• Aucun résultat trouvé

Interface dynamics and anisotropy effects in directional solidification

N/A
N/A
Protected

Academic year: 2021

Partager "Interface dynamics and anisotropy effects in directional solidification"

Copied!
14
0
0

Texte intégral

(1)

HAL Id: jpa-00246472

https://hal.archives-ouvertes.fr/jpa-00246472

Submitted on 1 Jan 1992

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Interface dynamics and anisotropy effects in directional solidification

S. de Cheveigné, C. Guthmann

To cite this version:

S. de Cheveigné, C. Guthmann. Interface dynamics and anisotropy effects in directional solidification.

Journal de Physique I, EDP Sciences, 1992, 2 (2), pp.193-205. �10.1051/jp1:1992133�. �jpa-00246472�

(2)

Classification

Physics

Abstracts

81.30 61.50 47.20

Interface dynamics and anisotropy effects in directional solidification

S, de

CheVeign6

and C. Guthmann

Groupe

de

Physique

des Solides

(*),

Tour23, 2 PI. Jussieu, 75251 Paris Cedex 05, France

(Received 6

May

1991, revised 4 October I991,

accepted

28 October 1991)

Rksum4. Une 6tude de la

dynarnique

de l'interface cellulaire en solidification directionnelle d'un

alliage

binaire dilu6 de CBr4 nous a

perrnis

d'observer un certain nombre d'instabilit6s secondaires

(propagation

de cellules incl1n6es, modes

optiques,

cellules anorrnales, etc.) dont certaines 6taient bien connues et traditionnellement attribu£es h des effets de

l'anisotropie

cristalline.

Cependant,

la

plupart

de ces instabilitds sont observables dons une

exp6rience qui

est

l'analogue hydrodynamique

de la solidification directionnelle, la

digitation Visqueuse dirig6e

ou

instabilit6 de

l'imprirneur

(M. RABAUD, S. MICHALLAND and Y. COUDER, Phys. Rev. Lent. 64 (1990) 184). Cette

analogie

va nous perrnettre de discuter (es r61es

respectifs

de

l'anisotropie

cristalline et de la

dynarnique

propre aux

systbmes

uni-dimensionnels.

Abstract. A

study

of the

dynamics

of the cellular

solid-liquid

interface in directional solidification of a dilute

CBr4 alloy

has allowed us to observe a number of secondary instabilities

(travelling

states, optical modes, anomalous cells, etc.) some of which were well known and

traditionally

attributed to the effect of

crystalline anisotropy.

However, most of these instabilities

are also observed in a

hydrodynamic analog

of directional solidification, directional viscous

fingering

(M. RABAUD, S. MICHALLAND and Y. COUDER, Phys. Rev. Lent. 64 (1990) 184). This

analogy

enables us to discuss the

interplay

of

crystalline anisotropy

and of the

dynamics generic

to one-dimensional systems in directional solidification.

Among

the various

dynamical

systems that exhibit

regular pattems

when out of

equilibrium,

those which can be considered as one-dimensional are of

particular

interest

II.

Their

dynamics

have been modelized and it has been shown that

they

can

develop

a number

of

secondary

instabilities as well as

spatio-temporal

chaos

[2].

In the present

article,

we

shali

describe some

aspects

of the

dynamics

of one such system, directional solidification of thin

samples

and their relations to

wavelength

selection. The

question

which will arise is : is the behavior of this

system generic,

identical to that of other lD

systems,

or does the

crystalline

substrate

play

a role ?

(*) Associated with Universities Paris 7 and Pierre-et-Marie-Curie and with the Centre National de la Recherche

Scientifique.

(3)

In directional solidification

(DS),

a

sample

is

pulled

at a

velocity

V in a constant

temperature gradient

G set up around its

melting temperature

in such a manner as to

progressively solidify

the

sample [3].

If the material is not pure, solute

rejection (or incorporation,

if the

partition

coefficient is greater than

I)

will occur at the

solid-liquid

interface. Above a critical

pulling speed,

for

given

concentration and temperature

gradient,

diffusion is no

longer

sufficient to assure the evacuation of solute and the interface

adopts

a

periodic

cellular pattem in which solute is accumulated in the cusps of the cells. The present

experiments

focus

particularly

on

non-stationary phenomena. They

were carried out on thin

(50

± 5 ~Lm unless otherwise

indicated) samples

of tetrabromomethane

(CBr4),

a transparent

organic material, containing

a fraction of a percent

impurities.

