Sie sind auf Seite 1von 12

Complex Systems 4 (1990) 139-150

Numerical Simulations of Hydrodynamics with a Pair


Interaction Automaton in Two Dimensions

D.A. Wolf-Gladrow
Alfred-Wegener-Institute for Polar and Marine Research,
D-2850 Bremerhaven, Federal Republic of Germany

R. N asilowski
UniverWit Oldenburg, FB Physik,
D-2900 Oldenburg, Federal Republic of Germany

A. Vogeler
Alfred- Wegener-Institute for Polar and Marine Research,
D-2850 Bremerhaven, Federal Republic of Germany

Abstract. We perform numerical experiments with a novel cellular-


automaton fluid model. The model has simple pair-interaction rules
and works in arbitrarily many dimensions; here we consider the two-
dimensional case. By observing the smoothing-out of the velocity
profile of a shear flow and the relaxation of a shear wave, we mea-
sure the kinematic shear viscosity and obtain a value of :::::: 0.5. To
demonstrate the fluid-like behavior of our automaton model, we also
simulate a flow past an obstacle (von Karman vortex street).

1. Introduction
A first attempt to simulate fluid dynamics by cellular automata has been
made by Hardy, de Pazzis, and Pomeau (HPP) [6]. They invented a fully
discrete two -dimensional lattice gas model on a square lattice with built-in
conservation laws for mass (particle number) and momentum. The HPP
model behaves qualitatively like a fluid, but cannot quantitatively simulate
interesting fluid dynamical phenomena because of anisotropies in the hydro-
dynamics of the lattice gas. The possibility of quantitative fluid -dynamic
simulation, however, has recently become available with the hexagonal Iat -
tice gas model by Frisch, Hasslacher, and Pomeau (FHP) [4] in two dimen-
sions, and the pseudo-4D face-centered hypercubic (FCHC) lattice gas model
by d'Humieres and Lallemand [3] in three dimensions. The hydrodynamics
of these models, although anisotropic too, can be shown by symmetry ar-
guments to be quasi-isotropic in the sense that the hydrodynamics in the

@ 1990 Complex Systems Publications, Inc.


140 D.A. Wolf-Gladrow, R. Nasilowski, and A. Vogeler

low-velocity incompressible limit becomes isotropic and is equivalent to the


usual incompressible Navier-Stokes dynamics [5].
Particularly the two-dimensional FHP model and variants thereof have
been successfully applied to problems like porous media flow [14], solids in
suspension [9], surface tension [1], immiscible fluids [15], autocatalytic re-
actions and flame p ropagation [2] . Three-dimensional simulations using the
FCHC model, however, are much more difficult , not only because of the 3D
space, but also because the rules of the FCHC model are very hard to formu-
late as an algo rithm which should be relatively simple and computationally
effect ive at the same time [7,16] .
These difficulties with the FCHC model in 3D, and also the awkward
hexagonal geometry of the FHP model in 2D, can possibly be overcome by a
new model that has recently been proposed by Ralf Nasi lowsk i [11,12]. The
new model works in arbitrarily many dimensions on orthogonal lattices and
has simple, determi nistic rules with pair int er acti ons between cells. Nev-
ertheless, the hydrodynamics of this lattice gas model can be shown to be
equiva lent to the usual (isotropic) Euler equati ons in the low-velocity in-
compressible nondissipat ive limit, so it is in a cer tain sense quasi-isot rop ic.
T heoretical result s for the practically interest ing dissipat ion transport coef-
ficients , part icu la rly the shear viscosity, have not been obtained yet .
In this article, we present first results of com puter experiment s which
have b een perfor med wit h the two-dimensional version of the new au tomaton
model. In t he first experiment, we observe how a velocity profi le of a shear
flow gets smoothe d out, whence we can determine the kinematic viscosity.
In t he second experiment, we try to simu late a von Ka rman vortex street by
let t ing the lat ti ce gas flow past an obstacle.

