• Aucun résultat trouvé

Cellular instabilities in directional solidification

N/A
N/A
Protected

Academic year: 2021

Partager "Cellular instabilities in directional solidification"

Copied!
10
0
0

Texte intégral

(1)

HAL Id: jpa-00210403

https://hal.archives-ouvertes.fr/jpa-00210403

Submitted on 1 Jan 1986

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.

Cellular instabilities in directional solidification

S. de Cheveigné, C. Guthmann, M.M. Lebrun

To cite this version:

S. de Cheveigné, C. Guthmann, M.M. Lebrun. Cellular instabilities in directional solidification. Jour-

nal de Physique, 1986, 47 (12), pp.2095-2103. �10.1051/jphys:0198600470120209500�. �jpa-00210403�

(2)

Cellular instabilities in directional solidification

S. de Cheveigné, C. Guthmann and M. M. Lebrun

Groupe de Physique des Solides de l’Ecole Normale Supérieure, Université Paris VII, Tour 23, 2 Pl. Jussieu, 75251 Paris Cedex 05, France

(Requ le 3 juillet 1986, accepté le 28 aoat 1986)

Résumé.

2014

Nous avons étudié la morphologie de l’interface solide-liquide lors de la solidification directionnelle d’échantillons minces (5-150 03BCm) d’un alliage binaire dilué (CBr4/~ 0,1 % Br2). Au-dessus

d’une vitesse de tirage critique, à concentration de soluté et gradient de températures données, l’interface qui

était plane au départ, présente une déformation cellulaire périodique. La bifurcation d’un front plan à un front

cellulaire est sous-critique. La vitesse de tirage critique est fortement dépendante de l’épaisseur de

l’échantillon : on ne peut pas considérer le problème comme bi-dimensionnel. Nous avons également mesuré

les longueurs d’onde au seuil et au-dessus ainsi que les temps de réponse. Nous comparons ces résultats aux

prédictions théoriques existantes.

Abstract.

2014

We have studied the morphology of the solid-liquid interface during the directional solidification of thin (5-150 03BCm) samples of a dilute binary mixture (CBr4/~ 0.1 % Br2) . Above a critical pulling speed,

for a given temperature gradient and solute concentration, the interface, initially planar, breaks down into a

periodic cellular pattern. The bifurcation from a planar to a cellular solidification front is sub-critical. The critical pulling speed is strongly dependent on the thickness of the sample : the problem cannot be considered

two-dimensional. We have also measured wavelengths at, and above the threshold and examined response times. These results are compared to existing theoretical predictions.

Classification

Physics Abstracts

64.70D

-

81.30D

1. Introduction.

The existence of morphological instabilities [1] defor- ming the solid-liquid interface of directionally solidi-

fied dilute binary alloys is a phenomenon well

known to metallurgists [2]. Samples are pulled at a

certain velocity V in a fixed temperature gradient G

and are thus progressively solidified. If the equili-

brium concentration of the solute is different in the

liquid and in the solid, an excess (in the case of the compounds studied here) or a lack of solute builds up in front of the moving interface and must be evacuated to allow solidification to progress. Above

a threshold pulling speed, for given temperature gradient and concentration, the interface becomes unstable and breaks down into a regular, periodic pattern (Fig. 1).

Metallic alloys, because of their opacity, are not

the most practical systems on which to study the

behaviour of such instabilities : samples have to be quenched or decanted to allow the observation of a

solid-liquid interface which may be affected by this

treatment. The study of thin samples of transparent organic compounds, first suggested by Jackson and

Hunt [3] in 1966, allows direct observation, under

the microscope, of the dynamics of the solid-liquid

interface during solidification. Some such

Fig. 1.

-

Well-developed cells (CBr4’ V = 3V+c, .

A - 50 JLm).

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198600470120209500

(3)

2096

compounds, in general plastic crystals, present a rough interface which does not facet, and in that respect they behave like metals. The most frequently

studied are tetrabromomethane CBr4 and succi-

nonitrile (CH2CN ) 2.

In the present study we have tried to pinpoint the

various factors which affect first the onset of cellular

instabilities, then the periodicity of the resulting pattern of interface deformation, in thin (5-150 pm) samples. Most of the experiments were performed

with CBr4, but the results were generally checked

with succinonitrile, both compounds being used as supplied. Impurity concentrations were therefore

fixed, pulling speeds, temperature gradient, and sample thicknesses were varied. We did not consider the problem of dendrite formation during directional

solidification, extensively studied in transparent organic materials [4-6]. This phenomenon takes place well above the threshold at which morphologi-

cal instabilities first appear (above about 10 times

the threshold pulling speed, for a given temperature gradient and impurity concentration). On the

contrary, most of the work described here was done to gain better knowledge of the behaviour of the system at, or near, the threshold.