The temperature

gradient

was

G

w loo °/cm which

gives

a threshold

V~

= 8 ~Lm/s for the

planar-cellular

front transition. The

experiments

were observed under a

microscope

and

video-taped. (The experimental

set-up

and

previous stationary-state

results

conceming

this material can be found elsewhere

[4, 5]).

In the present work we have made extensive use of

spatio-temporal diagrams

made

by extracting

a

given

line from the

digitalized

video

image

at

given

time intervals

(typically

0. I to I

s)

and

recreating

an

image (with

one

spatial

dimension and one

temporal one) by

juxtaposing

these lines

[6].

Coullet and Iooss

[7]

have

shown, by

symmetry arguments

only,

that one-dimensional

interfaces can be

expected

to present a series of

instabilities,

a number of which have

already

been observed in directional solidification of

liquid crystals [8]

and eutectic

alloys [9],

in one- dimensional

Rayleigh-Bdnard

convection

[10]

and in directional viscous

fingering (DVF),

also known as the

printer's instability

it

].

This last system is of

particular

interest for us because it is a

hydrodynamic equivalent

to DS the

physical

mechanisms

driving

the instabilities are

equivalent

but the medium in which

they

appear has no

crystalline anisotropy.

In the

experiment,

oil is

trapped

between two

nearly tangential rotating cylinders,

one inside the other. The interface observed is that between the oil and air forced into the oil as the

cylinders

come apart.

(DVF

can be in fact more

complex

than DS because the

cylinders

can be co- rotated or

counter-rotated.)

A close

comparison

of our results with those obtained in DVF will lead us to reconsider the role of

crystalline anisotropy

in DS.

Travelling

waves.

Tilted cells were observed in the very first

experiments

on transparent materials carried out

by

Jackson and Hunt

[12]

in 1965. The effect is

traditionally

attributed to

crystalline anisotropy (essentially

of the attachment

kinetics)

: for cubic

crystals

such as

CBr4, growth

in the

(100)

direction is faster than in others. If the direction of fastest

growth

is not

perpendicular

to the

interface,

the cells grow

asymmetrically.

The

importance

of attachment kinetics increases with

pulling speed

until the

growth

is oriented

along

the

preferred direction,

as is believed to be the case for dendrites.

Anisotropic

interface kinetics were indeed shown to

produce

such

effects,

both

analytically by

Coriell and Sekerka

[13]

and

numerically (in

the

case of a low

partition coefficient) by Young

et al.

[141.

However, a number of observations

bring

us to

question

this

interpretation.

Inclined cells,

by definition,

grow at an

angle

to the normal to the solid

liquid

interface

(Fig, la)

and appear

as

travelling

waves on a

spatio-temporal diagram (Fig. lb).

A similar

diagram

has been

presented by

Gleeson et al.

[5]

for succinonitrile-ethanol. The lateral

velocity V~

of the cells is

simply

related to their

angle

of inclination a:

V/V~

= cotan a.

Figure

2 shows how the lateral

velocity V~

of the

travelling

wave increases with

pulling speed

V in a

given

«

grain

».

(This

term

normally

refers to a zone of

given crystalline

orientation. We shall

only

mean here

a zone of cells of a

given inclination.)

Note the presence of a well defined

threshold,

at about 25

~Lm/s,

below which no lateral movement is observed, to within our

precision

of

(4)

b)

a)

Fig.

I. a) A cellular front

presenting

inclined cells and both a source and a sink (V

= 66 ~Lm/s). b) The

corresponding spatio-temporal diagram.

Note the

following

features : A) an

asymmetric

source

(cells only divide on the upper side) ; B) the same source has become symmetric (cells divide on both sides) C) an

asymmetric

sink where altemate cells are

pinched

off D) is a gas bubble.

8

6

~

E

~ 2

e

0

(5)

measurement

(w0.5 ~Lm/s).

If the video camera is at all

misaligned

an apparent lateral movement will appear in the

spatio-temporal diagrams.

Whenever necessary, it has been subtracted out for

quantitative

measurements. In our

experiments,

the curve was not

unique

:

in different

grains

and

samples,

the lateral

velocity

increased with

pulling velocity

from the

same threshold but its value

beyond

this threshold

varied, depending probably

on the

crystalline

orientation of the

grain.

