• Aucun résultat trouvé

On directional solidification of a faceted crystal

N/A
N/A
Protected

Academic year: 2021

Partager "On directional solidification of a faceted crystal"

Copied!
16
0
0

Texte intégral

(1)

HAL Id: jpa-00211003

https://hal.archives-ouvertes.fr/jpa-00211003

Submitted on 1 Jan 1989

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.

On directional solidification of a faceted crystal

R. Bowley, B. Caroli, C. Caroli, F. Graner, P. Nozières, B. Roulet

To cite this version:

R. Bowley, B. Caroli, C. Caroli, F. Graner, P. Nozières, et al.. On directional solidification of a faceted crystal. Journal de Physique, 1989, 50 (12), pp.1377-1391. �10.1051/jphys:0198900500120137700�.

�jpa-00211003�

(2)

On directional solidification of a faceted crystal

R. Bowley (1), B. Caroli (2, 3), C. Caroli (2), F. Graner (1), P. Nozières (1)

and B. Roulet (2)

(1) Institut Laue-Langevin, BP 156, 38042 Grenoble Cedex, France

(2) Groupe de Physique des Solides de l’Ecole Normale Supérieure, associé au Centre National de la Recherche Scientifique, Université Paris VII, 2 place Jussieu, 75251 Paris Cedex 05, France

(3) Département de Physique, U.F.R. des Sciences fondamentales et appliquées, Université de Picardie, 33 rue Saint-Leu, 80000 Amiens, France

(Reçu le 20 janvier 1989, accepté le 8 mars 1989)

Résumé. 2014 Nous étudions l’instabilité morphologique d’un mélange binaire en solidification directionnelle, dans le cas l’interface plan initial est une facette. Nous montrons que, pour des vitesses de tirage plus grandes que le seuil standard de l’instabilité de Mullins-Sekerka, il existe un continuum de formes de front stationnaires non planes périodiques et de petite amplitude,

formées d’une succession de facettes chaudes et froides reliées par des régions courbes. Ces solutions sont instables par rapport aux fluctuations d’amplitude ; elles peuvent donc être considérées comme définissant un seuil pour l’amplitude de l’instabilité ordinaire. Celle-ci devrait donc présenter une très forte hystérésis.

Abstract.

2014

We study the Mullins-Sekerka instability of a binary mixture submitted to directional solidification in the case where the basic planar solid-liquid interface is a facet. We show that, for pulling velocities larger than the standard MS instability threshold, there exist a continuum of

stationary non-planar periodic front profiles of small amplitude, consisting of an alternation of hot and cold facets connected by curved regions. These solutions are unstable as regards their amplitude, so that they may be viewed as an amplitude threshold for the usual cellular instability,

which should therefore exhibit an anomalously large hysteresis.

Classification

Physics Abstracts

61.50C - 64.60 - 64.70D

Dilute binary mixtures, when submitted to directional solidification (pulling at an imposed velocity V along the z axis in an external thermal gradient G parallel to V) exhibit, at a

threshold velocity Vc(G), a morphological instability. The solid-liquid interface, which is planar for Vu Vc, develops, for Vu Vc, a quasi-periodic structure. This instability results

from the competition between the destabilizing effect of chemical diffusion and the stabilization due to capillarity and to the temperature gradient. It was first analyzed by

Mullins and Sekerka [1] who studied the linear stability of the planar front for materials with

an atomically rough solid-liquid interface and an isotropic surface tension. Much experimental

and theoretical work has been performed since in order to analyze in detail both the vicinity of

the bifurcation [2-4] and the shapes of stationary front cells far above threshold [5, 6]. Most of

these studies deal with the case where the attachment kinetics is instantaneous (local

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198900500120137700

(3)

equilibrium on the front) and the surface tension isotropic, except for some extensions to systems with (isotropic or anisotropic [7]) linear kinetics and a smooth capillary anisotropy [8].

That is, all existing theories of cellular fronts are concerned with solids which do not exhibit facets on their equilibrium shape. More generally, the question of growth shapes of faceting

solids when the solidification dynamics is, at least partly, controlled by diffusion [9], has

received little attention with the noticeable exception of the work of Chernov and collaborators [10] on freely growing crystals. A significant opening has been made recently by

Ben Amar and Pomeau. In their work on faceted needle-crystals [11] they were able to

formulate explicitly the equations describing the growth of a front containing faceted parts.

