• Aucun résultat trouvé

The oscillatory instability in rapid solidification

N/A
N/A
Protected

Academic year: 2021

Partager "The oscillatory instability in rapid solidification"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00246412

https://hal.archives-ouvertes.fr/jpa-00246412

Submitted on 1 Jan 1991

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.

The oscillatory instability in rapid solidification

Herbert Levine, Wouter-Jan Rappel

To cite this version:

Herbert Levine, Wouter-Jan Rappel. The oscillatory instability in rapid solidification. Journal de

Physique I, EDP Sciences, 1991, 1 (9), pp.1291-1302. �10.1051/jp1:1991207�. �jpa-00246412�

(2)

Classification Physics Abstracts

44.30 68.45 81.30F

The oscillatory instability in rapid solidification

Herbert Levine and Wouter-Jan

Rappel

Department

of

Physics,

Institute for Nonlinear Science,

University

of Califomia, San

Diego,

La Jolla, CA 92093, U-S-A-

(Received 4 January 1991, revbed 25

April

1991,

accepted

24

May 1991)

Rksumk. Les

bquations

habituelles de la solidification directionnelle sont modifibes dans le cas de vitesses de

tirage

blevdes

(solidification rapide).

A

partir

de ces

d(uations,

nous ddcrivons une mdthode

numdrique

pour

analyser

la stabilitd des structures cellulaires obtenues

numdriquement.

Ces cellules

prdsentent

une instabilitd oscillatoire, ce

qui prddit

l'observation de

rdgions

en forme de bandes dans

l'expbrience.

La

largeur

de ces bandes, que nous avons estimde, est I peu

prds

en

accord avec celle observde

expdrimentalement.

Abstract. We modify the usual directional solidification

equations

for the case of

large pulling

velocities

(rapid solidification).

Using these

equations

we describe

a numerical method to analyze the

stability

for

numerically

obtained cellular structures. We find cells which are

oscillatory

unstable which will lead to the observation of banded

regions

in the

experiment.

The width of these bands are estimated and found to agree

roughly

with the

experimentally

observed ones.

I. Inboduction.

It is well known that the process of

rapid

solidification can lead to various

interesting

microstructures of the

solidifying

materials. One of the

techniques

consists of the

scanning

of

a

high intensity

laser or electron beam over a surface

of,

for

example,

a metal

alloy.

The laser will melt the surface up to a certain

depth

and after the

passing

of the laser the melt will re-

solidify.

The

speed

of the solid-melt front

depends

on the distance from the surface and is maximal at the surface

[I].

One can model this type of solidification with the well-known

equations

of directional solidification

[2]. Here,

one

places

a bath of a two-component

alloy

in an

externally

fixed

temperature gradient

and moves the bath with a certain

pulling velocity

towards the

(fixed)

cold contact.

Typically,

the

pulling

velocities are of the order of micrometers per second. If

one assumes that the laser will create a fixed temperature

gradient

we will have a directional

solidification

experiment

with

pulling

velocities close to the

scanning velocity

of the laser. Of course, this allows for very

high pulling speeds.

The observed microstructures after the melt has solidified

depend

on the

velocity

of the laser and the

depth

of the melt.

Typically,

one observes cellular and dendritic structures for small and moderate velocities

[I].

For

higher velocities,

one observes banded

regions

and for

(3)

9

even

higher

velocities one has

microsegregation

free structures. These banded

regions

are the

ones we are

investigating

in this paper.

The

high

velocities will alter the usual set of

equations

for directional solidification since we

can no

longer

assume local

thermodynamical equilibrium

at the interface. 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- a bifurcation where the

eigenvalue

corresponding

to the

growth

rate of the

perturbation

has a non-zero

imaginary

part when the

real part crosses zero. This will result in the

growth

of an

oscillating perturbation

of the

interface and will

give

rise to banded

regions.

In this paper, we will

present

a numerical

approach

to the

problem

of

rapid

solidification.

We solve

numerically

the set of

equations

for directional solidification for

steady-state growth,

I-e- we

neglect

the time

dependent

concentration field.