A curve similar to that of

figure

2 was

reported

in succinonitrile-ethanol

[13].

On the other

hand,

Trivedi

[16]

has studied the effects of

anisotropy

in

pivalic acid-ethanol,

a material for

which

they

are believed to be stronger than in

CBr4

or succinonitrile. In that case, no

threshold in lateral

velocity

is

reported

for misorientations of the

crystal

greater than 20° and the threshold observed

(at

15°

misorientation)

is

only

of 0.5

pm/s. Indeed,

no threshold in lateral

velocity

is

predicted by

the modelizations of the effect of cristalline

anisotropy

mentioned above

[13, 14]

: the lateral

velocity

was

predicted

to be

roughly

constant. So the threshold in lateral

velocity

does not seem to be related to cristalline

anisotropy.

In this context, it is

particularly interesting

to consider that inclined

travelling

states with

asymmetric

cells are observed in DVF

[10], although

no

crystalline anisotropy

is present.

Figure 3, kindly provided by Couder,

Michalland and

Rabaud,

shows a

spatio-temporal diagram

of such waves, on either side of a source

(time

runs from top to

bottom).

It is

possible

that in this system, translational symmetry

along

the air-oil interface be broken for instance

by

lateral flow of the oil. But there is no

equivalent

to the

microscopic anisotropy

of attachment kinetics.

These two observations lead us to believe that

microscopic

cristalline

anisotropy

may not be the

only

factor

influencing

the

growth morphology

of cells in directional solidification

although

cristalline orientation

apparently

affects their

precise angle

of inclination and that

more

generic dynamic

instabilities also

play

a role. We shall see further evidence of this in what follows.

Fig.

3. A

spatio-temporal diagram

with a

symmetric

source observed in directional viscous

fingering (kindly provided

by COUDER, MICHALLAND and RABAUD. The

spatial

coordinate is horizontal and time

increases downward). Note that a single central cell splits altemately to one side then to the other.

(6)

Let us first examine in more detail the

spatio-temporal diagram

of

figure16.

Boundaries

between

grains

appear on the

spatio-temporal diagrams

as sources where cells widen then

split

and sinks where cells narrow and

pinch

off. Sources can be more or less

symmetric

with cells

appearing

on both sides of the

boundary,

or

they

can be

asymmetric

with

only

cells

on one side

splitting.

In the upper part of

figure

16 one can see an

asymmetric

source become

symmetric.

We have

only

observed

asymmetric

sinks : one

grain

appears to be blocked

by

the other.

Figure

16

(feature C)

also shows a

particular

oscillation of the sink where the last cell and the second last cell are

altemately pinched

off.

In

DVF,

sinks and

asymmetric

sources appear to be identical to those

reported

here.

Symmetric

sources,

however,

are different

(see Fig. 3)

: a

single

central cell

splits altemately

to the left and to the

right

whereas in

figure

16 there are two central cells

clearly separated by

a continuous line

probably

a

grain boundary

and each cell

splits

to its own side. This is an illustration of one

possible

interaction between

dynamic

instabilities and the

crystal

:

sinks and sources seem to lock onto a defect of the

crystal,

I-e- a

grain boundary.

Asymmetric

cells and

wavelength

decrease.

A number of

secondary

instabilities can be described as cases where a

cell,

too wide

compared

to the normal

wavelength

at that

velocity, unsuccessfully attempts

to

tip-split (Fig. 4).

When

tip-splitting

succeeds, two new cells appear, but when it

fails,

the result is an

asymmetric

cell with a shoulder that is advected back towards the solid. The

instability

can

propagate

or not

along

the interface. It has been shown

[1,

7,

17]

that the presence of such modes is related to the

development

of the second harmonic of the interface deformation. The cell

shape

goes

from sinusoidal to « square » with a flattened center that can allow

tip-splitting.

The term

« mitten cells »,

suggested by

Cladis et al.

[18],

describes

large amplitude

cells with a thumb-like shoulder. A

given

cell

periodically begins

to

tip split,

but the

dip slips

off-

i

x ~

$ e

~

x ~ °

_

x

m

i lo loo

vl~L1n/s)

Fig.

4. Average

wavelength

I versus

pulling velocity

V for cells

presenting

optical modes (crosses) and mitten or

pre-splitting

instabilities (diamonds). The line represents the average

wavelength (dispersion

is about 10fb) for normal cells [4].

