Sie sind auf Seite 1von 12

International Journal of Modern Physics C, Vol. 8, No.

4 (1997) 685696
c World Scientific Publishing Company

DIGITAL PHYSICS SIMULATION OF LID-DRIVEN CAVITY FLOW

CHRISTOPHER M. TEIXEIRA
Exa Corporation, 450 Bedford Street
Lexington, Massachusetts 02173, USA
E-mail : teixeira@exa.com

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

Received 30 October 1996


Revised 26 February 1997
Digital Physics is a new extension of the lattice-gas concept for the simulation of fluid flow
which removes disadvantages that prevented practical application of the original method.
The extensions are summarized. Simulation results for a three-speed model demonstrate
the absence of artifacts and significant reduction in viscosity. Also, simulation results
for 2D and 3D lid-driven cavities for a range of Reynolds numbers and geometries are
compared with experiment, CFD and the lattice-Boltzmann BGK method. Accurate
results are obtained with the new method.
Keywords: Extended Lattice Gas; Simulation Validations; 2D and 3D Cavity Flow;
Digital Physics.

1. Introduction
Lattice-Gas Automata (LGA) was introduced over a decade ago as a potential
new method for efficiently simulating fluid dynamics. It was shown that this
discrete, particle-based method, which utilized simplified microdynamics incorporating important physical attributes (exact conservation laws, symmetry), solved
macroscopically equations that looked NavierStokes-like in both two1 and three
dimensions.2,3 The method generated a great deal of interest4 because of several
important advantages over conventional computational fluid dynamics (CFD) techniques. Specifically, the lattice-gas method is unconditionally stable, efficient computationally and inherently parallel, and permits complex boundaries to be easily
handled.
Unfortunately, disadvantages were also demonstrated. Nonhydrodynamic artifacts in the momentum equation were found.13 Large noise-to-signal ratios5 and
inherently large transport coefficients6 both made it difficult, if not impossible, to
simulate high Reynolds number transient flows. Finally, the equilibrium had Fermi
Dirac instead of MaxwellBoltzmann form, a consequence of the boolean nature of
the method, which was physically incorrect for ideal gases. While it was quickly
shown that the artifacts could be removed by the addition of dynamical transition
rates in the collision process,7 the remaining problems raised serious doubts about
the practical usefulness of the method as a fluid simulator.
685

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

686

C. M. Teixeira

To avoid these problems, attention turned to solving the Boltzmann equation directly.8 A variation of this technique, known as the lattice-Boltzmann BGK
model,9,10 corrected most of the problems with the lattice-gas approach. However ,
in doing so the method lost some of its computational efficiency since floating point
instead of boolean operations were used but, more importantly, the collision process employed no longer obeyed an H-Theorem. Thus, the method was subject to
instability like any other numerical method. This problem has only begun to be
addressed11 and is even more pervasive when thermodynamics is added.12
Digital Physics is an extension of the original lattice-gas concept which maintains the advantages of that method while removing the disadvantages. The method
recovers NavierStokes hydrodynamics, reduces fluctuations by allowing integer
state occupations, uses a multi-step collision process that has an H-Theorem and
achieves MaxwellBoltzmann equilibrium, and allows for variation and reduction of
the transport coefficients.
Elsewhere, the theoretical elements that allow these key features to be achieved
are presented.13 Here we present the highlights without proof. Simulation results
are then presented confirming the removal of the artifacts from the momentum
equation and the behavior of the viscosity. We then apply the method to a standard
benchmark problem, lid-driven cavity flow in both two and three dimensions, and
compare the Digital Physics results with experiments, direct numerical simulations
of the fluid equations and lattice-Boltzmann BGK results, where appropriate.
2. Description of Method
As with a lattice gas, particles reside in a world where space, velocity, time and
energy are discretized. The nature of the discretization of velocity and energy
is dictated by the underlying lattice which must have appropriate symmetry.1,3
Particle occupations in this discrete phase space are integers in the range zero
to Nmax , denoted Nji (x, t), where j is the energy of a state moving in discrete
direction i = (1, 2, . . . , dj ). The velocity of such a state is cji . Particles, with unit
mass, have energy j = 12 (cji cji ). The standard face-centered hypercube (FCHC)
lattice used for three-dimensional simulations3 can be generalized to accomodate all
non-negative energies, j, with symmetry requirements for each energy satisfied.14
The microdynamics consists of two steps, advection and collision, which may be
described as follows,
Nji (x + cji , t + 1) = Nji (x, t) + Cji {(Nji )} ,