2. T h e p air in t er a ction aut o maton


Our lattice gas model, in the case of two dimensions, is defined as follows. We
x
use the "alt ern at ing" lattice shown in figure 1, whose node po ints = (Xl,X2)
have coordinates

if t E Z.
if t E Zo
1,2) (2.1)

at ti mes t = 0, 1,2, . . ., where Z. an d Zo is t he set of even and odd integr al


numbers, respectively. Locat ed at each lattice p oint are 4 cells (compar e
figure 2), where each of the cells corresponds t o one of t he 4 poss ible discrete
v
velocit y vect ors = (VI, V2) wit h compo nent s

Vj = ±1 = 1, 2) .
(j
The state of a cell (t, x,v) is specified by 3 bit s (n o, nl, n 2)' T he value of t he
bit no in dicates whether the cell is occ upied (no = 1) or empty (no = 0);
we call no the m ass of the cell. T he two remaining bit s (nI, n2) are always
Numerical Simulations of Hydrodynamics 141

o D
¥- 7«-
,0 D
>rc X

**10 X D
D

-X-
D

X: Subgrid 1
0: Subgrid 2

Figure 1: Lattice structure of our model. Particle positions at even


and odd times are shown as white and black circles, respectively. The
arrows indicate the 4 pos sible particle velocities .

b a
Cb = (-1,1 ) C, = (1,1)

Co = (-1 ,- 1) Cd = (1,-1)
c d
Figure 2: Attempt to illust rate the cellular structure of our mo del.
T he left pict ure shows all the local cells (indicat ed by the squares )
belonging t o one of t he lattice points (indicated by the dot in the
middle); the right pict ure shows one of the cells in more detail. Each
circle corresponds to a bit in the computer.
142 D.A. Wolf-Gladrow, R. Nasilowski, and A. Vogeler

zero for an empty cell, but may take arbitrary Boo lean values (0 or 1) for
an occupied cell. We define, in a somewhat unconventional manner, the
momentum m= (ml,m2) ofthe cell (t,x, v) by
(j =1,2) .
Note that our definition of "momentum" differs from the usual one: in our
model, momentum f- mass x velocity; instead, a particle with velocity v =
(VI,V2) may have a momentum of m
= (VI, V2 )' (Vl, O) , (0, V2), or (0,0) ,
dep ending on t he values of the bits nl and n 2, which may be viewed as
int ernal degrees of freedom of t he particle.
The dynamics of our model is given by

nJ(t + 1, x + v, V) = n~(t, x, v) (J =0,1,2),


where nJ and n'.J denote the bit values immediately before respectively after
the interactions which may take place between particles when they meet at
times t = 0,1,2, . . . at the lattice points. Not e that in equation (2) an d
elsewhere in this text, we are using uppercase and lowercase index letters
to indica te different ranges of allowed value s: J E {a, 1, 2} b ut j E {1,2}.
T he interactions, n J -> n '.J , are in st ant aneous and local (between t he four
cells at a lattice point). Equat ion (2) states t hat the particles move free ly
v
according to t he ir velocity vectors during the unit t ime intervals between
the int eger-val ued inst ant s of time.
Our interaction rule is composed of pair in teract ions of cells and cons ists
of two successive steps (see figure 3):

(2.2)

In the first step, n} -> n}, the interacting pairs are formed by cells whose
v
velocity vectors differ only in the first component: = (±1, V2) with the same
value of V2 for bot h partners ("horizont al pair" ). In t he secon d step, n} -> n},
v
the partners differ only in the second velocity component : = (V I , ± 1) with
t he same VI ("vertical pair").
Our pair interaction rules conserve the to tal mass and momentum of each
interacting pair. For the interaction between two horizont al cells (here: cells
a and b, compare figure 2) they can be derived from the following table (see
also figure 4) :

ao bo al bl a' b'0 a'I b'I a'2 b'2


°
°11 ° °° °° °1 ° °°1 °°1 bbb
1

1
1
1
1
2
2
a2
a2
(2.3)

1
1
1
° °1 11 11 °1 ° bb
1 1 1
2
2
a2
a2

1 1 ° 1
° 1
° 1 °1 b 2
2
a2
a2
whe re ao, o are the mass bits and aI, a2, bl , ~ are the momentum bits
b
of cells a and b before interaction; the bits after int eract ion are indicated
Numerical Sim ulations of Hydrodynamics 143

b) a)
(1 , 1 ) (0,1) (0,1) (1 ,1 )

(1 ,0 ) (1, 0 )
( 0, 0) (0 ,0)

c) d)
(0 ,0) (0 ,0)
(1,0) ~-----... (1, 0 )

(1 ,1) ( 0,1) ( 0,1) (1 ,1 )

Figure 3: Subsequent formation of horizontal and vertical pairs of in-


teracting cells in the collision process. The "process" is instantaneous
(it takes no time in the model) and local (it does not involve any cells
located at different lattice points). The squares represent the 4 local
cells, their velocity vectors are indicated by the arrows.