Afterwards,

we use the obtained

steady

state cellular cells as the

input

for exact

stability

calculations.

Although

this

approach

has been used for usual directional

solidification,

it has never been

applied

to

rapid

solidification. Our main result is that the cellular solutions can go unstable to a

Hopf bifurcation, presumably giving

rise to banded cellular structures.

Kelly

and

Ungar [3]

have also found an

oscillatory instability,

however

they

have used a different model for the interface

temperature

and the modification of the

partition

coefficient

(see

Sect.

3)

for

high pulling velocity.

This paper is

organized

as follows. In section 2 we review the basic

equations

for directional solidification and

modify

these

equations

for the case of

rapid

solidification. In section

3,

we

review the linear

stability analysis

and

plot

the

regions

of

oscillatory

and

non-oscillatory

instabilities for the system of Sisb. We describe how we can obtain numerical

steady

state

equations

in section 4 and how we can

analyze

their

stability. Finally,

in section 5 we will present and discuss our results.

2. Basic

equations

and their modification.

The basic

equations

for directional solidification are well known and are described in many papers. In this paper we will

only

consider the one sided

model,

where we

neglect

the solute diffusion in the solid. In the frame

moving

with the

pulling velocity

uo we have for the

concentration field

c

~ 3c 3c

~~~~~°fi~&

uofi.y+u~

(fi Vc)~

=

(l k)

c~

D

Here,

D is the diffusion constant in the

liquid,

k is the

partition coefficient, i~

=

m~/G

is a

thermal

length (m~

is the

slope

of the

liquidus

and

T~

is the

melting temperature

of the pure

melt), do

=

«T~/Lm~

is a

capillary length («

is the surface free energy and L is the latent

heat)

and K is the curvature. The total interface

velocity

is uofi

y

+ u~ where fi is the normal to the interface. Of course, we have u~ =

0 for a

steady

state.

In

deriving

the above

equations

we have assumed local

thermodynamic equilibrium

at the interface. This

assumption

is correct for small

pulling

velocities but fails for

large

velocities. In that case we have to

modify

the above

equations.

As we will see, these modifications

give

rise to a different linear

stability

behavior and in

particular

there will now exist a

Hopf

bifurcation for

large

velocities.

(4)

Perhaps

the most

important change

within the

equations

is that the

partition

coefficient k has to be modified for

high

velocities. It will now

depend

on the

velocity

and should

approach unity

in the limit of

high speeds.

In this limit the interface moves so

quickly

that the solute can not diffuse out of the solid and will be

trapped

inside the

solid,

hence the name

« solute

trapping».

In

addition,

one can introduce the mechanism of attachment kinetics which will describe at which rate the molecules will attach to the

crystal.

There are a number of models for the

velocity dependence

of k. We believe that the actual functional

dependence

of k on v will not matter

greatly

for our purpose,

finding

the

oscillatory instability.

In

fact,

Coriell and Sekerka

[4]

have shown that for an

oscillatory instability

a non-

zero derivative

3k/3u

is sufficient. We have chosen a model

proposed by Boettinger

et al.

[5, 6]. They

introduce a

parameter flo

which measures the

departure

of k from its

equilibrium

(u

=

0 value and a parameter

Vo

which

mimici

the effect of kinetic attachment. In this paper

we will

only

consider the case

Vo

= tx~, which means that the atoms attach

infinitely

fast and there is no kinetic

undercooling, although

an extension for different values of

Vo

is

straightforward.

For the functional

dependence

of k

they

propose

k~+flou