Namely, they show that the kinetic growth relation on the front, which is local in the curved

regions, must be replaced, on the facet, by an integrated condition relating the averages of the relevant (temperature and/or concentration) fields along the facet with its size. They were

then able, on the basis of a scaling argument, to make some predictions about velocity

selection for the needle-crystal.

In order to make further progress on the problem of the coupling between diffusion and

faceting, and of its incidence upon dynamical front morphologies, it seems useful to address a question a priori simpler than dendritic growth, namely that of the Mullins-Sekerka instability

of a faceted front under conditions of directional solidification.

As already mentioned, when the equilibrium shape of the crystal is smooth i.e. when the Wulff plot of the surface energy is not cusped - the existence and position of the instability

can be established on the basis of the MS linear stability analysis. It would of course seem

natural to try and apply the same perturbative approach to the case where the basic planar

front is a facet. However, it appears that the presence of a cusp in the Wulff plot in the

direction of the unperturbed interface entails that any deformation of this front, however small its amplitude, belongs, from the point of view of the MS approach, to the strongly non-

linear regime. This can be easily understood qualitatively.

In all that follows, we will only consider 1-D front deformations. Consider a situation where :

(i) the liquid-solid interface energy a ( 0 ) (where Ois the angle between Oz and the normal to the interface pointing into the liquid ; see Fig. 1) is the equilibrium one [12], and its

temperature and concentration dependence are negligible in the T and C ranges of interest ; (ii) the equilibrium facet matches smoothly (tangentially) with the contiguous curved regions. That is [13],6 + d 2o--Id 02 is positive everywhere on the Wulff plot - except, of

course, for 0

=

0 and in the crystallographically equivalent directions, where it is undefined.

We assume for the moment that interface kinetics is everywhere instantaneous. This

assumption, which is unessential for the following qualitative argument, will be relaxed later.

A cusp in a(0 ) in the 0

=

0 direction gives rise to a contribution à o,’ . 8 ( e ) (with

Acr’ = (d a /d 0 )0 + - (d a /d 0 )o _ ) to the

«

surface stiffness

»

(o’ + 0, " ) which appears in the Gibbs-Thomson equation expressing local equilibrium on the front. Let us imagine that we

round off this cusp on a very small angular width e (see Fig. 2). (er + U") becomes a smooth function, so that the MS linear expansion can now be formally performed with the usual results in which the surface stiffness parameter has to be understood as lQ = (u + u")8 = o.

Note that, due to the smallness of E, lQ

=

£- 2 a (0 ) is much larger than for a rough system.

This entails that the region of the (G, V ) plane where the planar front is linearly unstable

becomes correspondingly small [3].

Of course, the next question to be asked is that of the limits of validity of the linear

perturbation treatment. The perturbation expansion in powers of the front deformation

amplitude relies on the Taylor expansion of the growth equations. In the presence of the

narrow peak in a + o-" (Fig. 2), the linear approximation therefore fails as soon as the slope

(4)

Fig. 1.

-

Schematic sketch of a directional solidification setup.

Fig. 2. - Schematic shape of the surface energy (a) and surface stiffness (b) around the facet direction 8

=

0. The dashed curves correspond to the smoothed cusp approximation.

of the front profile 03BE(x ) is such that 0

=

tan - 1 (d C ldx )

>

c. In other words, it is meaningful only for profiles containing no orientations but those spanned by the equilibrium quasi-facet resulting from the rounded-off cusp of 0152

That is, a cusp in a (£ -+ 0) corresponds to a vanishing range of validity of the linear

approach. This increases considerably the difficulties of an exploration of the dynamics of

deformation of a growing facet. For this reason, we limit ourselves in the following to the

(5)

much more restricted problem of investigating whether there exists stationary front shapes

which are non-planar, periodic and with a small amplitude of deformation with respect to the original facet. We show that this problem has a simple analytical solution in the highly non-

local limit where the space-period of the front structure is much smaller than the chemical diffusion length - a condition which is easily satisfied in usual experimental situations.

