4
VOLUME 78, NUMBER 20 PHYSICAL REVIEW LETTERS 19 MAY 1997 Kolmogorov Turbulence in Low-Temperature Superflows C. Nore, 1 M. Abid, 2 and M. E. Brachet 1 1 Laboratoire de Physique Statistique de l’Ecole Normale Supérieure, associé au CNRS et aux Universités Paris 6 et 7, 24 Rue Lhomond, 75231 Paris Cedex 05, France 2 Institut de Recherche sur les Phénomènes Hors Equilibre, UMR CNRS et Université d’Aix-Marseille I, service 252, Centre St-Jérôme, 13397 Marseille Cedex 20, France (Received 11 December 1996) Low-temperature decaying superfluid turbulence is studied using the nonlinear Schrödinger equation in the geometry of the Taylor-Green ( TG) vortex flow with resolutions up to 512 3 . The rate of (irreversible) kinetic energy transfer in the superfluid TG vortex is found to be comparable to that of the viscous TG vortex. At the moment of maximum dissipation, the energy spectrum of the superflow has an inertial range compatible with Kolmogorov’s scaling. Physical-space visualizations show that the vorticity dynamics of the superflow is similar to that of the viscous flow, including vortex reconnection. The implications to experiments in low-temperature helium are discussed. [S0031-9007(97)03046-9] PACS numbers: 67.40.Vs, 47.37. + q, 67.40.Hf Superfluid flows are described mathematically in terms of Landau’s two-fluid model [1]. When both normal fluid and superfluid vortices are present, their interaction, called “mutual friction,” must be taken into account as pioneered by Schwarz [2]. At temperatures low enough for the nor- mal fluid to be negligible (in practice below T 1 K for helium at normal pressure), an alternative mathematical description is given by the nonlinear Schrödinger equa- tion (NLSE), sometimes also called the Gross-Pitaevskii equation [3,4]. The NLSE reads t c sicy p 2 jdsc2 jc j 2 c1j 2 = 2 c d . (1) The complex wave field c is related to the superflow’s density r and velocity y by Madelung’s transformation r jc j 2 , ry j sicjy p 2 dsc≠ j c2 c≠ j c d, where j is the so-called “coherence length” and c is the velocity of sound (when the mean density r 0 1 [5]). The superflow is irrotational, except near the nodal lines of c which are known to follow Eulerian dynamics [6,7]. These topological defects correspond to the superfluid vortices that appear naturally, with the correct velocity circulation, in this model [8]. The basic goal of the present Letter is to qualify the degree of analogy between turbulence in low- temperature superfluids and incompressible viscous fluids. We will do this by comparing numerical simulations of NLSE with existing numerical simulations of the Navier-Stokes equations, in particular the Taylor-Green (TG) vortex [9]. The TG vortex is the solution of the Navier-Stokes equations with initial velocity field v TG s sinsxd coss yd cossz d, 2 cossxd sins yd cossz d,0d. This flow is well documented in the literature [10–12]. It admits symmetries that are used to speed up computations: rota- tion by p about the axis sx z p y2d, s y z p y 2d, and sx y p y2d, and reflection symmetry with respect to the planes x 0, p , y 0, p , z 0, p . The velocity is parallel to these planes which form the sides of the impermeable box which confines the flow. The TG flow is related to an experimentally studied swirling flow [13–15]. The relation between the experimental flow and the TG vortex is a similarity in overall geometry [13]: a shear layer between two counterrotating eddies. The TG vortex, however, is periodic with free-slip boundaries while the experimental flow is contained inside a tank between two counterrotating disks. We now show how to construct a vortex array whose NLSE dynamics mimics the vortex dynamics of the large scale flow v TG . The first step of our method is based on a global Clebsch representation of v TG and the second step minimizes the emission of acoustic waves [16]. The Clebsch potentials lsx, y, z d cossxd p 2j cossz dj, msx, y, z d coss yd p 2j cossz dj sgnscossz dd (where sgn gives the sign of its argument) correspond to the TG flow in the sense that =v TG =l 3 =m. The complex field c c , corresponding to the large scale TG flow cir- culation, is given by c c sx, y, z d sc 4 sl, mdd fg d y4g with g d 2 p 2ysp cjd (fg is the integer part of a real) and c 4 sl, md c e sl2 1y p 2, mdc e sl, m2 1y p 2 d 3c e sl1 1y p 2, mdc e sl, m1 1y p 2 d , where c e sl, md sl1 i md tanhs p l 2 1m 2 y p 2jdy p l 2 1m 2 . The second step of our procedure consists of integrating to convergence the advective real Ginzburg-Landau equa- tion (ARGLE): t c scy p 2jdsc2 jc j 2 c1j 2 = 2 c d 2 i v TG ? =c 2 fsv TG d 2 y2 p 2 cjgc , (2) with initial data c c c . This amounts to minimizing the functional F scy p 2 jd Z d 3 $ x f2jc j 2 1 jc j 4 y2 1j 2 j=c 2 si v TG y p 2 cjdc j 2 g . (3) It is shown in [17] that replacing v TG by a constant vector yields as a minimum of (3) a boosted straight vortex line that is an exact radiationless solution of (1). 3896 0031-9007y 97y 78(20) y3896(4)$10.00 © 1997 The American Physical Society