(1)

where Cji {(Nji )} denotes the collision operator. The collision operator applied
in the Digital Physics method is a sequential series of valid binary collisions that
transfers an integer number of particles from one pair of states to the other pair.
A valid collision exactly conserves the mass, momentum and energy at the node.
The number of particles to be transferred for each collision is a function of the
occupations of the states involved and a pair of transition rates. These rates must

Digital Physics Simulation of Lid-Driven Cavity Flow

687

be positive and themselves functions of quantities that are collisionally invariant.


Conditions under which a local and global H-Theorem may be demonstrated for
this collision operator are presented in Chen et al.13 The equilibrium can then be
eq
calculated from Cji {(Nji
)} = 0 and has the MaxwellBoltzmann form:
!#
"
c2ji
(eq)
(1)
,
(2)
Nji = Rj
exp + cji +
2

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

where the coefficients , and are associated with the conservation of mass,
momentum and energy respectively, and may be determined by enforcing their
definitions:
X eq
X
X1
1
eq
eq
=
Nji , u =
cji Nji
, Uint + u2 =
c2ji Nji
.
(3)
2
2
ji
ji
ji
The energy dependent factors, Rj , are the transition rates.a
With the equilibrium determined and the free parameters Rj initially set to
unity, the Euler components of the stress tensor, (0) , and energy flux, Q(0) , may
be calculated14


X
2
1 2
eq
(0)
=
cji cji Nji = I
Uint + u (1 g) + guu ,
(4)
D
2
ji
Q(0) =

X
ji

1
eq
cji c2ji Nji
=
2



D+2
1
g
Uint + g 0 u2 u ,
D
2

(5)

where D is the dimension of the underlying lattice. These results differ from their
continuum counterparts due to the presence of lattice artifacts.2,7,14 These artifacts
may be removed (g = 1, g 0 = 1) if the second and third order moments of the
microscopic energy are set to14
2
X  1 2 eq
D + 2 Uint
dj
c2j Nj =
,
(6)
2
D

j
X
j


dj

1 2
c
2 j

3
Njeq =

3
(D + 4)(D + 2) Uint
.
2
D
2

(7)

In the above, Njeq denotes the velocity-independent


of the equilibrium, Eq. (2),
 part
P
P
eq
eq
1 2
so that j dj Nj = , the mass, and j dj 2 cj Nj = Uint , the internal energy.
Since these higher order moments are not conserved in the collision process, they
can be set by adjusting the equilibrium appropriately. The introduction of two
rates, Rj and Rj0, will allow both artifacts to be removed, which requires at least
a four-speed model.14 Satisfying Eq. (6) requires one rate, Rj , and a three-speed
model which removes the artifacts from the Euler stress tensor only.
a Generally, the rates may be functions of state direction as well as energy.13

688

C. M. Teixeira

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

A consequence of removing the artifacts from the Euler terms using the above
equilibrium form is that the dissipative terms are artifact-free through second order
2
Uint and using an ideal gas equation
in the velocity.13 Setting the pressure P = D
b
of state, P = T , defines the temperature, T . With these substitutions the hydrodynamic equations that the Digital Physics fluid satisfies for mass, velocity and
temperature are, respectively,



+ u = ( u) ,
(8)
t



+ u u = P + + O(u3 ) ,
(9)
t


D

+ u T = P ( u) + (T ) + : u + O(u4 ) ,
(10)
2
t

u
u
2
=
+
( u)I ,
x
x
D
which reproduces all details of continuum hydrodynamics to the indicated order in
velocity.
In the above equations, represents the dynamic viscosity while is the thermal
conductivity. Explicit expressions for these transport coefficients can be calculated
given the collision operator. Chen et al.13 demonstrates a multi-step collision process that introduces multiple relaxation time scales which allows alteration of the
transport coefficients while still satisfying an H-Theorem. For example, if two time
scales, and , are introduced then the expressions for the coefficients are13




1
1
1 D+2
1
=

T
and
=

T .
(11)

2
2

where

With a single relaxation parameter, = , the lattice-BGK result9,10 is recovered. However, here this result has been accomplished using collisional processes
which ensures exact conservation of mass, momentum and energy, and positivity of
state occupations. To complete the specification of the fluid, the sound speed, cs ,

is14 cs = T where = (D + 2)/D is the ratio of specific heats.