We find that, for V Vcs (G - where Vcs (G) is the so-called constitutional supercooling

threshold [14]

-

no small amplitude stationary non-planar solution exists. Above

Vcs, we show that there is a continuum of such stationary

«

crenellated

»

profiles, consisting

of an alternation of hot and cold facets connected by curved parts, and calculate the relation between the height and size of the facets.

We then show - in the limit where the facets are much longer than the curved regions

-

that these solutions are unstable against small amplitude fluctuations. That is, the branch of crenellated states has to be « jumped over » in order for the locally stable facet front to

restabilize into a Mullins-Sekerka structure of partly faceted cells. This should result in a very strong hysteresis of the front morphology, related to the depth of the cusp in the Wulff plot,

i.e. to the magnitude of the step energy for the relevant facet.

1. The front equation.

We describe our two-phase system by the usual one-sided model [2], that is we neglect

diffusion in the solid. For the sake of simplicity, we assume that the two phases have equal

thermal diffusivities and consider thermal diffusion as instantaneous on the scale of chemical

diffusion, so that the thermal profile is unaffected by the solidification process [3]. Let

G

=

dT/dz be the constant externally imposed temperature gradient.

We consider here 1-D front deformations. Finally, we assume that the surface energy

a ( 0 ) is locally symmetric about the facet orientation, i.e. that 0

=

0 is a direction of high symmetry of the solid.

In the laboratory frame, the solidifying system, pulled at velocity V//Oz, is then described

by the following set of equations [1, 2] : a) in the liquid (z:>03BE (x, t ) )

b) on the front (z = ’(x, t»

-

Solute balance equation :

-

Gibbs-Thomson equation :

In equations (1-3) lengths, times and concentrations are measured respectively in units of the solute diffusion length f = 2 D /V, the diffusion time ’T = 4 D /V2 and the solute concentration gap AC

=

Coo(1 - K)/K for the planar stationary front (i.e. for C,

=

C (0).

Crois the imposed concentration of the liquid far ahead of the front, K the equilibrium partition coefficient of our (dilute) mixture, û the unit vector normal to the front,

k = - 03BE,l(l + ep2)3/2 the front curvature, and 8 (x, t )

=

tan-l(ô(x, t)lax).

u = ÎGI 1 ML AC 1,

,

where mL is the slope of the liquidus line on the phase diagram.

(6)

y ( 0 ) = Tm u ( 0 )/ (Lf mL LlC 1 ) ; L is the specific latent heat, TM the melting temperature of the pure solvent. e is measured from the T

=

Tm isotherm.

Finally, the Y term in equation (3) accounts for the undercooling effect associated with attachment kinetics. For the sake of simplicity, we assume this process to be instantaneous in all but the facet direction :

where 8 Tkin (v ) is the kinetic undercooling of the facet growing at velocity v.

The generalized Gibbs-Thomson equation (3) is local for 0 # 0, where the « surface stiffness » ( y + y " ) is well defined. In the absence of surface friction (F

=

o ), it expresses local equilibrium of the interface. For a facet, the situation is different, as equilibrium is now a global property, corresponding to a neutral balance of, say, terrace nucleation and spreading.

Equation (3) must be integrated across the facet, yielding the global constraint :

(remember that K

=

d 0 Ids). 0 y’ - y’, - yo

>

0. xo, xi are the abscissae of the endpoints

of the facet of height lf. Ef = - 1 (resp. + 1) for a

«

quasi-convex » (resp. concave) solid

facet. We define as quasi-convex a facet such that 0 (xo- )

=

0+ ; 9 (xl+ )

=

0_ (see Fig. 3).

When F

=

0, (5) is simply a statement that the energy is stationary when one full terrace, with height a, is added to the facet : the cost of the peripheral steps, 2 | yÓ 1 a = a 0 y’ exactly

balances the bulk energy gained by growing an extra solid volume a (xl - xo ).