~~~~~

l+flou

~~~

where

k~

is the

equilibrium

value for u

= 0.

Here,

u is the total normal

velocity

of the interface. As we can see, k

approaches unity

as u increases.

Boettinger

et al. showed that for a flat interface

moving

with

velocity

uo the

temperature

at the interface will be

given by

where

m(vo)

= mE

1 ikE k(I

In

(k/kE))i (4)

Here, Vo

measures the effect of attachment kinetics. As mentioned

before,

we will choose Vo = tx~ in the remainder of this paper.

Equation (4)

describes the

change

in the

liquidus slope

due to

non-equilibrium segregation.

As in

[7]

we suppose that the temperature at the interface is

given by

the above

expression

for the flat interface

plus

a Gibbs-Thomson term :

The above modifications will

change

the standard

equations

for the directional solidifi- cation. As in

[8]

we go into the frame

moving

with the

pulling velocity

uo and measure

length

in units of A

/2 (the wavelength

of the

cell),

concentration in units of c~ and time in units of

A~/4D. Then,

for the one sided

model,

we have

~ ~ 3c 3c

~~~

~~fi"A

(fi Vc)~

=

(l k(u))(2 pelf y

+

u~)

c~

~~

k()) f~ "*

~ ~~~

(5)

where y *

=

2

«Tp/Lm(u)

c~ A, pe

=

~~

,

the Peclet number and

f*

=

2 m

(u)

c~

Tp/GA.

4 D In our rescaled units the above

equations

read

k~

+

flu

~~~~

l +

flu

~~~

where

p

=

po

2

Dj> (8)

3. Linear

perturbation theory.

In this section we will review linear

perturbation theory

for the set of

equations (6)

as

described

by

Merchant and Davis

[7]

and

apply

the results to an actual system. As our system

we will choose the Sisb system described in

[3].

We will see that for this system we will not have an

oscillatory instability

if we do not

modify equations (I),

I-e- when we choose

po

=

0.

It is easy to see that

equations (6)

have a solution

corresponding

to a flat interface

~ ~

l

~-2pejy-yoj

°

k

~~~

Yo "

~

k where

k~

+ 2

pep

~

l + 2

pep

~~~~

We

perturb

the flat solution

c = co + eci e~~~ e~ QY e~~

(I I)

y = yo + eyi e~~~e~~

(12)

where

Q

= Pe +

~

(13)

and em I.

Substituting

these

expansions

in the above

equations

for

k(u)

and

m(u)

and

linearizing

for

e we get

k(u)

=

k

+ eA

j wyi + o

(e2)

m

(U)

= ni +

8A2

WYI + O

(8~) (14)

where

p (I k~) Ai

=

2Pe(1

+

fl

~

Ai

In

(k/k~)