The simulations in the next section use a three-speed model, j = (0, 1, 2), with
8-bit state occupationsc and a single relaxation parameter in the collision operator.
The explicit functional form of the equilibrium for this system may be calculated,13
 


j
1
eq
eq
2
Nji = Nj (, T ) exp
exp (cji u) + O(u4 ) ,
(12)
T
2T
where14
N0eq =

(3T 2 3T + 1); N1eq = 2T (2 3T ); N2eq = T (3T 1) ,


d0
d1
d2

b The Boltzmann constant for the simulated gas is set to unity, k


B(lat) 1.
c A m-bit representation means the maximum occupation, N
m
max is Nmax = 2 1.

(13)

Digital Physics Simulation of Lid-Driven Cavity Flow

689

and d1 = d2 = 24 and d0 , the number of stopped particles, is arbitrary. We choose


d0 = 6. The range of temperature is determined by positivity and finiteness of the
state occupations. For this system, the allowable ranged is 13 < T < 23 .
3. Simulation Results

To demonstrate the removal of the Galilean-invariance breaking artifact, g, and


measure the viscosity of the three-speed model, the decay of an advecting momentum shear wave was analyzed for a series of velocity amplitudes, |u| = (0.14, 0.28),
and temperatures, T = (0.42, 0.50, 0.58) as a function of the relaxation parameter, . Results are presented in Fig. 1. When g = 1 the momentum equation
becomes artifact free, Eq. (4). Figure 1(a) demonstrates that g = 1.00 0.02 for
the range of parameters tested. For the T = 0.42 case we find that the error in g is
increased when the velocity is increased indicating the influence of fourth order and
higher velocity effects. Next, we compare the measured value of kinematic viscosity,
= /, to the theoretical result from Eq. (11), = T 1 12 , in Fig. 1(b). For
< 1.8 excellent agreement with theory is demonstrated. In Fig. 1(c), we highlight
the region 1.8 2.0 and now compare results using 8-bit and 16-bit states
for T = 0.42. For 8-bit states, the error increases as is increased. The value of
viscosity saturates for > 1.95 to ' 0.008 and ' 0.011 for velocity magnitudes 0.14 and 0.28 respectively. However, when 16-bit states are used, measured
viscosity agrees well with theory to = 1.99 ( ' 0.0013), the largest value tried.
As with lattice-Boltzmann BGK, the method requires < 2.0 for stability.11 With
finite precision states, there is a max < 2.0 above which the viscosity reducing
process violates the H-Theorem13 due to inaccuracy in achieving the maximum entropy state at each time step. The value of max increases as precision is increased.
For reasons not entirely understood, when a two-step collisional process is used the

a)

g artifact vs lambda

b)

viscosity vs lambda

c)

1.04

viscosity vs lambda at T=.42

0.025

1.03

T=.42
T=.42
T=.50
T=.58

1.02

u=.14
u=.28
u=.14
u=.14

T=.42
T=.42
T=.50
T=.58

0.25

T=.42 Theory
T=.50 Theory
T=.58 Theory

0.20

viscosity

1.01
1.00
0.99
0.98

u=.14
u=.28
u=.14
u=.14

0.15
0.10
0.05

u=.14
u=.28
u=.14
u=.28

0.020

8 bits
8 bits
16 bits
16 bits

Theory

viscosity

0.30

measured value of g

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

3.1. Basics

0.015

0.010

0.005

0.97
0.00
0.96
1.00

1.25

1.50
lambda

1.75

2.00

1.00

1.25

1.50
lambda

1.75

2.00

0.000
1.80

1.85

1.90
1.95
lambda

2.00

Fig. 1. Measured values of Galilean-invariance artifact, g, and viscosity, , as a function of temperature, T , velocity magnitude, u, and relaxation parameter, . (a) g versus , (b) versus ,
< 1.8, (c) versus , 1.8 > > 2.0 for 8-bit and 16-bit states.
d More generally, allowable range of temperature is 2 < T < 4 .
D+2
D+2