Note that since, as mentioned above, we assume that there are no missing orientations on

the equilibrium shape ( y + y "

>

0 for all 0 # 0), mechanical equilibrium of the curved parts, which is instantaneous on the time scale of interest here, imposes that the curved parts of the profile match tangentially with the facets.

Fig. 3.

-

Non planar periodic front profile. The planar region - b x b is what we define as a quasi-

concave facet, À /2 - a .: X .: À /2 + a corresponds to a quasi-convex one.

(7)

The basic stationary solution corresponding to an infinite facet is immediately obtained by looking for an x-independent solution of equations (1), (2) and taking in equation (5) the

(xl - xa )

-->

00 limit. It is given by :

The planar front position (0) is :

which simply describes the kinetically induced recoil of the facet from the local equilibrium position.

We now look for small amplitude non-planar stationary periodic solutions of equations (1)- (3). The shape of such a solution is depicted in figure 3. Since y (0) is even, we may restrict ourselves to front shapes with reflection symmetry about the centre of a facet. We call

2 a, e + (resp. 2 b, 03BE - ) the length and height of a quasi-convex (resp. concave) facet, c the

extension along Ox of each curved region. À = 2 (a + b + c ) is then the space period of the profile.

2. A simple treatment of stationary front profiles.

We first give simple arguments that are sufficient to provide stationary shapes in the strongly

non local limit à 1 (i.e., wavelength much smaller than the diffusion length f), but which

cannot account for the dynamics of these fronts. Detailed calculations will be given in the following section.

A stationary shape can be constructed in two steps :

(i) first find a diffusion profile that obeys the conservation law (2) for a given front profile ).

(ii) Then express surface equilibrium (3) and (5) as a constraint that specifies 03BE(f).

For the one-sided model in the limit À « 1, the first step is indeed trivial : the diffusion

profile that satisfies conservation laws is just the zeroth order one, (6). Assume for a moment

that the exponential in (6) is approximated by a linear law

The diffusion current in the liquid phase is unaffected by interface deformations

03BE(x), and it guarantees solute conservation for any stationary shape of the interface (the

volume swept by the interface is shape independent). In a two-sided model, the concentration fluctuation along the distorted interface would induce a short-circuit current in the solid phase

that would spoil conservation ; here this complication does not occur. Thus C (0)(z) is the

appropriate diffusion field, but for corrections due to curvature of the exponential. In order to

estimate the corresponding error, we note that the defect of current at the interface is

-(AC)V03BE/l (we restore the original units in order to make the argument more

transparent). Thus an extra correction 8C must appear, yielding that missing current on a

scale - À in the liquid, hence :

The error is of order (03BEÀ / £2), very small if à f. In contrast, the short-circuit current in the

(8)

solid in a two sided model would be of order (D/A). (oc ) elf, yielding a correction -- ’1 À, comparable to the main term in the equilibrium condition (3). In this respect, the one

sided model is much simpler than the two sided one.

.

From here on, we assume À « l, and we take C (0)(z) as the stationary bulk diffusion

profile. We first express the equilibrium condition (3) on the curved parts, where there is no

friction. To first order in Ç, the condition is obeyed if the local wave vector, q

=

’TT lc, is

neutral with respect to the usual cellular instability, corresponding to an edge of the instability

range, which exists only if V exceeds the appropriate instability threshold Vcs (G). Here, it

must necessarily correspond to the upper edge, since we have à f. Thus, the curved parts

are half arches of a sinusoid, with wavevector qmax, and with an arbitrary amplitude 03BEf (within our first order approximation). For a non-faceted interface, these stationary shapes

are unstable as regards their wavevector (Eckhaus instability).

We now turn to the integrated equilibrium condition for the facet. The facet height, ::t ’f, is fixed by the curved parts, and it is symmetric for upper and lower facets : we need only

calculate their stationary widths, 2 a and 2 b. Since we ignore for the moment surface dissipation, equation (5) predicts equal widths for the upper and lower facets :

The smaller the displacement 03BEf, the larger the equilibrium width. The relationship is more apparent if we write it in the form :

where Fo is the surface stiffness near 0

=

0. Equation (11) may be viewed the opposite way,

as yielding the stationary amplitude ef for a given global wavelength À. Contrary to the non-

faceted case, there exists for each À « f a stationary deformed front profile with a given amplitude, f that varies roughly as A -1 for long wavelengths.