JOURNAL DE PHYSIQUE I T 2, N'2, FEBRUARY 1992 8

(7)

center so the deformation is advected

back, giving

the characteristic shoulder. The cell then

regains

its initial

shape

and the whole process starts

again.

The same

thing

can

happen

to its

neighbours, slightly dephased

in time, hence the

impression

of

propagation.

This is best seen

in

figure

5b. The

phenomenon,

first

reported by Venugopalan

and

Kirkaldy [19]

and observed in succinonitrile

by

Heslot and Libchaber

[20],

can also be described as a failed

tip- splitting.

The cells are

abnormally

wide

(here

68 ~Lm on average instead of a normal value of 50 ± 5 ~Lm)

(Fig. 4).

Mitten cells can also be observed in DVF

[I il.

G ""

,. "z

~' -~ 'T '$ ~~'

~ w'~

' r-L'

~~ Z '.S~_- S,

i n©' j~b ~G'

~

', ~$

"~ f .'

.I

~~ 'l '[

z ''

"

»,' M')_. t~)~.-~ ~l't

~Q~- j

~

~"~' -"-'

, .~~

&'~ ~~" -~

3 /- /.

_' ~- -~

_' ,' l-?i~

i -,,'

~

II- /,

t

b)

a)

Fig.

5. a) An interface

presenting

« mitten cells ». A

tip split

off center

gives

a shoulder that is advected back (V

= 10 ~Lm/s). b) The

corresponding spatio-temporal diagram.

Figure

6a shows two

pairs

of « anomalous cells

».

Again,

these cells are also observed in DVF

by

Rabaud et al.

[I II

and in DS of

liquid crystals [2 II.

The cells are

abnormally

wide and each one is flattened on the side

facing

the other. Different

aspects

of their

dynamics

are

illustrated in

figure

6b. The

pair

can remain stable for

quite long periods,

in which case it has

approximately

the width of three normal cells. It can be

progressively squeezed

out

(feature C).

If it is too wide it can

disappear through

the

splitting

of both cells

(feature D).

In these two cases, the

pair disappears

without

leaving

any trace. But another type of

instability

can also be observed the

pair

widens and the wall between the cells appears to oscillate until one of the two

splits (feature E,

or

Fig. 6c).

In fact the two cells

altemately undergo

a series of unsuccessful

tip-splittings

the oscillation of the middle wall is due to the advection back of

the shoulder on altemate sides until one of them

finally

succeeds. This can

happen

repeatedly.

In this case the

pair

seems to be locked onto a

grain boundary (we

have checked this

by examining

the

sample

at rest :

grain

boundaries form cusps where

they

reach the solid-

liquid interface).

Rabaud et al.

[22]

observe a similar

locking phenomenon

in DVF on a scratch made on one of their

rotating cylinders.

(8)

A-

C)

a)

w-

b)

Fig.

6. a) A front

presenting optical

modes (A) and 4

pairs

of anomalous cells (B). The lower cell of the lowest

pair

is splitting. b) The

corresponding

spatio-temporal diagram, with : A)

optical

modes B)

a

pair

of anomalous cells C) a

pair

becomes

progressively

normal D) a

pair disappears

after both cells

split

; E) a

pair

that

regularly splits

after

presenting

«

pre-splitting

» oscillations. c) A

blow-up

(x 2.5) of a «

pre-splitting

» oscillation

(T~~

w 0.6 s) and an

optical

mode

(T~~~ w 1.8 s).

(9)

Figure

7 shows the

propagation along

the front of a

single

cell deformation on a low-

amplitude

cellular front. These deformations are

analogous

to the mittens that are observed

on

higher amplitude cells,

in that

they

appear to be unsuccessful

tip,splits. They

are also observed in DS of

liquid crystals [81,

but above about 3 times the threshold

velocity

for the

transition from a

planar

to a cellular front. Here

they

can be observed much closer to

threshold,

due no doubt to the fact that the bifurcation is subcritical: the interface deformation is never

purely sinusoidal, higher

harmonics are present from the start and the

cells look square. This can also be seen on the

marginal stability

curve for

impure CBr4

(critical velocity V~

vs. wavenumber

q),

that is

particularly

flat-bottomed

[4]

: modes at 2 q

(and beyond)

are accessible

practically

at threshold.

Fig.