2. Theory.

A number of authors have attempted to give a dynamical description of the phenomena occuring during directional solidification of dilute binary alloys in thin samples, when the system is definitely

out of equilibrium. The linear stability of the planar

solution for the solid-liquid interface was originally

studied by Mullins and Sekerka [7] in 1964. They

described the balance between the destabilizing

effect of the accumulation of solute in front of the

solid-liquid interface and the stabilizing effects of the temperature gradient and of the surface tension.

Wollkind and Segel [8] reformulated this analysis

and went beyond, giving a non-linear stability analy-

sis using perturbation methods valid only close to the

threshold. Caroli, Caroli and Roulet [9] extended

this analysis to the case where diffusion in the solid

phase is not negligible. A number of numerical

calculations have also been performed [10-21]. All

these theories tacitly suppose that the sample may be considered infinite in the direction perpendicular to

the plane of the sample, i.e. that the deformations

are one dimensional along the interface. We shall

see below that this is not the case.

Let us recall some of these results. Diffusion

equations for heat and solute are written in both the solid and the liquid phases. The interface conditions

are the source of the non-linearity of the problem,

because the temperature and concentration fields

depend non-linearly on the shape of the interface. In the linear stability analysis, first performed by Mul-

lins and Sekerka [7], a small sinusoidal perturbation

is superimposed on a planar interface, and the

conditions under which it is amplified or damped are

determined. At low velocities the equations defining

neutral stability can be expanded in terms of the

wavenumber, which is also small, and one in fact obtains, to the lowest order, a modified (1) version

of what is known as the « constitutional supercooling

criterion » [22] :

where G, is the critical temperature gradient, Vc the

critical pulling speed, Coo the mean solute concentra- tion, m the liquidus slope, k the partition coefficient and DL the solute diffusion coefficient in the liquid.

The ratio v

=

V C 00 /G appears as the control parameter.

Under the conditions of the present experiments,

the second order correction to relation (1) is at most

of a few percent. It can be important, particularly at higher pulling speeds [23, 24].

To the same order of approximation, one can also

write the wavelength A MS of the « critical mode »

(i.e. the first neutrally stable mode encountered when the control parameter is increased) :

where q is characteristic of the material :

y is the solid liquid surface tension, C the latent heat,

p the density. The expression for A ms is obtained

assuming an infinitesimally small amplitude perturba-

tion.

A non-linear analysis can describe analytically the

way in which an initially instable perturbation evol-

ves close to the threshold, using an « amplitude equation » :

The various possible evolutions are discussed in references [8] and [9]. Under the present experimen-

tal conditions (low velocity, a partition coefficient k 0.45 and a temperature distribution decoupled

from the instability because it is imposed from the

outside by the glass cell [25]) a subcritical bifurcation is predicted - and as we shall discuss below, it is indeed observed. This means that the’ amplitude of

the deformation is never infinitesimal so the ampli-

tude equation cannot be used to make further predictions:

One would nevertheless like to have predictions concerning the cell shape and the wavelength at

(1) Mullins and Sekerka’s result replaces the tempera-

ture gradient by the average of its values in the solid and in the liquid. In the present experiments the gradient is imposed externally and has the same value in the solid and in the liquid. In that case, as in the case of equal thermal

conductivities in both phases, Mullins and Sekerka’s result

reduces to the « constitutional supercooling criterion ».

(4)

velocities above the threshold. To answer this ques-

tion, a certain number of numerical calculations have been performed. The first models relied on the

amplitude equation and the hypothesis of infinitesi-

mally small deformations. Langer [10,11] developed

a « symmetrical » model - equal solute diffusion

coefficients in the solid and in the liquid

-

which, although it was not very realistic, gave a qualitative picture of the phenomena. For example, with this model, Dee [12], assuming that pattern selection

takes place by propagation along the front of the

« wall » between two regions of different cellular

wavelengths, finds a well-defined wavelength that

decreases with increasing velocity. On the other hand he does not predict cell shapes.

Using a more realistic « one-sided » model

-

which neglects solute diffusion in the liquid

-

but

still using the amplitude equation, Kerszberg [13-15]

derived the equation of movement of a deformed

front. He obtained quite a good representation of

the shape of a cellular interface. The wavelength was

not uniquely determined but is restricted to an

interval.

To approach the problem of wavenumber selec-

tion, Kerszberg studied the effect of Gaussian white

noise on the dynamics of the interface, again by a

numerical simulation. A unique wavevector is selec- ted in this manner. The noise induces the disappea-

rance or the division of cells in a manner very similar to what is observed experimentally (compare Fig. 4

of Ref. [15] to Fig. 4 of the present article). The

value of k « selected » by the noise is given as a

function of the constraint parameter v a V/DL G.