We will check in section 4 that such a stationary state is unstable as regards its amplitude

-

a quite unsurprising result, since ultimately one must recover the usual Mullins-Sekerka cellular profiles at large amplitudes when faceting effects become minor. The stationary state

described here must consequently be viewed as an amplitude threshold for the cellular

instability, which must be overcome before the deformation grows to its usual, non-linear

value. Since that state is unstable, the issue of phase stability is of no concern - we should remember only that the amplitude threshold is lower at long wavelength.

The above discussion is easily generalized to the case of a finite surface dissipation, F 96 0. As will be shown in the next section, it is found that the upper facets broaden, while

the lower facets shrink :

in which £ is a characteristic length which measures friction :

As the wavelength grows, (f decreases and the upper facets become progressively dominant.

When ef f-->. L, the upper facets invade the front altogether and the distorted stationary state

disappears.

(9)

3. Detailed description of the stationary front.

Since we only investigate deformations of small amplitude (03BE +- 03BE- ), we linearize the diffusion problem described by equations (1), (2) [15]. We set :

where k

=

2 7r/À, with À the period of the profile.

After linearization of equation (2), one gets [1, 2] :

Inserting this expression of the concentration field into the generalized Gibbs-Thomson

equation (3) and linearizing C (x, (0) + ((1)(.X)) in ((1), one obtains the front equation for

our small amplitude problem as :

Integrating equation (14) along each facet, one obtains the linearized version of equation (5),

so that the front equation explicitly reads : (i) for a quasi-concave facet (- b - x - b ) :

where we have made use of the reflection symmetry of C (x) ; (ii) for a quasi-convex facet (À /2 - a : X : À /2 + a ) :

(iii) in each curved part

where :

(10)

In equation (18), since the singular part of y ( 0 ) does not appear, we have linearized the

capillary term.

We must now solve the three coupled equations (16), (17), (18) together with the matching

conditions :

Note that (16), (17), (18) are a system of linear but inhomogeneous equations. It is the

inhomogeneous capillary term, characteristic of the presence of a facet, which accounts for

the fact that our problem is not fully linearizable in the sense of the Mullins-Sekerka analysis.

The fact that the kinetic term also gives rise to a non-homogeneous contribution results from

our assumption (Eq. (4)) that the 8-dependence of 37(v, 6 ) is discontinuous ; this model- dependent feature is unessential to our results, since it does not modify the basic structure of equations (16), (17) which is determined by the capillary singularity.

The strongly non-local limit is, a priori, characterized by A « 1, i.e. the physical wavelength

À must be much smaller than the diffusion length.

In this limit, for n _ 1, mn - nk > 1 and the sum on the r.h.s. of equation (14), which

describes the distortion of the diffusion field by the deformation of the front profile, is easily

checked to be, at most, of order À . In k . Max ( 1 C(l) 1

,

l ,03BE(1)1). So, it can be neglected as

compared with terms on the I.h.s. of equation (18) as soon as :

which provides a more accurate definition of what we call the strongly non-local limit. That is,

in this limit we approximate the diffusion field in the presence of the actual distorted front by

that associated with its planar average.

The same argument applies to equations (16), (17) so that, in the non-local limit, we are left

with solving the following simplified front equations :

together with equations (20), (21) and with the consistence condition :

The solution of equation (23c) is :

with

(11)

It is immediately seen that the continuity conditions (20), (21) together with the constraint

that £’ # 0 along the curved part except at its ends impose that q be real, 0

=

0 and :

That is, we find that the length c of each curved part is fixed. Note that, as expected, q is the

wavevector of the neutral mode (more precisely, of that with q larger than the critical value

qMS) of the classical MS problem with surface tension Fo in the q > qMS limit where space modulations of the diffusion field can be neglected.