~~

ni

(k~ )

~~~~

ni=ni~(1-_~[~ (k~-k(i-in(k/k~)))j.

(6)

Next,

we substitute

equations (14)

in

equations (6)

and we find

k~°~~l~i~~'~)~~~~j ~~

+

[~-2pe ~~-2pe ~~ w-2pe(1-k)( +yq~) +4pe~~ ~~=0 (16)

k

k k f k

where

2

«T~

~'

LmC~

A

(17)

2

nic~ TM f

"

~>

To find the neutral

stability

curve we set w = 0. This will

give

us a closed curve in

parameter (A

vs. u

)

space inside which the flat interface is unstable with a pure real

eigenvalue.

However,

it is also

possible

that an

w with non-zero

imaginary part

satisfies

equations (16).

If we write

Q

as

Q

= pe +

R,

we get a third order

equation

for R from which we can solve for

w. For each

wavelength

we get three solutions for w which the determine the nature of the

instability.

If the real part of

w crosses zero and its

imaginary

part is non-zero we will have an

oscillatory instability

and we will encounter a

Hopf

bifurcation. As we can see, a non-zero

Aj

or

equivalently

an non-zero

3k/3u, gives

us a cubic

equation

and for realistic parameter values we will find a

Hopf

bifurcation.

As an

example,

we have chosen the Sisb system as described in

[3].

The relevant material parameters are

given

in table I. We have

plotted

the boundaries of the

stability regions

for three different values of

flo

in

figures

1, 2 and 3. In

figure

I

fl

o = 0 and we see that there is no

oscillatory instability.

The

instability

in the unstable

region

is the Mullins-Sekereka

instability

and we expect to find cells

and/or

dendrites inside this

region.

For

flo

= I

x10~?

and

flo

= I x

10~~ however,

we do find an

oscillatory instability

for

high

velocities

(see Figs.

2 and

3).

There is a fundamental difference between the

figures

2 and 3. In

figure

2

flo

= I x 10~

?)

maximum of the neutral

stability

curved lies below the minimum of the

oscillatory

unstable

region. Therefore,

if one increases the

velocity,

one

expects

that the cellular structures which form inside the neutral

stability

curve bifurcate back to a

planar

interface. This

planar interface,

in turn, will

undergo

a

Hopf

bifurcation if one increases the

velocity

even more.

Since the

planar

interface will go unstable for a

perturbation

with

= tx~, the

planar

interface will oscillate in the direction of the

growth velocity.

The concentration left behind in the solid

Table I. Parameter values

of

the Sisb system used in the numerical calculation.

do

=

7.72 x lo ~~ m G

= 1.33 x

166 K/m T~

=

427 K mt =

452

Klat9b

c~ = I x

10~~

atfb D

=

1.5 x lo ~~

m~/s

k

=

0.023

(7)

~=0

lo ~

stable

j~

~

(

io 6

~/

unstable

>

io 5

lo 4

lo .I lo ° lo I lo 2

~w

(~llll)

Fig. I. The

regions

of

instability

for po 0.

J=lx10~~

lo 8

~

oscillatory unstable

~

E

~ lo 6

~

>

lo 5

io 4

10 .1 10 ° 10 10 2 10 3 10 4

~w

(~llll)

Fig. 2. The

regions

of instability for po = I x 10~~

depends

on the

velocity (k

is a function of

u)

and we expect to see banded

regions.

In

figure

3 however

(po

= I x

lo~~)

the cellular structures should not bifurcate to a stable

planar

interface since the

top

of the neutral

stability

curve is

already

in the

oscillatory

unstable

region. Instead,

one should find a

composite pattern

of bands of cells.

Although

we need a

(8)

J=lX10~~

oscillatory unstable

lo ~

~

~s

~

~~ ~

unstable

>

lo 5

~~/o

'i lo ° lo I lo 2 lo 3 lo 4

~w

(~Lm)

Fig.

3. The

regions

of

instability

for

po

=

I x 10~~

nonlinear

analysis

to examine these

patterns,

a banded cellular structure seems to be a resonable guess.

4. Numerical method.

Next,

we will describe how we can obtain numerical solutions of

equations (6).

Since this has been described elsewhere

[9, 10]

in great

detail,

we will be very brief. Once we have found a

steady