Kolmogorov Turbulence in Low-Temperature Superflows

  • Upload
    m-e

  • View
    213

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Kolmogorov Turbulence in Low-Temperature Superflows

VOLUME 78, NUMBER 20 P H Y S I C A L R E V I E W L E T T E R S 19 MAY 1997

6 et 7,

252,

38

Kolmogorov Turbulence in Low-Temperature Superflows

C. Nore,1 M. Abid,2 and M. E. Brachet1

1Laboratoire de Physique Statistique de l’Ecole Normale Supérieure, associé au CNRS et aux Universités Paris24 Rue Lhomond, 75231 Paris Cedex 05, France

2Institut de Recherche sur les Phénomènes Hors Equilibre, UMR CNRS et Université d’Aix-Marseille I, serviceCentre St-Jérôme, 13397 Marseille Cedex 20, France

(Received 11 December 1996)

Low-temperature decaying superfluid turbulence is studied using the nonlinear Schrödinger equationin the geometry of the Taylor-Green (TG) vortex flow with resolutions up to5123. The rate of(irreversible) kinetic energy transfer in the superfluid TG vortex is found to be comparable to that of theviscous TG vortex. At the moment of maximum dissipation, the energy spectrum of the superflow hasan inertial range compatible with Kolmogorov’s scaling. Physical-space visualizations show that thevorticity dynamics of the superflow is similar to that of the viscous flow, including vortex reconnection.The implications to experiments in low-temperature helium are discussed. [S0031-9007(97)03046-9]

PACS numbers: 67.40.Vs, 47.37.+q, 67.40.Hf

mdlleedr

icak

o].uity

fywidseee

tso

o

w

nd:hes

nk

seea

G

r-

tingua-

ht).

Superfluid flows are described mathematically in terof Landau’s two-fluid model [1]. When both normal fluiand superfluid vortices are present, their interaction, ca“mutual friction,” must be taken into account as pioneerby Schwarz [2]. At temperatures low enough for the nomal fluid to be negligible (in practice belowT ­ 1 K forhelium at normal pressure), an alternative mathematdescription is given by the nonlinear Schrödinger eqution (NLSE), sometimes also called the Gross-Pitaevsequation [3,4]. The NLSE reads

≠tc ­ sicyp

2 jd sc 2 jcj2c 1 j2=2cd . (1)The complex wave fieldc is related to the superflow’sdensity r and velocityy by Madelung’s transformationr ­ jcj2, ryj ­ sicjy

p2 d sc≠jc 2 c≠jcd, wherej is

the so-called “coherence length” andc is the velocityof sound (when the mean densityr0 ­ 1 [5]). Thesuperflow is irrotational, except near the nodal linesc which are known to follow Eulerian dynamics [6,7These topological defects correspond to the superflvortices that appear naturally, with the correct veloccirculation, in this model [8].