From equation (26), and since A > 2 TT / q, our non-local approximation (22) is valid only

when

where Qcs = 2 D/Vcs ; V cs == V cs (G) = DG 1 1 mL dC 1 is the constitutional supercooling instability threshold, to which the MS one reduces when qms » 1, i.e. when the system is

highly non-local at the bifurcation. do

=

T M( U + U ">0+ IL 1 ML AC | 1 is the capillary length of

our problem. Note that condition (28) precisely guarantees that q > qMS, i.e. that the

approximate expression (26) of the wavevector of the neutral MS mode is justified.

So, we cannot predict exactly the position of the threshold which is only defined, in the present approximation, to within an uncertainty measured by (do/fcs)1I3. However, typically,

for chemical diffusion, do cannot exceed a few hundred angstrôms, while Vcs lies in the 1-

10 u/sec range, so that (do/fcs)1I3 is of order 10-1 1 at most. So, although our non-local

assumption forbids us to analyze the mathematical details of the « bifurcation », the restrictions that it implies should be of very little consequence for the analysis of experiments.

From equations (24), (25), one obtains :

We are thus left with the five equations (20), (23a, b) (27) relating the six unknowns À, a, b,

e (1), e (1), A -

That is, one of these quantities, which we choose to be À, remains a free parameter. The lengths and heights of the faceted regions of a profile of wavelength À are then given by :

with, of course,

(12)

Note that, from equations (30), (31), a :::. b, i.e. the hot (quasi-convex) facet is longer than

the cold (quasi-concave) one, and the average displacement of the deformed front from the infinite facet position (JI)

=

2 7T’F(V)/qÀjL is positive. These features of course result from the fact that, due to kinetic effects, the curved parts of the front grow more

«

easily

»

than the facets, which favors the hot against the cold facet parts of the crenellated profile.

The deformation amplitude, for a given À, is :

It is seen on equation (32) that (e, - e- ) is a decreasing function of A. In the range of

validity of our calculation (À 1 ), it remains larger than the limit value 2 F / (2 - g ). Since Dry’ - do LBu’ / f ( u 0 + o- 0 ") - 10- 4 LBu’ / u 0 this limit can be approached rather closely for

materials with slow facet kinetics. The amplitude defined by equation (32) diverges when À

tends towards its minimum value 2 TT /q, i.e. when the facets become vanishingly small. This

can be understood in the following way : when a, b -->+ 0, the profile reduces almost

everywhere to a cosine deformation, i.e. to the standard MS linear mode. However, the difference between our problem and the standard MS one lies in the fact that each zero-slope region of the profile, even if its length vanishes, costs a finite capillary energy oc 0 y’. When À --+ À min’ this plays the role of a finite driving term for the relevant MS mode, whose amplitude, since this mode is neutral, blows up. This divergence is of course an artefact of our

linearization of both the curvature and the diffusion terms in equations ((1), (3)). So, the

above results are valid only when non-linear corrections are negligible, i.e. when :

where al is the coefficient of the third order term of the amplitude expansion of the standard MS problem [2, 3]. Note that, due to the above mentioned smallness of 0 y’, this is not a very

stringent condition.

4. Linear stability of the crenellated front.

We now want to analyze the stability of the continuum of solutions determined above with respect to small perturbations. Modulations of such periodic structures can be classified into

amplitude and phase modes. Usually (i. e. , for the solidification problem, in the case of rough interfaces), phase modes correspond to space modulations of the wavelength À, while amplitude ones describe perturbations of the basic structure at constant À. Here, due to the presence of facets, a new class of

«

optical » phase modes should appear, corresponding roughly to modulations of the relative lengths of hot and cold facets at constant À. We will

only study here the amplitude modes, since it will appear that the crenellated structures are

unstable against such perturbations.

That is, we look for the time evolution of an infinitesimal deviation 8( (x, t ) from the stationary solution (x) = (0) + (l)(x), where be has the same wavelength and the same

facet lengths as 03BE (x). That linear eigenmodes satisfying these constraints exist is of course not obvious and, indeed, only true, as will appear in the following, in the non-local limit and when the facets are much longer than the curved front regions.

Again, we linearize the, now time-dependent, diffusion problem, and then repeat the procedure (Eqs. (12), (13)) leading to the quasi-linearized front equation. We moreover

assume, and will check later, that the quasi-static approximation [16], which amounts to

(13)

neglecting the eC/et term in equation (1), is valid. Setting Z(x, t) = e (1)(x) + 8( (x, t), one easily finds that the dynamical front equation which generalizes equation (14) now reads :

where :

and Zn (t )

=

azn/at.