state

solution,

we

perturb

this solution with a normal

shift,

linearize about the

amplitude

of this shift and solve the

resulting eigenvalue problem.

The trick to solve

(6)

in a numerical efficient way is to rewrite the

equations

as a

integro-

differential

equation

and solve for the

(unknown

and

free) boundary.

We will do this for both the

quasi-static approximation

and for the full

equation.

In the

quasi-static approximation

we

neglect

the time derivative term in the diffusion

equation.

In the

quasi-static approximation

we find

(fl+y*K(s))

= i-

@+y*K(s'))(4"v'G)ds'+

+ 2

pek(u @

+ y * K

(s'i nj

G ds'

+

I

[k(u) @

+ y * K

(s') (()

+ y * K

(s')

G ds'

(18)

where the Green's function G is the solution of

V~G

+ 2 pe ~~

= 3

~(x x') (19)

°Y

(9)

The

explicit

form of G is

given

in

[8].

For the non

quasi-steady

case we go back to the rest frame and we find

(Y(S,

t

)

2 Pet

)

~ ~ ~ ~~ ~~ ~

j d~,

~

d~i

(Y(S', t')

2

Pet')

~ ~ ~ ~~, ~,

f*

'

-~

f*

'

fi'. V'G

(x(s,

t

), x'(s', t')

t,

t')

+ ds'

~ dt'( (~(~" ~') j ~P~~'~

y * K

(s',

t'

)

x

#

-w

where G is a solution of the full diffusion

equation

V~G

~~

=

3

~(x x')

3

(t t') (21)

To find

steady

state solution we

parameterize

the

(unknown) boundary

with N

points

of

equal ardength.

Our

independent

variables are then the normal

angles

at the

midpoints

of the

equal ardength

sections and the

position

of the

tip.

We solve the

integro-differential equation

on the

points

of

equal ardength,

except at the

endpoints (x

=

0 and x

=

I)

and

require

that the

slope

vanishes at both the

tip

and the tail. Once we have found a solution we can fit a

spline through

the solution to obtain a smaller discretization

step.

We

perturb

the

steady

solution

by adding

the normal shift

x(s)

=

xo(s)

+

ho (s)

3

(s,

t

(22)

where we assume that 3

(s,

t

= 3

(s)

e~~ and 3

(s)

« I. We substitute this

perturbation

in

(18)

and linearize in 3. This

gives

us the rather messy

looking equation

for the

quasi steady

state

approximation

ids'

yKo +

(3(s') fi(

V' + 3

(s) b0'

V

b~'

V'G +

f

+

@

y

(3"(s')

+

Kj

3

(s')) hi

V'G

+

lYKo+1°1(-3(S)to.vG+Ko3(S)no.vG)1

-2pek

ds'

b~ yKo+~° (3(s')fi(.V'+3(s)bo.V)G f

+

(yKo +9

(- 3'(s') ijG

+

4~

Ko

3(s') G)j

+

j- Y(3"(S')+ K13(s'))j

=w

ids'((I-k)~~°+yKo~ 3(s')G- ~9+yKo) bo.VGA~3(s')

f f

2

Pen)- °

+ YKo GA 3

(S)

+ 2

Pekni1 °

+ YKo GA 2 3

(S 1 (23)

(10)

Once we have chosen M

points

with

equal arclength ds,

the discretization of the

integrals gives

us a set of

equation

of the form

M M

£

A~j 3

j = w

£

B~y 3

j.

(24)

j o j o

This is

just

an

ordinary eigenvalue problem

for

w and can be solved with one of the standard numerical

packages.

For the non

quasi-static approximation

we

proceed along

the same lines. Substitution of the

perturbation

in

(20)

and

linearizing gives

us

ids' ~ yKo+~° (3(s')b(.V'+3(s)fro.V)b(.V'G(w)+

f

+

fi y(3"(s')

+

Kj3(s'))j hi V'G(w)

+ yKo +

) (- 3'(s') j

V'G

(w)

+ Ko 3

(s') hi

V'G

(w ))j

-2pek ds'fi~ yKo+~° f~ (3(s')fi(.V'+3(s)fro.V)G(w)

+

4~( @

y(3"(s')+ Kj3(s'))j G(w)

+

(yKo+9 (- 3'(s') ijG+

fi~ Ko

3(s') G(w))j

+

j- Y(3"(s'i

+

K13(S'ii)

= W

ids~lkli°+ Kol 3(S~IG(Wi- li°+ Kol no.vG(WiA23(S~i

2

pen(

+

yKo)

G

(w ) Ai

3

(s')

+ 2

peknj

+

yKo)

G

(w ) A~

3

(s')j (25)

f f

where an

explicit

form of

G(w )

is

given

in

[8]. Upon discretizing

the

integrals

we get an

equation

like

£

M

Cq(w

3~ = 0

(26)

j =o

where C is a function of w and therefore in

general complex.

To

satisfy

the above

equation

we

have to vary w until C has a zero

eigenvalue.

This is done with a Newton's

algorithm

which will converge to the

right

w after we have

guessed

an initial value for w. In