The basic goal of the present Letter is to qualithe degree of analogy between turbulence in lotemperature superfluids and incompressible viscous fluWe will do this by comparing numerical simulationof NLSE with existing numerical simulations of thNavier-Stokes equations, in particular the Taylor-Gre(TG) vortex [9]. The TG vortex is the solution of thNavier-Stokes equations with initial velocity fieldvTG ­sss sinsxd cossyd cosszd, 2 cossxd sinsyd cosszd, 0ddd. This flowis well documented in the literature [10–12]. It admisymmetries that are used to speed up computations: rtion by p about the axissx ­ z ­ py2d, sy ­ z ­ py2d, and sx ­ y ­ py2d, and reflection symmetry withrespect to the planesx ­ 0, p , y ­ 0, p, z ­ 0, p. Thevelocity is parallel to these planes which form the sidesthe impermeable boxwhich confines the flow. The TGflow is related to an experimentally studied swirling flo

96 0031-9007y97y78(20)y3896(4)$10.00

s

d

-

al-ii

f

id

-s.

n

ta-

f

[13–15]. The relation between the experimental flow athe TG vortex is a similarity in overall geometry [13]a shear layer between two counterrotating eddies. TTG vortex, however, is periodic with free-slip boundariewhile the experimental flow is contained inside a tabetween two counterrotating disks.

We now show how to construct a vortex array whoNLSE dynamics mimics the vortex dynamics of the largscale flowvTG . The first step of our method is based onglobal Clebsch representation ofvTG and the second stepminimizes the emission of acoustic waves [16].

The Clebsch potentialslsx, y, zd ­ cossxdp

2j cosszdj,msx, y, zd ­ cossyd

p2j cosszdj sgnssscosszdddd (where sgn

gives the sign of its argument) correspond to the Tflow in the sense that=vTG ­ =l 3 =m. The complexfield cc, corresponding to the large scale TG flow ciculation, is given byccsx, y, zd ­ sssc4sl, mddddfgdy4g withgd ­ 2

p2yspcjd (f g is the integer part of a real) and

c4sl, md ­ cesl 2 1yp

2, mdcesl, m 2 1yp

2 d3 cesl 1 1y

p2, mdcesl, m 1 1y

p2 d ,

where cesl, md ­ sl 1 imd tanhsp

l2 1 m2yp

2jdypl2 1 m2.The second step of our procedure consists of integra

to convergence the advective real Ginzburg-Landau eqtion (ARGLE):

≠tc ­ scyp

2jd sc 2 jcj2c 1 j2=2cd 2 ivTG ? =c

2 fsvTGd2y2p

2 cjgc , (2)

with initial datac ­ cc. This amounts to minimizing thefunctional

F ­ scyp

2 jdZ

d3 $x f2jc j2 1 jcj4y2

1 j2j=c 2 sivTGyp

2 cjdcj2g . (3)

It is shown in [17] that replacingvTG by a constantvector yields as a minimum of (3) a boosted straigvortex line that is an exact radiationless solution of (1

© 1997 The American Physical Society

Page 2: Kolmogorov Turbulence in Low-Temperature Superflows

VOLUME 78, NUMBER 20 P H Y S I C A L R E V I E W L E T T E R S 19 MAY 1997

ty

-lr

Tio

e

bar

t

r

n

d

ld

avevior

-

x

datei-

f

ae-

m

the

e

m,

a

l

Minimizing (3) with the space-dependentvTG appliesto the vortex lines a local Galilean boost with velocivTG. The TG symmetries can be used to expandcsx, y,z, td ­

PNy2m­0

PNy2n­0

PNy2p­0 csm, n, p, td cosmx cosny 3

cospz where N is the resolution andcsm, n, p, td ­ 0,unlessm, n, p are either all even or all odd integers. Implementing this expansion in a pseudospectral code yiea saving of a factor 64 in computational time and memosize when compared to general Fourier expansions.ARGLE converged periodic vortex array obtained in thmanner is displayed on Fig. 1. Note that the radiuscurvature of the vortex lines is large compared to thradius.

The total energy of the vortex array, conservedNLSE dynamics, can be decomposed into three pEtot ­ s1y2pd3

Rd3x sEkin 1 Eint 1 Eqd, with kinetic

energy Ekin ­12 ryjyj, internal energy Eint ­ sc2y

2d sr 2 1d2, and quantum energyEq ­ c2j2s≠jp

r d2.Each of these parts can be defined as the integral ofsquare of a field, for example,Ekin ­ 1

2 spryjd2. UsingParseval’s theorem, the angle-averaged kinetic enespectrum is defined as