Unfortunately his calculation does not go far beyond

the threshold ( A v / v == 5 X 10-3) . The selected

wavelength increases with increasing pulling speed,

at a given temperature gradient.

McFadden and Coriell [16] apply a one-sided

model to an Al-Ag alloy, without using the ampli-

tude equation. They use the solute concentration

Coo as the control parameter. (It should be noted that for the experimentalist, V is a more « natural »

parameter since it can be varied rapidly and easily).

Their predictions concerning the wavelength are particularly pertinent to the case of a subcritical bifurcation : a relatively large amplitude perturba-

tion of wavelength smaller than that of the most unstable linear mode can develop. It appears explici- tly, as expected, that the critical wavelength calcula-

ted by the linear analysis is not valid in the case of a

subcritical bifurcation. Above the threshold concen- tration C * , the wavelength evolves towards smaller values. When values larger than the most unstable

linear wavelength A * are tested, they are shown to decay and to be replaced by their first harmonic at half wavelength which is smaller than A *.

Ungar and Brown [17-20] also avoid the amplitude equation. They use -a one-sided mode and suppose

equal thermal conductivities in the solid and in the

liquid. They work with the temperature gradient as a

control variable at fixed pulling speed (once again,

these are not the most practical experimental condi-

tions). Using finite element methods, they predict

the presence of two successive bifurcations, the first

to a deformed interface with a periodicity À, the

second to a periodicity A/2, but the second bifurca- tion is predicted to be so close to the first (AV/V - 191,,) as not to be visible experimentally.

The predictions of Ungar and Brown [20] concer- ning deep cell morphology well above the threshold

are particularly interesting. They are based on the hypothesis that the wavelength does not change with increasing constant

-

a hypothesis contrary to expe- rimental results. Nevertheless, the evolution of cell

shape is very well described (compare for example Fig. 10 of Ref. [20] to Fig. 1 of the present article- both are at V - 3-5 Vc.)

Very recently, Karma [21] has applied a boundary

model to simultaneously obtain the shape and wave- length of well developed cells. This work is still in progress, but the preliminary results seem promising.

3. Experimental.

Most of the work presented here was done with

tetrabromomethane CBr4 . Some experiments

were also carried out on succinonitrile (CH2CN) 2’

for comparison. Succinonitrile has been extensively

characterized both when pure and mixed with ace- tone, by Glicksman [4]. Table I gives data concerning CBr4. We use the compound as provided by Fluka,

Table 1.

-

Physico-chemical constants for CBr4.

without further purification. The main impurity is given as Br2. We performed Differential Scanning Calorimetry measurements on the material. This

provided us with a curve of the energy absorbed

versus temperature during melting. By fitting succes-

sive melted fractions (from 1/8 to 1/3) versus

temperature to the van t’Hoff formula, we obtained

an impurity concentration Coo

=

0.12 ± 0.02 mol. %,

a liquidus slope m

=

2.9 ± 0.2 K/mol. % and a parti-

tion coefficient k

=

0.16 ± 0.01. A melting tempera-

ture of 93.3 ± 0.2 °C for the pure compound was calculated. The phase diagram for Br2 in CBr4 reported in reference 26 gives m

=

2.73

± 0.2 K/mol. % with TM

=

92.5 °C. We also perfor-

(5)

2098

med the same measurements and calculations on

succinonitrile (Koch-light and Eastman) ; they gave Coo

=

0.45 ± 0.05 mol. %, m

=

2.20 ± 0.05 K/mol.

% and k

=

0.12 ± 0.01. The melting temperature for the pure compound was calculated as 58.1 ± 0.3 °C.

The diffusion coefficient for Br2 is missing but we

shall estimate below a value of 1.2 ± 0.4 x

10- 5 cm2/s.

The pulling stage is essentially similar to that proposed by Jackson and Hunt [3]. Cell preparation

is described in reference [25] ; sample thicknesses

are measured to within ± 2-3 pm and are uniform to within ± 5 pm. The temperature of the sample is imposed by the glass cell which is a far better conductor of heat than the sample itself. Particular

care was taken to avoid residual vertical temperature gradients which strongly affect the experimental

results (see below).

The experiments are video-taped which allows the

pulling speed to be measured precisely (± 0.1 Rm/s).

Wavelengths are averages over at least 10 cells.

4. Results and discussion.

Before presenting our experimental results, we

believe that some preliminary remarks are necessary.