by an apostrophe ('). In cases not listed in the table, we set aJ = aj and


b~ = bJ . After interaction in the horizontal direction between cells a and b
respectively c and d (n} -+ nJ), the four cells will interact in the vertical
direction (cells a and c respectively b and d) in the same manner (nJ -+ n}) .
The dynamics of our automaton is now comp lete ly defined.
The pairwise mass and momentum conservation follows from the proper-
ties

a~ + b~ (2.4)
a~ - b~ (2.5)
a~ + b~ (2.6)

of table (2.3). Furthermore, we not e that not only two-particle int eract ions
(3rd through 6th line of the table), but also one-particle interactions (1st and
2nd line) may occur. In a one-particle interaction, the "self-colliding" particle
changes its velocity while retaining its momentum. T his and all the other
144 D.A. Wolf-Gladrow, R. Nasilowski, and A. Vogeler

...
.20

...
. 10

-
r---.

>
>-
~

<;:»
.0'

- .0 5

-.t o

-.15

-.20

-.25
10 20 30 .to 50 80 70 80 110 100 110 120 130

y
Figure 4: The pair int eraction rules, illustrated for a "hori zontal" in-
teraction. A legenda of the diagrams is given at the bottom, where
also the conser vation laws are indi cated. V, V+, nJ, and nj: denote
t he velocity vector s resp. state bit s of t he par tn er cells. When only
one cell contains a par ticle and th e ot her is empty, that particle can
"spontaneously" change its velocity (jump to th e partner cell) wit h-
out changing its momentum (rule 1). When both cells are occupied
by particles, th ese particles necessarily retain t heir velocities, but a
transfer of momentum from one particle to t he ot her is possible (r ules
2 th rough 4).

tran sitions listed in (2.3) are only possible because of our unconventional
defini tion of momentum by equation (2).
Since we want to simulate hydrodynamic phenomena, t he observables of
interest in our model are not t he states of ind ivid ual cells, but rat he r the
ma croscopic densities p and q of mass and momen tum , given by
Numerical Simulation s of Hydro dynamics 145

~ \~mj(t,x,V)) = ~ \~ Vjnj(t, x,V) )


(j =1 ,2 ), (2.7)
where the sum extends over all four po ssible valu es of v with Vj = ± 1 and
the average (-) is taken over many lattice points in som e neighborhood of x.
By applying to our model the usual Gibbs formalism of statistical mechanics
and then making a Chapmau-Enskog type expansion, in a similar way as this
has been done for the conventional cellu lar-automaton fluids [5], one obtains
the hydrodynamic equations [12]

V · 11 o
a11 ~ ~
at + (11 . \7)11 + \74> o (2.8)

with

Eq uations (8), which have the usual form of fluid-dynamic E uler equations,
are valid for the la ttice gas only in the low-velocity, incompressible, nondis-
sipative limi t , for a mean mass density of p = 1/2. When these limiting
con ditions are not satisfied, the automaton's hydrodynamics is describ ed by
an isotropic equations whose form is considerably more complicated than (8) .
Unfortunately, the theoretical prediction of dissipative effects, particu-
larly t he shear viscosity, which is of great practical interest, is very difficult,
and no results are avai lab le yet. We can, however, "measure" the viscosity
by computer experiments.

3. Measurement of the shear viscosity


3 .1 The smoothing-out of a di scontinu ity in v elo city
A simple method to measure the shear viscosity of lattice gases has been
described by Lim [10]. T he init ial set -up consists of two ad jacent streams
movi ng at the same velocity U but in opposite directions:

- U for y < 0
{ +U (3.1)
for y > 0
o (3.2)
(att =O ) (3.3)
where 11 = : (Ul,U2)' The analytical solution of t he incompressible Navier-
Stokes equations with this initial condition is known:
Y
Ul = U erf rrr: > v = 0,
y4vt
146 D.A. Wolf-Gladrow, R. Nasilowski, an d A. Vogeler

. 2>

.2 0

. IS

. t0

.0'
/"""'.,
+-
'"
>-
'-../
> - .Q'

-.1 0

-."