Ekinskd ­12

Zk2 sinu du df

3

Ç1

s2pd3

Zd3r eirjkj

pr yj

Ç2,

which satisfies Ekin ­ 1ys2pd3R

d3x Ekin ­R`

0 dk 3

Ekinskd. The angle-average is performed by summiover shells in Fourier space. A mode (m, n, p) belongs tothe shellk ­ f

pm2 1 n2 1 p2 1 1y2g. Ekin is further

decomposed into compressibleEckin and incompressible

Eikin parts, using

pr yj ­ spr yjdc 1 spr yjdi with

=spr yjdi ­ 0. This simple decomposition has the a

FIG. 1. Three-dimensional visualization of the vector fie= 3 sr $yd for the Taylor-Green flow at timet ­ 0 withcoherence lengthj ­ 0.1ys8

p2 d, sound velocityc ­ 2, and

resolutionN ­ 512 in the impermeable boxf0, pg 3 f0, pg 3f0, pg. The visualization is obtained by drawing the30 000vectors of highest norm.

dsyhesf

ir

yts

he

gy

g

-

vantage over the more conventional one,p

r yj ­p

r syjdc 1p

r syjdi , of not involving a mixedcompressible-incompressible energy spectrum. We hchecked that both decompositions give the same behafor Ei

kinskd in the runs presented below.An exact solution of NLSE, describing a 2D axisym

metric vortex, is given bycvortsrd ­p

rsrd expsimwd,m ­ 61, wheresr , wd are polar coordinates. The vorteprofile

prsrd , r as r ! 0 and

prsrd ­ 1 1 Osr22d

for r ! `. It can be computed numerically using mappeChebychev polynomials expansions and an approprifunctional [17]. The corresponding velocity field is azmuthal and is given byysrd ­

p2 cjyr . Using the ex-

pansion forp

rsrd, the 2D angle-averaged spectrum op

r yj can then be computed with the formulaEvortkin skd ­

sc2j2y2pkd fR`

0 dr J0skrd≠rp

r g2, whereJ0 is the zerothorder Bessel function. It is shown in [17] that, when3D isolated vortex line is almost straight, the 3D anglaveraged spectrum of the line is given byEline

kin skd ­sly2pdEvort

kin skd, wherel is the length of the line.Indeed, the incompressible kinetic energy spectru

Eikinskd of the ARGLE converged vortex array of Fig. 1

displayed on Fig. 2(a) is well represented, at largek,by an isolated line spectrumEline

kin skd with total vortexlength given byly2p ­ 175. In contrast, the small wavenumber region cannot be represented byEline

kin skd. Thisstems from the average separation distance betweenvortex lines in Fig. 1. Calling this distancedbump ,k21

bump ­ 1y16, the wave number range between thlarge-scale wave numberk ­ 2 and the characteristicseparation wave numberkbump can be explained by

FIG. 2. Plot of the incompressible kinetic energy spectruEi

kinskd. The bottom curve (a) (circles) corresponds to timet ­ 0 (same conditions as in Fig. 1). The spectrum ofsingle axisymmetric 2D vortex multiplied bysly2pd ­ 175 isshown as the bottom solid line. The top curve (b) (plusses)corresponds to timet ­ 5.5. A least-square fit over the interva2 # k # 16 with a power lawEi

kinskd ­ Ak2n givesn ­ 1.70(top solid line).

3897

Page 3: Kolmogorov Turbulence in Low-Temperature Superflows

VOLUME 78, NUMBER 20 P H Y S I C A L R E V I E W L E T T E R S 19 MAY 1997

c

l

vhm

re

s

-ea-n-

u

ktie

nuo

n

at

ion

is

t.in

tex

so

be

nof

interference effects. Because of constructive interferenthe energy spectrum atk ­ 2 has a value close toits corresponding value in TG viscous flow (name0.125), which is much above the value ofEline

kin sk ­ 2d.In contrast, for2 , k # kbump , destructive interferencedecreasesEi

kinskd below Elinekin skd. This spectral situation

can be understood by analogy with the reproductiona grey scale picture using black dots. The small wanumber spectrum of the reproduction will be that of toriginal picture, while the large wave number spectruwill be that of the individual dots. Here the dots avortex lines and the picture isvTG .