C. M. Teixeira

viscosity reducing step ceases to function above max causing saturation of the
viscosity rather than instability.
The impact of using integer states on the noise level was investigated by measuring the mean state occupation (the signal) and the standard deviation about this
mean (the noise) in a simulation volume of size 16 16 cells seeded quiescently
and with periodic boundary conditions. Statistics are accumulated for each of the
54 states for 4000 time steps resulting in a sample size of just over 1106 . Since the
flow is isotropic, we also accumulate over all states with a given speed. The ratio
of the standard deviation of the mean to the mean occupation is then computed
for each of the three speeds. The test case was run using m = (1, 2, 4, 6, 8, 12, 16)
number of bits to represent the state occupations while keeping the normalized density constant, /Nmax = 10, with two different temperatures, T = (0.42, 0.56). The
relaxation parameter, , had the value = 1 for all cases. Figure 2 shows the noise
levels for these six sets of data as a function of number of bits, m. For m > 4 we
find that the noise decreases linearly with the inverse of the occupation value (note
line with slope 1). The standard deviation is < 1 and relatively constant in
this range while the mean occupation increases linearly with the number of bits.
This demonstrates the ability of the collision process to drive the distribution to
the closest integer about the real-valued mean quantities. However, for m 4 the
data flattens out where the noise-to-signal ratio is order unity or larger indicating
that the hole states play a significant role2 there. This clearly demonstrates the
significant reduction in noise possible with integer states in comparison with the
single-bit lattice gas.
Noise-to-Signal Ratio vs. Number of Bits
5

log (sigma/N)

-5

-10

T=.42 Speed 0
T=.42 Speed 1
T=.42 Speed 2

-15

T=.56 Speed 0
T=.56 Speed 1
T=.56 Speed 2

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

690

Slope = -1
-20
0

8
10
Number of Bits

12

14

16

18

Fig. 2. Noise-to-signal ratio for each of the three speeds for temperatures, T = (0.42, 0.56) as a
function of the number of bits, m, used to represent the states. For m > 4, noise decreases in
inverse proportion to the number of bits, as illustrated by the line with unit slope. For m 4, the
noise does not decrease as significantly.

3.2. Two-dimensional lid-driven cavity


Two-dimensional flow in a lid-driven cavity is a very attractive problem computationally due to its geometric simplicity and complex flow structure. Simulations

Digital Physics Simulation of Lid-Driven Cavity Flow

691

Fig. 3. 2D Streamlines for Re = 3200. A clockwise rotating primary vortex along with three
counter-rotating secondary vortices are visible. The lid is moving from left to right.
a)

b)

X Velocity(y) Along Geometric Centre

1.0

0.5

0.9

0.4
f

0.2

0.7

Reynolds #
b

0.5
0.4
c

0.2
e
g

Reynolds #
-0.0
a
b
c
d
e
f
g

-0.1
-0.2

DNS - Ghia
100
400
1000
3200
5000

0.3

0.1

0.1

10
100
400
1000
2000
3200
5000

Vy/U

a
b
c
d
e
f
g

0.6

0.0
-0.50

Y Velocity(x) Along Geometric Centre

0.3

0.8

y/L

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

have been performed for a series of Reynolds numbers,e Re = 10, 100, 400, 1000,
2000, 3200 and 5000, on a 256 256 lattice.f Lid velocities, U , and relaxation
parameters, , in the range 0.039 < U < 0.30 and 1.00 < < 1.82 were used
to achieve these values. Bounce-back boundary conditions were used at the
three walls.
Figure 3 presents the simulated streamlines for Re = 3200 and demonstrates the
complex flow structure that results. A primary clockwise-rotating vortex generates
counter-rotating secondary eddies in the bottom two corners of the cavity and one
in the top left corner (the lid is moving from left to right). The location and size of
these vortices are strong functions of the Reynolds number. In Table 1, normalized
locations of the primary and secondary vortices are presented and compared with
the direct numerical simulations (DNS) of Ghia et al.15 and the lattice-BGK results
of Hou et al.16 The current results are within 1% of DNS for the primary vortex
location and differ by no more than half a lattice cell for the secondary vortices.