fact,

we can look for the

eigenvalues corresponding

to different modes.

S. Results and discussion.

The

quasi-static stability

calculation can be checked in several ways. First of

all,

we know from the linear

stability analysis exactly

for which

velocity

and

wavelength

the flat interface goes unstable. We have checked that our

stability

program does

predict

the

right

onset for

instability

of the flat

interface,

both for small velocities and

large

velocities.

(11)

Furthermore, carrying

out the

expansion

described in section 3 to third order

gives

us an

equation

for the

amplitude

A of the cell of the form

~ =aiA+a~A~ (27)

Depending

on the

sign

of a~, the so-called Landau

coefficient,

we have a sub- or

supercritical

bifurcation. It can be shown that for the one sided model and for values of k ~ 0.45 we will have a subcritical bifurcation

[11]. Indeed,

we do find a subcritical bifurcation at onset. The lower branch of the subcritical branch has to be

unstable, regaining stability

when it bends in the forward direction

again.

We have checked this behavior for the system at hand.

The full

stability

program can be checked as

well, using

the flat interface as our

steady

state solution. The linear

stability predicts

for which values of u and A the flat solution will lose

stability

to a cellular structure, will

regain stability

and lose

stability again

to an

oscillatory instability.

We are of course

especially

interested in the last

instability.

Let us take the case of

po

= I x

10~?

In table II

we have listed the

predicted eigenvalues

from the linear

stability

calculation of section 3 and the obtained values from our numerical

analysis.

We see that the agreement is excellent. We have also checked the lower branch of the subcritical bifurcation at onset and have found that the lower branch is unstable while the

upper branch is

stable,

which is as

expected.

Table II.

Eigenvalues for d#ferent flat interfaces (flo

=

I x

10~?).

~

(Jll/Sl

A

(iLJlll

t°numencal

linearstabil,ty

0.023 5 3.51 3.52

0.32 5 89.8 + I 54.3 89.9 + I 56.0

0.9 5 15.3 + I 1673.7 9.I + I 1681.6

Next, we have examined the

stability

of the cellular structures for

high velocities,

I-e- at the

top

of the neutral

stability region.

For

flo

=

I

x10~~

we have indeed found an

oscillatory instability

for the cellular solution. In

figure

4 we have

plotted

two

steady

state solution with

u =

2.3m/s

and u

=2.32m/s.

Here x and y are the

position

of the solution in the

computational

box

(the wavelength

of the solution is thus 0.15 ~Lm). We can search for the value of w

corresponding

to any mode. We expect the « zero mode which

corresponds

to a translation of the interface in the direction of the

growth velocity

to be the first mode to go unstable. The results of our numerics for this mode are

presented

in table III and the modes

are

plotted

in

figure

5. We have calculated w for different number of discretization

points

of the interface to check that the answer converges. We see that the real part of the

growth

rate

w crosses zero somewhere between u

=2.3m/s

and u

=2.32m/s. Thus,

the solution

corresponding

to an u = 2.3

m/s

is stable while the solution

corresponding

to an u = 2.32

m/s

is unstable with an

eigenvalue having

a non-zero

imaginary

part. This solution is thus

oscillatory

unstable The other modes were found to stable.

The cellular structures with

flo

= I x 10~ ? however are found to be

stable,

even at the

top

of the unstable

regime. This,

of course, is related to the fact that in the case of

larger

flo

the

planar

interface can restabilize before it

undergoes

the

Hopf

bifurcation.

(12)

1.34

v=2.32x10~

v=2.3x10~

~

/s

~

-l.35 i'>

~~'~~.00

1.50 3.00 4:50 6.00 7.50x10.2

X

(~m)