- .20 f-------==~=------

o '0 20 30 ..0 50 SO 70 80 10 100 110 120 130

y
Figure 5: Measuremen t of the shear viscosity /I : Ul (y , t) over y (after
6000 time steps) : data point s (0) and fitted error fun~tion.

whe re erf is the error function , and /I is the the kinematic shear viscosity coef-
ficient . Linearizing the er ror function for small values (~ 1) of t he argument,
we have

The viscosity can be calculated by linear regression applied to the observed


velocity profile Ul(t,y) for sufficiently small y at different times t . T he size
of our computational domain is 2048 x 2048 un its, and we use macro-cells for
coarse-graining (local averaging over x) of size 64 x 16. T he vertical velocity
component v and the density deviation from 1/2 remain small during t he
simulat ion : after t = 6000 timesteps we observe U2 = 0.000 ± 0.014 and
p = 0.500 ± 0.015. The x-averaged mean values U(t,y) = (Ul (t , X, y))x
and the corresponding standard deviations are calculated from 32 macro-cell
values along the x-direction. The results of the linear fit y = 0 (see figure 5)
Numerical Simulations of Hydrodynamics 147

are shown in the following table:

t 2000 4000 6000


v 0.49 0.52 0.48

The value of the kinematic shear viscosity is similar to than that of the
FHP-2 model at the same density (FHP-l: v = 1.2; FHP-2: v = 0.68; FHP-
3: v = 0.075 [5]). Only recently Rothman [13] has given (non-local) collision
rules for FHP which result in even negative viscosity.

3.2 Shearwave r elaxation measurements


As shown by d'Humieres and Lallemand [3] the relaxation of an initial shear-
wave on a square lattice with periodic boundary conditions can be used to
measure the shear viscosity v , the bulk viscosity T) and the speed of sound c.
The initial shearwave consists of a longitudinal and a transversal mode and
has the form

(3.4)

111 is the longitudinal mode, 11t is the transversal mode and k is the wavevector
of the shearwave. By solving a linearized Navier-Stokes equation with an
relaxing shearwav~ (4) one gets the relaxation of the initial velocity modes
and a density perturbation op coupled to the longitudinal mode:

k2(V + T))
(
~
11(f, t) = UI cos(wt) exp( 2 t)

+ 11t exp( -k 2v /2t) ) cos(k. i')


2(v
op(r,t) ((PU I/C) sin(wt) exp( _ k sin(k.i') (3.5)
2+T))t))

with w = ck and C(~:C'1) r


~ 1. A typical set of the relaxation curves
is shown in figure 6. To obtain the relaxation curves the transversal and
longitudinal velocities and the density have been averaged perpendicular to
the wavevector at each time step. Then the interesting component of the
initial wave has been extracted by a Fourier transformation as a function
of time. The viscosities v and T) and the speed of sound c were calculated
by least squares fits of the above equations (5) to the measured relaxation
curves. We performed our measurements on a lattice of 512 x 512 nodes with
wavelengths between 64 and 128 nodes at a density per cell of p = 0.5. And
we let the shearwave propagate in x- and y-direction. This is easily attained
by exchanging the succession of interacting pairs in the lattice gas. Within
the uncertainty of our measurements we obtained the same values for both
directions. For sufficiently small mean velocities the results are independent
of wavelength and mean velocity. We have obtained v = 0.46 ± 0.02 for the
kinematic shear viscosity, which is in good agreement to the previous method.
148 D.A . Wolf-Gladrow, R. Nasilowski, and A. Vogeler