We are looking for a pure amplitude mode, that is a deviation from the stationary solution

which leaves the average position of the front invariant, i.e. such that :

As appears in equations (33), (35), the dynamics of front deviations is controlled by two

mechanisms :

-

facet kinetics, which gives rise to the 2VY’ /2 term in equation (33),

-

diffusion, which gives rise to the non-stationary terms proportional to Zn

=

stn in

jé (Eq. (35)). It is of course necessary to compare the importance of this effect with that of the kinetic one, the strength of which varies widely with V and with the microscopic nature (e.g.

dislocation spirals, or 2-D nucleation) of the growth mechanism.

We have not been able to build a complete solution of the set of dynamical front equations generated by equations ((33), (35)), i.e. to get a general consistent solution for the infinite set of d03BEn (t ). We therefore limit ourselves to the simple case where the facet half-lengths a, b are

much larger than the x-extension 7r /q of the curved regions of the front. In this case, the contribution of these curved regions to the diffusion term can safely be neglected. That is, in

jé, Z can be approximated by :

where, in order for (36) to be satisfied :

That is :

o o

Plugging expression (39) into equation (33), neglecting, for the same reason as in the

stationary case, the Zn terms in fC, integrating equation (33) along the hot facet, and

subtracting out the stationary facet equation (23b), we obtain :

(14)

with :

The same procedure, when applied to the cold facet, generates an equation for d03BE_ (t ). It is easily checked that, when use is made of equation (38), this degenerates with equation (40).

That is, to lowest order in the ratios 7T /qa, 7T /qb, we find that the amplitude mode is

described by :

where :

F ( V ) is proportional to the kinetic undercooling (Eq. (4)), that is increases with V. So, w (Eq. 43)) is always positive, and we find that the stationary crenellated states are unstable

against amplitude fluctuations.

Of course, the above calculation of the amplitude growth rate is only a lower order

estimate, since it ignores completely the problem of the dynamical shape of the curved parts, which would come into play at the next step of an expansion in 7T /qa, 7T /qb, and would certainly lead, in a more elaborate calculation, to small variations of the facet lengths.

5. Conclusion.

Inspite of its limitations (small amplitude and strong non-locality approximations) the above analysis permits to draw a first picture of the destabilization of a faceted directional solidification front.

Two main results emerge from our calculation :

(i) small amplitude crenellated stationary solutions, characteristic of faceted fronts, exist beyond a threshold V

=

V cs (G ) . (1 + o «do/fcs)1I3» which is, in practice, the usual MS

one.

There is, for each value of V beyond this threshold, a continuum of such states. Our

approximations, although they do not permit to locate precisely the limits Àmin, Àmax of this continuum, enable us to determine the front shapes for space periods À such that (in physical variables)

(ii) these crenellated solutions are unstable against small amplitude fluctuations. This entails that, for V

>

Vcs, the planar front is unstable against finite amplitude deformations.

These, of course, cannot be treated within our approximations, but it can reasonably be guessed that the corresponding stable branch of stationary states corresponds to globally

cellular fronts with faceted regions, analogous to what is observed on faceted dendrites [17].

(15)

The crenellated solutions must be

«

jumped over

»

by the front configuration in order for cellular shapes to be reached. So, we predict that the facet to cell transition should exhibit a

very large velocity hysteresis.

Whether or not this transition can be termed a bifurcation is not clear to us, since no

linearization of the basic facet state is legitimate. Our calculation does not predict any upper

velocity threshold for crenellated states. However, we cannot describe the high velocity regime since, when f

=

2 D/V decreases, our non-local approximation fails. In particular, we

cannot exclude that, in our model, there might be no such upper threshold. However, we have assumed here that the facet is perfect, i.e. that the Wulff plot exhibits a perfect cusp whatever the velocity. This approximation becomes incorrect at high growth rates, where the

cusp is rounded-off by kinetic roughening [12]. In real systems, this effect will always come

into play to produce a cross-over between the regime analyzed here and a high-V one where

the standard MS analysis is, again, valid, giving rise to an ordinary inverted bifurcation.