7. A sequence of

pictures

of the interface taken every 5.7 s

showing

the

propagation

of a solitary mode (V

= 14 ~Lm/s).

The

periods

of these «

pre-splitting

»

oscillations,

I.e. those of the central cell wall of anomalous cells and of the mitten

instability,

decrease as the

pulling speed increases, approximately

as

T~~~ oc

V~~ (Fig. 8).

This is also the time

scaling

of the diffusion time

(D/V~w

lo

s for V

= lo

~Lm/s),

as remarked

by

Gleeson et al.

[15].

The

periodicity

of

dendrite

side,branching

in DS of

CBr4

falls on the same curve

[23] (in

succinonitrile-acetone

[24]

the order of

magnitude

of the

periodicity

is the same but it decrease as V

~i.

One

can

speculate

that the same type of fluctuation of the

tip

can

(a) split

a

flat-tipped (rich

in second

harmonic) cell, (b)

be a little off center, fail to

split

it and be advected back or

(c)

in the case of

sufficiently pointed cells,

where

tip-splitting

is blocked

[5],

form side-branches I.e.

high frequency

shoulders that have room to

develop laterally.

Optical

modes and

wavelength

increase.

In

regions

of the

sample

where the cellular

wavelength I

is too small

(Fig. 4),

either because of a recent local

adjustment

of cell widths a cell has

split

in the

vicinity

or because the

(10)

loo

lo

]

w

~~4

o

~

I

T

i io

8. -

of

(11)

oscillation ceases and the

adjustment

then

spreads

out over the

neighboring

cells. Even in

cases where the cell walls do not touch at the

interface,

a rearrangement can take

place

behind

the interface

causing

the cells to

pinch

off

alternately, producing

the checkered pattem seen in

the middle of

figure

6a

[26].

The

dependence

of the

period

on

pulling velocity

is about

T~~~ oc

V~~'~ (Fig. 9),

I-e- it decreases more

slowly

than

T~~~. The reason for this different

dependence

is not clear.

Combined instabilities.

In another type of

secondary instability,

known as a tilt wave, a zone of inclined cells propagates

through

normal ones, as has been

clearly

observed in eutectic solidification

[9].

Tilt waves are not

frequent

in the

experiments

described

here,

but we have observed three

cases.

Figure

lo shows a tilt wave associated with an

optical

mode. This is once

again

very

similar to a

phenomenon reported

in DVF

[27].

The cases we observed were all at low

velocities

(6

to lo

~Lm/s)

and the inclined cells were at 7 ± 1° off the normal to the interface.

We sometimes observe

spatio-temporal diagrams (Fig. ii)

with an

apparently

chaotic

mixture,

in time and space, of anomalous

cells,

mitten instabilities and

optical

modes. Similar

diagrams

can be observed in DVF

[I ii

when the

cylinders

are co-rotated. « Chaotic »

grains

can coexist with

quite

stable ones at the same velocities so the orientation of the

grain

may

play

a role as

suggested by

Heslot and Libchaber

[24].

A proper

spectral analysis

of the time evolution of the

interface,

such as has been carried out on convection

experiments [28],

is be needed to be able to say if it is true

« chaos » we

observe, and,

if so, of what

type.

Fig.

10. -A

spatio-temporal diagram presenting

a «tilt wave». The tilted cells are affected by oscillations of the

optical

mode type (V

=

20 ~Lm/s).

Tip-splitting

in dendrites.

To

complete

this discussion of the

respective

roles of cristalline

anisotropy

and

dynamic

instabilities in

DS,

we wish to mention some

experiments

carried out on dendritic

growth

in very thin

samples (e

it ~Lm). The

stability

of dendrites is a case in which cristalline

(12)

b)

a)

Fig.

ll.-a) A «chaotic» interface, b) The

corresponding spatio-temporal diagram showing

a

mixture of anomalous cells,

optical

modes, etc. (V

= 80 ~Lm/s).

anisotropy,

either of attachment kinetics or of interfacial

tension,

is essential. This has been established

by

numerical simulations

[29-34], analytical

calculations

[35, 36], experiments

in viscous

fingering [37, 211,

where « cristalline »

anisotropy

is introduced

locally by scratching

to

plates

of the Hele-Shaw cell and in solidification

[38, 391

and

phase

transitions in

liquid crystals [40, 411.

