.
front The region where some physical process occurs. Usually the
front includes the locations in space that are first affected by the
phenomena. For instance, in a reaction process between two spatially
separated reactants, the front describes the region where the reaction
takes place.
Galilean invariance Fundamental physical properties characterizing a
system that is invariant under a constant velocity shift (i.e. the system
1
obeys the same equation whether or not it moves with respect to the
observer). Galilean invariance is violated in several LGA and this is
reflected by a factor g =I= in front of the convective term uVu in the
hydrodynamical description.
geometrical phase transition Qualitative change of behavior (similar
to a thermodynamical phase transition) in a system without thermal
effects. Percolation and directed percolation are examples of geometrical
phases transitions.
Glauber dynamics Dynamical model aiming at describing the
relaxation of an Ising spin system, from an initial state to the equilibrium
state, with no conservation law.
Hamming distance A way to define a distance between two
microstates by counting the numbers of sites on which the two
microstates differ.
HPP model Abbreviation for the Hardy, de Pazzis and Pomeau
model. The first two-dimensional LGA aimed at modeling the behavior
of particles colliding on a square lattice with mass and momentum
Glossary 331
conservation. The HPP model has several physical drawbacks that have
been overcome with the FHP model.
invariant A quantity which is conserved during the evolution of a
dynamical system. Some invariants are imposed by the physical laws
(mass, momentum, energy) and others result from the model used to
describe physical situations (spurious, staggered invariants). Collisional
invariants are constant vectors in the space where the Chapman-Enskog
expansion is performed, associated to each quantity conserved by the
collision term.
Ising model Hamiltonian model describing the ferromagnetic
paramagnetic transition. Each local classical spin variables s i = ± 1
interacts with its neighbors.
isotropy the property of continuous systems to be invariant under
any rotations of the spatial coordinate system. Physical quantities defined
2D
on a lattice and obtained by an averaging procedure may or may not be
isotropic, in the continuous limit. It depends on the type of lattice and
the nature of the quantity. Second-order tensors are isotropic on a
square lattice but fourth-order tensors need a hexagonal lattice.
LB Abbreviation for lattice Boltzmann.
lattice Boltzmann model A physical model defined on a lattice where
the variables associated to each site represent an average number of
particles or the probability of the presence of a particle with a given
velocity. Lattice Boltzmann models can be derived from cellular
automata dynamics by an averaging and factorization procedure, or be
defined per se, independently of a specific realization.
lattice gas A system defined on a lattice where particles are present
and follow a given dynamics. Lattice gas automata (LGA) are a
particular class of such system where the dynamics is performed in
parallel over all the sites and can be decomposed in two stages : (i)
propagation : the particles jump to a nearest-neighbor site, according to
their direction of motion and (ii) collision : the particles entering the
same site at the same iteration interact so as to produce a new particle
distribution. HPP and FHP are well-known LGA.
lattice spacing The separation between two adjacent sites of a
regular lattice. Throughout this book, it is denoted by the symbol A.
LBGK model Abbreviation for Lattice BGK models. See BGK for a
definition.
332 Glossary
1
two velocity amplitudes (fast and slow particles) are allowed. For
instance, on a 2D square lattice, a two-speed model may include particles
of speed which jump to the nearest neighbors in one time step, while
speed- .J2 particles jump to the next nearest neighbors (along the
diagonal). Interactions conserving mass, momentum and energy are
expected between the different particle populations. Multispeed models
are useful in having a distinct mass and kinetic energy conservation laws.
p
where p is the density and P the pressure. The Navier-Stokes equation
expresses the local momentum conservation in the fluid and, as opposed
to the Euler equation, includes the dissipative effects with a viscosity
term vV2 u. Together with the continuity equation, this is the fundamental
equation of fluid dynamics.
334 Glossary
rate equation The evolution equation for the local particles densities
in a reaction-diffusion systems in which the reaction terms are treated in
mean-field approximation.
Glossary 335
reaction-diffusion systems Systems made of one or several species of
particles which diffuse and react among themselves to produce some new
species.
renormalization group Method which systematically implements
some form of coarse-graining to extract the properties of the large-scale
phenomena, in physical systems where many scales are important.