-0.3

a
b
c

DNS - Ghia
100
400
1000
3200
5000

-0.4

10
100
400
1000
2000
3200
5000

-0.5

d
e

-0.6
-0.25

0.00

0.25
Ux/U

0.50

0.75

1.00

0.0

0.2

0.4

0.6

0.8

g
1.0

x/L

Fig. 4. 2D Cavity Velocity Profiles as a function of Reynolds Number. (a) X component of velocity
as a function of y through centre; (b) Y component of velocity as a function of x through centre.
Velocity is normalized by the lid velocity and space by the length of the cavity. Comparisons are
with Ghia et al.15
e Re = U L/ where U is the velocity of the lid and L the length of the cavity.
f Except for Re = 10 where a 64 64 lattice was used.

692

C. M. Teixeira
Table 1. Vortex Locations for 2D Cavity.

Primary Vortex
x
y

Re
10
100

400

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

1000

2000

3200

5000

Lower Left
x
y

Lower Right
x
y

0.5156

0.7656

0.0059

0.0117

0.9609

0.0391

0.6172

0.7344

0.0313

0.0391

0.9453

0.0625

0.6196

0.7373

0.0392

0.0353

0.9451

0.0627

0.6152

0.7383

0.0275

0.0352

0.9493

0.0587

0.5547

0.6055

0.0508

0.0469

0.8906

0.1250

0.5608

0.6078

0.0549

0.0510

0.8902

0.1255

0.5588

0.6053

0.0537

0.0425

0.8906

0.1172

Upper Left
x
y

0.5313

0.5625

0.0859

0.0781

0.8594

0.1094

0.5333

0.5647

0.0902

0.0784

0.8667

0.1137

0.5352

0.5664

0.0859

0.0741

0.8691

0.1128

0.5255

0.5490

0.0902

0.1059

0.8471

0.0980

0.5234

0.5484

0.0850

0.1010

0.8477

0.0977

0.0207

0.8699

0.5165

0.5469

0.0859

0.1094

0.8125

0.0859

0.0547

0.8984

0.5195

0.5430

0.0828

0.1152

0.8320

0.0898

0.0488

0.8941

0.5117

0.5352

0.0703

0.1367

0.8086

0.0742

0.0625

0.9102

0.5176

0.5373

0.0784

0.1373

0.8078

0.0745

0.0667

0.9059

0.5156

0.5352

0.0723

0.1391

0.8086

0.0742

0.0586

0.9102

al.15 ;

al.16 ;

Note: a, Ghia et
b, Hou et
c, present work. All lengths normalized by the
length of the cavity. Origin is at bottom left of cavity.

The lattice BGK results show comparable accuracy. The appearance of the vortex
in the upper left corner between Re = 1000 and 2000 is consistent with previous
findings.17
Figure 4 compares the normalized horizontal and vertical components of velocity
along the geometric centre lines for all simulated Reynolds numbers with DNS. The
agreement is good. At higher Reynolds numbers, the boundary layers are restricted
to a thin region close to the wall and the profiles become linear indicating a large
central region of uniform rotation where near inviscid conditions exist.
3.3. Three-dimensional lid-driven cavity
Flow structure in the cavity becomes more complex when a third dimension is
added, as was shown by Koseff and Street.18 Their experiments demonstrated that
the presence of side-walls produced a three-dimensional structure that significantly
altered the primary flow in the central plane, behavior that two-dimensional simulations simply could not capture. For laminar flow (Re < 6000) they found two
sources of three-dimensional structure. Examining a plane parallel to the downstream wall, corner eddies were caused at the juncture of the side-walls and the

693

ground. Secondary vortices caused as well by centrifugal forces along the downstream eddy separation surface were found along the span. These came to be known
as Taylor-Gortler-Like (TGL) vortices in reference to their curvature-induced origins. The number and location of these vortices were a function of Reynolds number,
Spanwise-Aspect-Ratio,g and time.1820
Simulations were performedh for cavities of SAR 1:1 (a cube), lattice size 128
128 128, and SAR 3:1, 128 128 384, at Re = 3200 and compared with experimental results. The normalized primary flow velocity profiles in the plane midway
between the side-walls are shown for both geometries in Figs. 5(a) and 5(b). Mean
profiles of simulation and experiment have been averaged over comparable periods
of nondimensional time,i . The agreement is good. Koseff and Street19 reported
that the nondimensionalized height of the downstream secondary eddy grows from
0.28 0.02 at SAR 1:1 to 0.35 0.02 for SAR 3:1 at Re = 3300. We measured
values of 0.29 and 0.34 for the SAR 1:1 and 3:1 cases respectively, which agrees with
experiment.
a)