More

indirectly,

dendrite

instability

has been observed when the effect of

cristalline

anisotropy

is

thought

to

weaken,

I.e. when

growth

slows down

[42, 43]

or in

homeotropic liquid crystal samples [44].

In the absence of

anisotropy,

or under conditions where

crystalline

and surface tension

anisotropy, aligned along

different

directions,

cancel each

other,

dendrites are unstable to

tip-splitting

and the result is a dense

branching morphology (DBM).

In

experiments [181

on

samples

of the order of I ~Lm

thick,

one

easily

obtains dendritic structures at velocities of the order of a few micrometers per second

(Fig. 12)

because the critical

velocity

for the transition from a

planar

to a cellular interface decreases

sharply

as

sample

thickness

decreases,

due to meniscus effects.

Therefore, simply by decreasing

the thickness

along

the

sample,

one increases the constraint. These dendrites

are unstable to

tip-splitting

and

produce

a structure similar to DBM. We suppose that a such

velocities, anisotropy

effects in

particular

kinetic

anisotropy

are too weak to stabilize

dendritic

growth.

Conclusion.

We have observed a number of

secondary

instabilities of cellular fronts in DS. These

instabilities

(optical modes,

anomalous

cells, solitary modes,

mitten

cells, travelling waves)

(13)

a) b)

Fig.

12.

Tip-splitting instability

of dendrites in very thin

samples

(= I ~Lm) : a) V

= 3 ~Lm/s b) V

=

6 ~Lm/s.

appear

quite

similar to those observed in

DVF,

the

hydrodynamic analogue

of DS.

Crystalline anisotropy

is not therefore essential to their

apparition although

the

crystalline

structure of the material affects inclination

angles

and locks

particular

structures

(anomalous cells,

sources

and sinks of

travelling waves).

When

growth

becomes

dendritic, crystalline anisotropy

is

necessary to stabilize the dendrite

tips against tip-splitting

and we have shown that in our

experiments

it is not

strong enough

at low velocities.

Acknowledgments.

It is a

pleasure

to thank Y. Couder, C.

Misbah,

S. Michalland and M. Rabaud for very fruitful discussions and for

allowing

us use their results

prior

to

publication.

This work was

supported by

the Centre National d'Etudes

Spatiales.

References

[1] FLESSELLES J.-M., SIMON A. J., LIBCHABER A. J., to appear in Adv. Phys.

[2] CHATt H., MANNEVILLE P.,

Phys.

Rev. Lett. 58 (1987) l12.

[3] LANGER J. S., Rev. Mod. Phys. 52 (1980) 1.

[4] DE CHEVEIGNt S., GUTHMANN C., LEBRUN M. M., J. Phys. France 47 (1986) 2095.

[5] KuRowsKI P., DE CHEVEIGNt S., GUTHMANN C., Phys. Rev. A 42 (1990) 7368.

[6] Caillou software from Stone

Age

on

Apple

Macintosh.

[7] COULLET P., Iooss G., Phys. Rev. Lett. 64 (1990) 866.

[8] SIMON A. J., BECHHOEFER J., LIBCHABER A.,

Phys.

Rev. Lett. 61 (1988) 2574.

[9] FAIVRE G., DE CHEVEIGNt S., GUTHMANN C., KuRowsKI P., Europhys. Lett. 9 (1989) 779.

(14)

[10] DuBoIs M., DAVIAUD F., BONETTI M., in Nonlinear Evolution of

Spatio-Temporal

Structures in

Dissipative

Continuous

Systems,

F. H. Busse and L. Kramer Eds. NATO ASI series (Plenum Press, 1990).

[ll] RABAUD M., MICHALLAND S., COUDER Y.,

Phys.

Rev. Lett. 64 (1990) 184.

[12] JACKSON K. A., HUNT J. D., Acta Metall. 13 (1965) 1212.

[13] CORIELL S. R., SEKERKA R. F., J. Cryst. Growth 34 (1976) 157.

[14] YOUNG G. W., DAVIS S. H., BRATTKUS K. E., J.

Crystal

Growth 83

(1987)

560.

[15] GLEESON J. T., FINN P. L., CLADIS P. E.,

Phys.

Rev. Lett. 66 (1991) 236.

[16] TRIVEDI R.,

Appl.

Mech. Rev. 43 (1990) 579.

[17] FAUVE S., DOUADY S., THUAL O., J.

Phys.

II France 1 (1991) 311.