The evolution in time via NLSE (1) of the incompresible kinetic energy is shown in Fig. 3 for various valueof j, the resolutionN being adjusted to maintain accuracy. The main quantitative result of this Letter is the rmarkable agreement of the energy dissipation r2dEi

kinydt with the corresponding data in the incompressible viscous TG flow. Both the mometmax , 5 10 of maximum energy dissipation (the inflection point of Fig. 3) and its valueestmaxd , 1022 at thatmoment are in quantitative agreement with the viscodata [10,18]. Furthermore, bothtmax andestmaxd dependweakly on j. This is remarkably similar to the weadependence of the corresponding viscous dissipain the limit of small viscosity. This weak dependencis considered a hallmark of numerical evidence forKolmogorov regime in decaying turbulence [18].

Another important quantity studied in viscous decayiturbulence is the scaling of the kinetic energy spectrduring time evolution and, especially, at the moment

FIG. 3. Total incompressible kinetic energyEikin plotted

versus time for j ­ 0.1ys2p

2 d and resolution N ­ 128( long-dash line); j ­ 0.1ys4

p2 d, N ­ 256 (dash); j ­

0.1ys6.25p

2 d, N ­ 400 (dot); and j ­ 0.1ys8p

2 d, N ­512 (solid line). All runs are realized withc ­ 2. Theevolution of the total vortex filament length divided by2p(crosses) for theN ­ 512 run is also shown (scale given othe righty axis).

3898

e,

y

ofe

e

-s

-te

t

s

on

a

gmf

maximum energy dissipation, where ak25y3 range can beobserved [10]. Figure 2(b) shows the energy spectrumt ­ 5.5. A least-square fit over the interval2 # k # 16with a power lawEi

kinskd ­ Ak2n givesn ­ 1.70 (solidline). For 5 , t , 8, a similar fit givesn ­ 1.6 6 0.2(data not shown). Although uncertain, the value ofnis compatible with Kolmogorov’s5

3 value. The timeevolution of ly2p obtained by representing the high-kregion of Ei

kin by a line spectrumElinekin is displayed in

Fig. 3. The length saturates beyondtmax at roughly threetimes its t ­ 0 value. Although the volume occupiedby the vortices has increased, it remains a small fractlpj2ys2pd3 , 0.4% of the total volume of the box. Thecomputations were performed withc ­ 2 correspondingto a root-mean-squareMach numberMrms ; jvTG

rmsjyc ­0.25. As it is very costly to decreaseMrms, we checked[17] that compressible effects were nondominant at thvalue ofMrms. In particular,Ei

kin is well above the otherenergy spectra throughout the run fork , kbump.

It is known [19] that NLSE vortex lines can reconnecThe vortex lines are visualized in physical spaceFigs. 4 and 5 at timest ­ 4 and t ­ 8. At t ­ 4, noreconnection has yet taken place while a complex vortangle is present att ­ 8. Detailed visualizations (datanot shown) demonstrate that reconnections occur fort .

5. Note that in the viscous TG vortex reconnection alsets in fort . 5.

As seen above, the spectral behavior of NLSE cancompared to viscous turbulence only fork # kbump ,d21

bump , where dbump is the average distance betweeneighboring vortices. We now estimate the scalingkbump in terms of the flow integral scalel0 and velocityu0 and of the velocity of soundc and coherence lengthj. The numbernd of vortex lines crossing a large-scale l2

0 area is given by the ratio of the circulationl0u0 to the quantum of circulationG ­ 4pcjy

p2, i.e.,

nd , l0u0ycj. Assuming that the vortices are uniformly

FIG. 4. Same visualization as in Fig. 1, but at timet ­ 4.

Page 4: Kolmogorov Turbulence in Low-Temperature Superflows

VOLUME 78, NUMBER 20 P H Y S I C A L R E V I E W L E T T E R S 19 MAY 1997

-

lueergt

h

-

u

y

sl

-

-as

diss

fen

t.

.

e)

.

).

,

FIG. 5. Same visualization as in Fig. 1, but at timet ­ 8.

spread over the large scale area givesnd , l20yd2

bump.Equating these two evaluations ofnd yields dbump ,l0

pscjdysl0u0d.

In the case of helium, the viscosity at the criticapoint (T ­ 5.174 K, P ­ 2.2105 Pa) is ncp ­ 0.27 3

1027 m2 s21 while the quantum of circulation,G ­hymHe has the value0.99 3 1027 m2 s21. Thus,4ncp ,G anddbump , l0y

pRcp , ll, whereRcp is the integral

scale Reynolds number at the critical point andll theTaylor microscale. The value ofdbump in a superfluidhelium experiment atT ­ 1 K is thus of the same orderas the Taylor microscale in the same experimental setrun with viscous helium at the critical point.