Centre Profiles 3D SAR 1:1 Cavity

b)

Centre Profiles 3D SAR 3:1 Cavity

x/L
0.00
1.0

0.25

0.50

x/L
0.75

1.00
1.00

0.00
1.0

Re # 3200
Digital Physics
Experiment

1.00
1.00

0.75

0.0

0.50

0.0

0.50

-0.5

0.25

-0.5

0.25

0.0
U /U
x o

0.5

1.0

0.00

-1.0
-1.0

U y/U o

0.5

o
y

U /U

0.75

y/L

0.75

-0.5

0.50

Digital Physics
Experiment

0.5

-1.0
-1.0

0.25
Re # 3200

y/L

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

Digital Physics Simulation of Lid-Driven Cavity Flow

0.00
-0.5

0.0
U /U
x o

0.5

1.0

Fig. 5. 3D Cavity Central Plane Velocity Profiles for Re = 3200. (a) Spanwise-aspect-ratio (SAR)
1:1 comparison with experiment, Prasad et al.21 ; (b) SAR 3:1 comparison with experiment, Koseff
et al.20

Next, we examine the flow structure in a plane parallel to the downstream wall
about three-quarters of the way down the length (x/L = 0.77) and at nondimensional time = 125 (about 53500 timesteps). Figure 6 shows projected streamlines
for the cube. Two strong corner eddies along with two relatively weak pairs of
vortices are readily visible along the bottom. Figure 7 shows streamlines for the
SAR 3:1 case. Besides the two corner eddies, we now count eight pairs of vortices
of varying strength, some of which appear to be as large as the corner eddies.
g SAR ratio of span to length.
h Again, bounce-back was used at the walls to enforce zero velocity there.
i = time U/L, where U is the velocity of the lid and L is the cavity length.

694

C. M. Teixeira

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

Fig. 6. Projected Streamlines for SAR 1:1 3D cavity at x/L = 0.77 and = 125. Two corner
eddies and two pairs of Taylor-G
ortler-Like vortices are visible along the span.

Fig. 7. Projected Streamlines for SAR 3:1 3D cavity at x/L = 0.77 and = 125. Two corner
eddies and eight pairs of Taylor-G
ortler-Like vortices are visible along the span.

Experiments indicate that, in the mean, there are in fact two of these TGL vortices
for the SAR 1:1 cavity21 and eight for SAR 3:118 at Re ' 3200. However, the
experiments also indicate that the location and number of these vortices vary with
time and that symmetry is not necessarily maintained. We find that the number of
TGL vortices varies between two and three for the cube and from seven to nine for
the SAR 3:1 case. The time-dependent behavior of these simulations along with an
examination of a possible correlation between the TGL structure on the bottom and
on the upstream wall, as first suggested by Hou,22 where the curvature of the lower
upstream eddy also triggers secondary vortices, is ongoing and will be reported in
a later publication.
Finally, a note about computational performance. All of the cavity simulations
presented here were performed using Exas PowerFLOWTM software on a SUN
UltraSparc-2 workstation. The number of cell updates per second was about 90000.j
Work is currently under way to reduce the number of needed states from 54 to
34 which would boost the update rate to about 140000 cells per second. Succi
et al.23 reported about 330000 cell updates per second with their uniform 3D LB
code. While currently moderately slower, the Digital Physics method offers the
considerable advantages of unconditional stability and reduced storage requirements
(integers versus single- or double-precision floating point) in comparison with the
lattice-Boltzmann method.
j With a 200 MHz clock rate.