..
. ---
;
~ - -
~
-
- - ..
:"",:;::::::::::::::::::::::::::
,
:::::: :::.. .. .. ..... . .... .. .. ... . .. ..

~~~~~~~~~~~~~~~i~i~! 111 ii111 iii i


;~F:::::::::::: ::: ::::::::: : ::: :
11'· ·· · ........·· __·· · ·· ·· ···· · ·· · ·· · · · · · ·
:
Il l· ····· · ·· ·· · ·· ·· ·· ··· ·· ··· ········ ·

j j \HHHH :H:llHH HH j:~ ~ [ ~ ~ ~ j ~ .~


.. ..
. . .. ...
. ...... .. ...
...

... ........ . .. - -
n
. .. . ... .. . .. .. .. . . .. . .. . .. . . . . .. . . ... . .. .. ..
. . . .. . .. ... ... . .. . . .. _ .

Figure 6: Shearwave relaxation measurements: A typical set of mea-


sured relaxation curves.

The bulk viscosity "l seemed to have its value below 0.02, so we werr- unable
to determine it . For th e sound velocity c we have measured c = 0.58±0.02, a
confirmation of the predicted theoretical value of c = ~ 0.577 for p = 0.5 VI
in Nasilowski [12]. The results of the measurements are summarized in the
following table:

u p x direction v "l c
0.05 0.5 128 y 0.46 < 0.02 0.58
0.05 0.5 128 x 0.48 < 0.02 0.58
0.07 0.5 128 y 0.46 < 0.02 0.58
0.07 0.5 128 x 0.46 < 0.02 0.58
0.05 0.5 64 y 0.47 < 0.02 0.59
0.10 0.5 64 y 0.4(; < 0.02 0.58

4. Von Karman vortex street

In the literat ur e, the von Karman vortex street serves as a classical example
for t he comp lexity of flow patterns which may occur in systems governed by
the Navier-Stokes equations from a phenomenological point of view [8]. We
performed the vortex street simu lations on a domain of size 6400 x 3200,
with periodic boundary conditions in the x-direction and free slip at the y-
boundaries . The width of the obstacle is 400 units. With a speed of u = 0.1
and a kinematic shear viscosity of 0.5 one get s a Reyno lds number of 80.
F igure 7 shows the flow pattern at t = 80000.
N umerical Simu lations of Hydrodynamics 149

Figure 7: Von Karm an vortex street: The flow pattern after 80000
time steps (the mean flow has been subtracted) on a lattice of 6400 X
3200 nodes with a velocity of u = 0.1. With an obst acle of length 400
nodes the Reynolds number is 80.

5. Conclusions
Our simul ations show t hat the new cellular-automaton model indeed shows
the expect ed fluid-dynamical behavior . Moreover, we have obtained by
measurement a numerical va lue of the shear viscosity coefficient and the
sou n d speed. Whether or not dissipative effects are (at least approximately)
isotropic in our lattice gas , is still an open qu estion.

Acknowledgments
We are grateful to Mr . Charilaos Kougias for teaching us important program-
ming techniques for lattice gas automata.

References
[1] D. Burge ss, F. Hayot, and W.F. Saam, "Model for surfac e \«nsion in lattice-
gas-hydrodynamics ," Phys. Rev. , A38(7) (1988) 3589--35'J2.

[2] P. Clavin, P. Lallemand, Y. Pomeau , and G. Searby, "Simulation of free


boundaries in flow systems by lattice-gas models," J. Fl uid Mech. , 188
(1988) 437-464.

[3] D. d'Humieres and P. Lallemand, "Lattice gas automata for fluid mechanics,"
Physica, 140A (1986) 326-335.
150 D.A. Wolf-Gladrow , R. Nasilowski, and A. Vogeler

[4] U. Frisch, B . Hasslacher, and Y. Pomeau, "Lattice-gas automata for Navier-


Stokes equations, " Phys . R ev . Lett., 56 (1986) 1505.

[5] U. Frisch, D. d'Humieres, B. Hasslacher, P. Lallemand, Y. Pomeau, and


J .-P. Rivet, "Lattice gas hydrodyna mics in two and three dimensions," Com-
plex Systems, 1 (1987 ) 649-707.
[6] J. Hardy, O. de Pazzis, and Y. Pom eau , "Molecular dynamics of a classical
lattice gas : Transport properties an d time correlation functions," Phy s. Rev.,
A 13 (1976) 1949.

[7] M. Henon, "Isome tric collision rules for t he four-dimensional FC HC lattice


gas ," Complex Sys tems, 1 (1987) 475-494.

[8] L.P. Kadanoff, "On two levels," Phy sics Today, 39 (1986) 7-9 .

[9] A.J .C. Ladd and M.E. Colvin, "Application oflattice-gas cellular automata
to the Brownian motion of solids in suspension," Phys. Rev. Lett., 60 ( 11 )
(1988) 975-978.

[10] H.A. Lim, "Lattice gas automata of fluid dynamics for un st eady flow," Com -
plex Sys tems, 2 (1988) 45- 58.
[11] R . Nasilowski, "An arb itrary-dimension al cellula r-aut omaton fluid model
with simple rules", to appear in P roceedings in Dissipati ve St ruct ures in
Transport Processes and Comb ustion (Interdisciplinary Seminar, Bielefeld
1989), Springer-Verlag, Heidelberg.

[12] R . Nasilowski, "A cellular-automaton fluid model wit h simp le rules in arb i-
t rarily many dime nsions ," sub mitted to J . St at . Phys. (1990).

[13] D.H. Rothman, "Negative-viscosity lat ti ce gases," J. Stat. Phys., 56(3 / 4)


(1989) 517-524.

[14] D.H. Rothman, "Cellular-aut omat on fluid s: A model for flow in porous me-
dia," Geophysics, 53(4) (1988) 509-518.

[15] D.H. Rothman and J.M . Keller , "Immisci ble cellula r-automaton fluids," J.
St at . Phys ., 52 (1988) 1119-1127.
[16] S. Succi, P. Santangelo , and R. Benzi, "High resolution lattice gas simulatio n
of two-dimensional turbulence," Phys. Rev. Lett., 60 (1988) 2738-2740 .

Das könnte Ihnen auch gefallen