Reynolds number In fluid dynamics, a dimensionless quantity defined
as the ratio (uwd)jv, where u00 is the generic speed of the flow (e.g speed
far away from obstacles), d the typical size characterizing the system
(characteristic size of an obstacle) and v the kinematic viscosity of the
fluid. For the same geometry, flows with identical Reynolds numbers are
similar in the sense that they differ by scaling factors.
scaling hypothesis A hypothesis concerning the analytical properties
of the thermodynamic potentials and the correlation functions in a
problem invariant under a change of scale.
scaling law Relations among the critical exponents describing the
power law behaviors of physical quantities in systems invariant under a
change of scale.
self-organized criticality Concept aimed at describing a class of
dynamical systems which naturally drive themselves to a state where
interesting physics occurs at all scales.
spatially extended systems Physical systems involving many spatial
degrees of freedom and which, usually, have a rich dynamics and show
up complex behaviors. Coupled map lattices and cellular automata
provides a way to model spatially extended systems.
spin Internal degree of freedom associated to particles in order to
describe their magnetic state. A widely used case is the one of classical
Ising spins. To each particle, one associates an "arrow" which is allowed
to take only two different orientations, up or down.
telegraphist equation The partial differential equation
atP + (D j c2 )a� pDV2 p, where p is the quantity which is considered, D a
=
C0-0 2 ).
Index
337
338 Index
crack, 261 exclusion principle, 38, 67, 76, 140, 209
creeping, 305, 308 exponent
critical critical, 3 1 , 233
behavior, 249 mean-field, 197
exponent, 3 1 , 233, 252
phenomena, 2 1 9 FCHC model, 1 1 2
point, 250 ferrofluid, 293
current of particle, 139 FHP model, 6, 75, 77, 78, 1 1 8, 276
FHP-III, 77, 107, 109
D2Q9 model, 306 Fick's law, 139
damped waves field theory, 220
one dimension, 265 fluctuations, 86, 1 1 1, 1 22, 200
two dimensions, 264 fluid,multiparticle method, 298
decay process, 222 Fock space, 220
dendrite, 203 force
deposition, 305 external, 286
detailed balance, 106 gravity, 1 36, 296
diffusion long range, 292
constant, 1 39, 146, 1 50, 2 1 2, 2 1 8 short range, 292
front, 163 forest fire model, 31, 62
microdynamics, 140 fractal, 3 1 , 1 63, 201
one dimension, 70, 146 fracture, 261
rule, 1 39, 141, 1 76 Fredkin rule, 8
three dimensions, 149 front
two dimensions, 1 57 reaction, 193
with traps, 1 77 width, 194
diffusion-limited deposition, 1 7 1
dimer, 1 7 7 Galilean invariance, 97, 102, 103, 1 1 6, 1 24, 1 29,
directed percolation, 29 1 30
dispersion relation game of life, 3, 1 8
diffusion, 1 5 6 Gaussian approximation, 2 1 3 , 2 1 5 , 3 0 1 , 3 1 0
random walk, 6 9 geometrical phase transition, 30
wave model, 266 Glauber dynamics, 295
dissipation, 276 glider, 3, 19
DLA, 166, 1 77, 200 granular, 305
driven diffusion, 1 77, 255 gravity, 1 36, 296, 309
dynamical system, 2 1 Green-Kubo
formalism, 105
eddies, 1 10 relation, 149
Ehrenfest model, 175 Greenberg-Hasting model, 181, 228
Einstein, 28
Einstein relation, 14 7 H-theorem, 106
elastic collision, 60 Hamming distance, 255
energy, 96, 1 1 5 heterogeneous catalysis, 242
conservation, 276 hexagonal lattice, 75, 94, 295, 296
equipartition, 1 3 5 mapping on a square lattice, 287
internal, 1 1 5 homogeneous, 1 8 8
epidemic model, 228 HPP
equilibrium states, 232 collision rules, 39
equipartition, 96, 1 1 3, 1 3 5 model, 5, 6, 38, 1 3 5
ergodicity, 3 6 two-speed, 121
erosion, 305
Euler equation, 94, 96 immiscible flows, 7, 280
excitable media, xi, 179 inertia, 141
excluded surface, 1 72 inhomogeneous systems, 193
Index 339