An experiment corresponding to the numerical resuof the present Letter must be performed at a temperatlow enough for the normal component of the flow to bneglected. In this regime, second sound attenuation msurements cannot be performed. Preliminary measuments (J. Maurer, private communication) in the swirlinflow of Ref. [14] did not seem to show any significanchange in energy dissipation for temperatures as low1.6 K where the normal fluid and the superfluid are in thsame proportion. It would be interesting to know if sucbehavior persists atT , 1 K. In viscous turbulence, itis well known that Kolmogorov’s theory is only approximate since it neglects intermittency [18]. Inertial rang“intermittency corrections” are measured [14,20,21] ovelocimetry data by monitoring the scaling of high ordemoments of velocity increments. If the corresponding sperfluid quantities could be measured experimentally blow T ­ 1 K by an as-yet-to-be-developed velocimetrprobe, significant differences might appear.

In summary, turbulent solutions of NLSE—and thulow-temperature superfluid turbulence—approximate

l

up

tsre

a-e-

ase

enr-

e-

y

obey Kolmogorov’s scaling. A question, open to experimental investigations, is to know the exact limitsof this analogy between superfluid and viscous turbulence. Some important experimental properties, suchKolmogorov’s scaling, still evade first principle deriva-tion from the Navier-Stokes equations [18]. These harproblems may be easier to solve, when the analogyvalid, using the NLSE rather than the Navier-Stokeequations.

Computations were performed on the C94-C98 othe Institut du Développement et des RessourcesInformatique Scientifique. We would like to thankL. Tuckerman for her helpful discussions on this work.

[1] L. Landau and E. Lifchitz,Fluid Mechanics(PergamonPress, Oxford, 1980).

[2] K. W. Schwarz, Phys. Rev. B31, 5782 (1985).[3] E. P. Gross, J. Math. Phys.4, 195 (1963).[4] V. L. Ginzburg and L. P. Pitaevskii, Sov. Phys. JETP34,

858 (1958).[5] R. J. Donnelly, Quantized Vortices in Helium II(Cam-

bridge University Press, Cambridge, England, 1991).[6] J. C. Neu, Physica (Amsterdam)43D, 385 (1990).[7] F. Lund, Phys. Lett. A159, 245 (1991).[8] P. Nozières and D. Pines,The Theory of Quantum Liquids

(Addison-Wesley, New York, 1990).[9] G. I. Taylor and A. E. Green, Proc. R. Soc. London A,

158, 499 (1937).[10] M. E. Brachet, D. I. Meiron, S. A. Orszag, G. Nickel,

R. H. Morf, and U. Frisch, J. Fluid Mech.130, 411 (1983).[11] M. Brachet, C. R. Acad. Sci. Ser. 2311, 775 (1990).[12] J. Domaradzki, W. Liu, and M. Brachet, Phys. Fluids A5,

1747 (1993).[13] S. Douady, Y. Couder, and M. E. Brachet, Phys. Rev. Let

67, 983 (1991).[14] G. Zocchi, P. Tabeling, J. Maurer, and H. Willaime, Phys

Rev. E50, 3693 (1994).[15] S. Fauve, C. Laroche, and B. Castaing, J. Phys. II (Franc

3, 271 (1993).[16] C. Nore, M. Abid, and M. Brachet, C. R. Acad. Sci. Ser

2 319, 733 (1994).[17] C. Nore, M. Abid, and M. Brachet, Phys. Fluids (to be

published).[18] U. Frisch, Turbulence, the Legacy of A. N. Kolmogorov

(Cambridge University Press, Cambridge, England, 1995[19] J. Koplik and H. Levine, Phys. Rev. Lett.71, 1375 (1993).[20] Y. Gagne, E. Hopfinger, and U. Frisch, inA New

Universal Scaling for Fully Developed Turbulence: TheDistribution of Velocity Increments,NATO ASI, Vol. 237,edited by P. Coullet and P. Huerre (Plenum, New York1990), p. 315.

[21] F. Belin, P. Tabeling, and H. Willaime, Physica (Amster-dam)93D, 52 (1996).

3899