Fig. 4. The steady state solutions for v

=

2.3 m/s and u = 2.32 m/s

(po

=

I x

10~~).

Table III.

Eigenvalues for

the cells in

figure

4

for

a

d%ferent

number

of

discretization

points of

the

interface (A

= 0.15 ~Lm,

fl

=

I

x10~~).

v

(m/s)

w

(50 points)

w

(100 points)

w

(150 points)

2.3 0.I I + I 1.46 0.15 + I 1.41 0.15 + I 1.40

2.32 + 0.041 + I 1.43 + 0.036 + I 1.44 + 0.035 + I 1.44

We can also estimate the width of the

resulting

bands. This is

just

the

growth velocity

times the

period

of oscillation. For

w we take the value

corresponding

to the

marginal

stable mode

(I.e.

where the real

part

is

equal

to

zero).

Since we measure time in units of ~/4 D we have

aruo A~ W

=

(28)

In the above case we have

u=2.32m/s,

A

=0.l5~Lm

and w~l.5. Thus we find

W 3 ~Lm which is close to the

expeRmentally

observed bandwidths. Of course this is

just

a

rough

estimate since we have chosen a

quite arbitrary

value of

flo. Nevertheless,

the result

makes sense in a

physical

way.

In

conclusion,

we have

developed

a numerical way to examine the

stability

of

steady

state cellular structures for the one-sided model. We have found that modifications due to

large pulling

velocities can result in

oscillatory

unstable

cells, depending

on our modification

parameter

flo. Also,

the

resulting

bandwidths agree

roughly

with the

experimentally

observed

ones. Of course this

approach

cannot

explain

the

alternating

bands with and without structure

as observed in the

experiment.

For the exact

dynamics

of the cells we have to

perform

a

nonlinear

analysis.

(13)

xlo.I -o.70

0.80

o.90

-1,oo

V=2.32X10~

~ v=2.3x10~

i>~ ~'~°

-l.20

1.30

-1.40

0.00 1.50 3.00 4.50 6.00 7.50x10.2

X

(~m)

Fig.

5. The modes corresponding to the tigenvalues in table III.

Acknowledgments.

This work was

supported

in

part by

DARPA Grant No. N00011-86-K-0758.

References

[1] ZIMMERMANN M., CARRARD M. and KuRz W., Acta Metall. 37

(1989)

3305.

[2] See e.g. LANGER J. S., Rev. Mod. Phys. 52 (1980) 1.

[3] KELLY F. X. and UNGAR L. H.,

Phys.

Rev. B 34

(1986)

1746.

[4] CORIELL S. R. and SEKERKA R. F., J.

Crystal

Growth

3/4 (1976)

157.

[5] BOETTINGER W. J. and PEREPEzKO J. H, in

Rapid

Solidified

Crystalline Alloys,

S. K. Das, B. H.

Kear and C. M. Adam, Eds., Proc. of a TMS-AIME Northeast Regional

Meeting,

Morriston, New Jersey

(1985).

[6] BOETTINGER W. J. and CORIELL S. R.,

Rapid

Solidification Materials and

Techniques,

P. R.

Sahm, H. Jones and C. M. Adam, Eds. NATO

(1986).

[7] MERCHANT G. J. and DAvis S. D., Acta Metall. Mater. 38

(1990)

2683.

[8] KESSLER D. and LEVINE H.,

Phys.

Rev. A 41

(1990)

3197.

[9] KESSLER D. and LEVINE H.,

Phys.

Rev. A 39 (1989) 3041.

[10] LEVINE H. and RAPPEL W. J., Phys. Rev. A 42

(1991)

3197.

[ll]

CAROLI B., CAROLI C. and ROULET B., J.

Phys.

France 43

(1982)

1767.

Références

Documents relatifs

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

ﺺﺨﻠﻤﻟا Ostreopsis Ovata يﺮﮭﺠﻣ ﺐﻠﺤﻃ ، ﺔﯿﻠﺨﻟا ﺪﯿﺣو ﻢﺳ ﺞﺘﻨﯾ palytoxin ﺔﯿﺋاﻮﺘﺳﻻا رﺎﺤﺒﻟا هﺎﯿﻣ ﻲﻓ ةدﺎﻋ ﻢﯿﻘﯾ ،. ﮫﻧا ﻦﻣ ﺔﻋﻮﻤﺠﻣ ﻰﻟا ﻲﻤﺘﻨﯾ طﻮﺴﻟا

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

We show that the corresponding shape of the phase diagram induces a strong reduction of the scale of the bifur- cation curve which makes its complete exploration

a threshold pulling speed, for given temperature gradient and concentration, the interface becomes unstable and breaks down into a regular, periodic pattern (Fig.

interface observed during the controlled solidification of a dilute binary alloy, we performed a three-dimen- sional extension [3] of an existing

We can therefore conclude that, although the assumption made in (I) that amplitude and phase motion decouple is not justified at finite dissipation, its

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