Clearly, the deeper the equilibrium cusp, i. e. the larger the step energy relative to the facet, the higher the velocity at which this cross-over occurs.

It is of course essential at this stage that systematic quantitative experimental studies of

directional solidification morphologies of transparent faceting materials be developed. On the

basis of the present study, we believe that they should aim primarily at the observation of the

anomalously large hysteresis predicted above.

Let us also remark that, as appears on equation (43), large values of VF’ (V ) lead to small amplification rates for amplitude fluctuations. That is, as is reasonable, strongly

«

locked »

facets correspond to slow amplitude dynamics. So, it might not be excluded that, in strongly faceting materials, crenellated states - possibly triggered by an external perturbation - be

observable as transients.

References

[1] MULLINS W. W., SEKERKA R. F., J. Appl. Phys. 35 (1964) 444.

[2] WOLLKIND D. J., SEGEL L. A., Philos. Trans. R. Soc. London 51 (1970) 268.

[3] CAROLI B., CAROLI C., ROULET B., J. Phys. France 43 (1982) 1767.

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

[5] See for instance : KESSLER D. A., LEVINE H., Preprint INLS 1007 and references therein.

[6] SAITO Y., MISBAH C., MÜLLER-KRUMBHAAR H., Proc. of the 3rd UC Conf. on Statistical Mechanics Univ. Calif. Davis (1988) Ed. C. Garrod, to be published.

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

[8] MCFADDEN G. B., CORIELL S. R., SEKERKA R. F., J. Cryst. Growth 91 (1988) 180.

[9] This excludes the case where facetted growth is controlled by kinetics only (see UWAHA, M., J.

Cryst. Growth 80 (1987) 84).

[10] CHERNOV A. A., NISHINAGA T., Morphology of Crystals, Part A, Ed. I. Sunagawa (Terra Scient.

Publ. Co.) 1987, p. 207.

[11] BEN AMAR M., POMEAU Y., Europhys. Lett. 6 (1988) 609.

[12] We assume here that the melting temperature is far enough below the roughening one for us to neglect any dynamical roughening effect (see NOZIERES P., GALLET F., J. Phys. France 48 (1987) 353).

[13] NOZIERES P., Cours au Collège de France (1984), unpublished ;

See also CAHN J. W., HOFFMAN D. W., Acta Metall. 22 (1974) 1205.

[14] TILLER W. A., The Art and Science of Growing Crystals, Ed. J.J. Gilman (J. Wiley & Sons, New

York) 1966.

(16)

[15] Note that, in the highly non-local limit we will restrict to here (03BB ~ 1), this is a much more severe

limitation for our one-sided model than for the symmetric one (see LANGER J. S., TURSKI

L. A., Acta Metall. 25 (1977) 1115). Indeed, it can be checked that the ratio of the coefficients a1 of the cubic terms in the amplitude expansion are such that, in this limit (see Ref. [3]) :

aOS1/aSym1 ~ 03BB.

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

[17] MAURER J., BOUISSOU P., PERRIN B., TABELING P., Europhys. Lett. 8 (1989) 67.

Références

Documents relatifs

Zuazua study the long time behaviour of a homogeneous version of (1); we also refer the interested reader to [11], in which M. Zuazua extend the analysis performed in [12] to the

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

Abstract 2014 We study, for directional solidification or fusion of binary mixtures, the stability of planar fronts against one-dimensional deformations.. We extend the

In section 4, we focus on the one-dimensional space case, and prove theorem 2 using the lemma of Matano about the number of sign changes.. 2 Existence of

Typical examples are three-dimensional red blood cells whose boundaries minimize the second-order Helfrich energy [ 15 ], two-dimensional soap films which are connected solutions to

For waves of long wavelength, the negative signature of the action Hessian is found to be exactly governed by M 00 (c), the second derivative with respect to the wave speed of

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

Figure 6 shows two typical recoil curves measured at the same pulling velocity in two different samples of undoped CBr4.. These curves have been corrected for the instrumental