[18] CLADIS P. E., GLEESON J. T., FINN P. L., in Defects, Patterns and Instabilities, D.

Walgraef

Ed.

(Kleuwer Academic Publishers, 1990).

[19] VENUGOPALAN D., KtRKALDY J. S., Acta Metall 32 (1984) 893.

[20] HESLOT F., LIBCHABER A.,

Physica Scripta

T9

(1985)

126.

[21] OSWALD P., J.

Phys.

II France 1 (1991) 571.

[22] RABAUD M., MICHALLAND S., COUDER Y., Private communication.

[23] KuRowsKI P., DE CHEVEIGNt S., GUTHMANN C.,

unpublished

results.

[24] ESAKA H., Thesis, Ecole

Polytechnique

Fdd6rale de Lausanne (1986).

[25] DE CHEVEIGNt S., GUTHMANN C., LEBRUN M. M., J. Phys. France 47 (1986) 2095.

[26] DE CHEVEIGNt S., GUTHMANN C., in

Propagation

in Far from

Equilibrium Systems,

J.E.

Wesfreid Ed.

(Springer-Verlag,

1988).

[27] MICHALLAND S., RABAUD M.,

unpublished

(1991).

[28] CILIBERTO S., BIGAzzI P.,

Phys.

Rev. Lett. 60

(1988)

286.

[29] KESSLER D., LEVINE H., Phys. Rev. B 33 (1986) 7867.

[30] BEN-JACOB E., GARIK P., MUELLER T., GRIER D.,

Phys.

Rev. A 38 (1988) 1370.

[31] NITTMAN J., STANLEY H. E., Nature 321 (1986) 663.

[32] FONG LIU, GOLDENFELD N.,

Phys.

Rev. A 42 (1990) 895.

[33] CLASSEN A., MISBAH C., MULLER-KRUMBHAAR H., SAITO Y., to appear in

Phys.

Rev. A.

[34] BRENER E., MULLER-KRUMBHAAR H., TEMKIN D.,

unpublished

(1991).

[35] BEN AMAR M., POMEAU Y.,

Europhys.

Lett. 2 (1986) 307.

[36] BRENER E. A., Sov. Phys. JETP 69 (1989) 133.

[37] BEN-JACOB E., DEUTSCHER G., GARIK P., GOLDENFELD N. D., LEREAH Y., Phys. Rev. Lett. 57 (1986) 1903.

[38] HONJO H., OHTA S., MATSUSHITA M., J.

Phys.

Sac. Jpn. 55 (1986) 2487.

[39] HONJO H., OHTA S., MATSUSHITA M.,

Phys.

Rev. A 36 (1987) 4555.

[40] ARORA S., BUKA A., PALFFY~MUHORAY P., RAcz Z., VORA R.,

Europhys.

Lett. 7 (1988) 43.

[41] BUKA A., PALFFY~MUHORAY P., J. Phys. France 49 (1988) 1319.

[42] SAWADA Y.,

Physica

A 140 (1986) 134.

[43] SAWADA Y., DOUGHERTY A., GOLLUB J.,

Phys.

Rev. Lett. 56 (1988) 1691.

[44] BECHHOEFER J., OSWALD P., LIBCHABER A.,

Phys.

Rev. A 37 (1988) 2574.

Références

Documents relatifs

The shift of the convective bifurcation, though much larger, can also be calculated with very good accuracy in the same range of values of the thermal gradient with the

Si on suppose que le domaine géométrique Ω du système (2.1) est connu avec une certaine précision, alors de cette précision prés, le choix d’un seul actionneur ( p = 1) peut

The, the implémentation of a law of non- linear control, high performance adaptive applied to a synchronous generator connected to an infinite network, to improve

Dans ce travail, une nouvelle théorie trigonométrique non locale de déformation de cisaillement est proposée pour l'analyse de la stabilité thermique des nano-plaques FG

When solute disturbances in the liquid meet a cellular solidification front, they first affect the cell tips and are then convected down the roots at the pulling

It can be shown that a linear stability analysis of the standard equations of directional solidification modified for high pulling velocities exhibit a Hopf bifurcation, I-e-

agrees better with the prediction of Karma and Pelc4. For this value of k, our numerics agree better with the solid curve for small values of pe. This is expected since their

order to separate thermal effects due to convective motions from the ones imposed by solidification, experiments have been performed without pulling.. Therefore no