Digital Physics Simulation of Lid-Driven Cavity Flow

695

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

4. Summary
Digital Physics is an extended lattice-gas method that removes the disadvantages
while maintaining the stability of that method. We have demonstrated the absence
of Galilean-invariance breaking artifacts and shown that the minimum viscosity is
about min 0.01 for an 8-bit model and at least a factor of ten smaller for a
16-bit model. The maximum cell Reynolds number ( Ux
) for the 8-bit model is
then about 40, significantly larger than with the standard lattice gas.6 Simulation
of flow in both two-dimensional (over a range of Reynolds numbers) and threedimensional (for two separate geometries) cavities agree well with other numerical
simulations, and experiments, respectively, thus demonstrating the ability of the
method to accurately and efficiently reproduce hydrodynamic behavior. Different
CFD methods produce results of varying quality for the 3D cavities, none of which
have been deemed acceptable when compared with experiments.24 Thus, this case
is especially potent for demonstrating the utility of alternative methods. Analysis
of time dependence in the 3D cavities is ongoing and will be presented in the future.
Work is under way to extend the method to other applications including heat
transfer, complex fluids, and high Reynolds number flows. Recently, results for high
Reynolds number flow (Re 106 ) around a car-like shape have been presented and
are very promising.25
Acknowledgments
The author wishes to thank David Freed for use of his multi-bit Digital Physics code
and Joe Tompkins and David Hill for assistance in data acquistion for the cavity
simulations.
References
1. U. Frisch, B. Hasslacher, and Y. Pomeau, Phys. Rev. Lett. 56, 1505 (1986).
2. U. Frisch, D. dHumi`eres, B. Hasslacher, P. Lallemand, Y. Pomeau, and J-P. Rivet,
Complex Systems 1, 649 (1987).
3. S. Wolfram, J. Stat. Phys. 45, 471 (1986).
4. G. D. Doolen (ed.), Lattice Gas Methods: Theory, Applications and Hardware, Physica,
D47 (North Holland, Amsterdam, 1991).
5. J. Dahlburg, D. Montgomery, and G. Doolen, Phys. Rev. A36, 2471 (1987).
6. M. Henon, Complex Systems 1, 763 (1987).
7. K. Molvig, P. Donis, J. Myczkowski, and G. Vichniac, in Discrete Kinetic Theory,
Lattice Gas Dynamics and Foundations of Hydrodynamics, ed. R. Monaco (World
Scientific, Singapore, 1989), p. 409.
8. G. McNamara and G. Zanetti, Phys. Rev. Lett. 61, 2332 (1988).
9. Y. Qian, D. dHumi`eres, and P. Lallemand, Europhys. Lett. 17, 479 (1992).
10. H. Chen, S. Chen, and W. Matthaeus, Phys. Rev. A45, 5339 (1992).
11. J. D. Sterling and S. Chen, J. Comp. Phys. 123, 196 (1996).
12. Y. Chen, private communication (1996).
13. H. Chen, C. Teixeira, and K. Molvig, Int. J. Mod. Phys. C4, 675 (1997).
14. C. Teixeira, Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, 1992.
15. U. Ghia, K. N. Ghia, and C. Y. Shin, J. Comp. Phys. 48, 387 (1982).

696

C. M. Teixeira

Int. J. Mod. Phys. C 1997.08:685-696. Downloaded from www.worldscientific.com


by 201.141.62.54 on 03/18/13. For personal use only.

16. S. Hou, Q. Zou, S. Chen, G. Doolen, and A. C. Cogley, J. Comp. Phys. 118, 329
(1995).
17. A. S. Benjamin and V. E. Denny, J. Comp. Phys. 33, 340 (1979).
18. J. R. Koseff and R. L. Street, J. Fluids Eng. 106, 21 (1984).
19. J. R. Koseff and R. L. Street, J. Fluids Eng. 106, 385 (1984).
20. J. R. Koseff and R. L. Street, J. Fluids Eng. 106, 390 (1984).
21. A. K. Prasad and J. R. Koseff, Phys. Fluids A1, 208 (1989).
22. S. Hou, Ph.D. Thesis, Kansas State University, Manhattan, 1995.
23. S. Succi, G. Amati, and R. Benzi, J. Stat. Phys. 81, 1/2 (1995).
24. M. Deville, T-H. Le, and Y. Morchoisne (eds.), Notes on Numerical Fluid Mechanics
36 (Vieweg, Wiesbaden, 1992).
25. A. Anagnost, A. Alajbegovic, H. Chen, D. Hill, C. Teixeira, and K. Molvig, Society of
Automotive Engineers, Warrendale, PA, SAE Paper 970139, 1997.

Das könnte Ihnen auch gefallen