First, when studying dynamical phenomena, where

fluctuations play such an important role, a certain

amount of dispersion in the results is inevitable.

Another point is that all the important parameters of the problem have not yet been isolated and can

easily be missed

-

this was the case of the sample thickness, the effect of which is decribed below. Just

as in the case of hydrodynamic instabilities, nume-

rous and repeated experiments must be performed

before definite conclusions can be drawn.

4.1 RESPONSE TIMES.

-

In the present experiments

we have tried to observe the solid-liquid interface in

a stationary state and this can take a long time. To

illustrate this we have plotted the time it takes the interface to begin to deform visibly after the pulling speed is suddenly changed from a zero value (the

time it takes to reach a steady state is even longer).

Figure 2 shows the curves of T versus the final speed

V for CBr4 ; the points we obtain for succinonitrile fit the same curve, as do most of the response times

reported by Somboonsuk et al. [27] for a succinonitri- le-acetone solution. The final speed is the pertinent parameter since it is at that speed that the concentra-

tion profile adjusts.

The response time in fact contains two terms. The

first, a solute diffusion term, is roughly the time it takes to establish the steady state concentration

profile after begining to solidify a sample. This can

be estimated as [22] :

The second term is due to the thermal response of

Fig. 2.

-

Response time versus pulling speed (see text).

the glass cell and is expected to be preponderant at higher pulling speeds. To explain this, the cell can be

considered as a heat conducting bar moving at velocity V across the two heat reservoirs (Fig. 3a). In

the laboratory frame, the temperature 7y at z satisfies, in the stationary state, the equation :

where DG is the thermal diffusivity of glass. With Ty(O) = T, and Ty(L) = T2, the solution is :

Fig. 3.

-

a) Schematic representation of the cell moving at velocity V across the heat reservoirs. Tl > T M > T2;

b) Temperature profiles for V

=

0 and V > 0.

(6)

i.e. the temperature profile is no longer linear as it is

when V

=

0 (Fig. 3b) and the average temperature of the cell has increased. The term in the response time which corresponds to the time it takes the cell to fit the new thermal conditions is preponderent at high pulling speeds, It is the value at which T levels off: about 10 s in the present case. The thermal component is negligible at speeds lower than

~

10 f.L/s. When that term is subtracted from the

experimental values of T (Fig. 2) the result varies as

V-2 . The proportionality factor is

-

2-3 x 10-5

cm2/s which is less than the value of D/k estimated

as 1.2 x 10-5/0.16

=

7.5 x 10-5 cm2/s. One must

however remember that T is not sufficiently precisely defined, experimentally or theoretically, to allow

any quantitative conclusions to be drawn.

The important result is that at pulling speeds of

10 R/s, one must wait about 60 s for the system to react, at 1 R/s about 30 min and extrapolating to

0.1 ii/s, about 16 hours ! Such times make extensive and repeated studies extremely difficult. That is why

we found it preferable to work under conditions (in particular at low concentrations) where critical speeds are above 1 R/s.

4.2 A SUBCRITICAL BIFURCATION.

-

We have

previously shown [25] that the bifurcation from a

planar to a cellular morphology of the interface is

subcritical, i.e. the critical pulling speed V+c for increasing speeds is greater than the critical value

Vy for decreasing speeds. Figure 4 shows the ampli-

tude of the deformations of the interface (measured

from the cell tip to the point where the cell sides

meet) versus the pulling speed. Above V’ a finite amplitude deformation sets in. As mentioned above,

the non-linear analysis of the problem [8, 9] predicts

this result which is of practical consequence for calculations. It means that the amplitude of the

deformations is never infinitesimally small, so the amplitude equation cannot be used above the thres- hold. Note that the threshold values are somewhat lower than those given in reference [25] : this is due to the presence of a residual vertical gradient in the

Fig. 4.

-

Amplitude of the periodic deformation versus

pulling speed (e

=

30 J..Lm, G =120°/cm). Full segments : increasing speed. Dashed segments: decreasing speed.

Curve : interpretation in terms of a subcritical bifurcation.

earlier experiments. The vertical temperature gra- dient produces an inclined interface which favorizes the macroscopic evacuation of solute along the

bottom of the cell, since the concentration gradient

has a component parallel to the interface. This process, also observed with succinonitrile, is suffi-

cient to retard the ’destabilization of the interface.

According to relation (1), the critical speed predic-

ted by the linear theory (V ’ in the notation of the

previous paragraph) should be proportional to the temperature gradient G, with a slope that is a

function of the diffusion coefficient DL’ We checked (for relatively thick samples - see explanation below) that the linear relation holds and used it to estimate DL = 1.2 ± 0.4 x 10- 5 cm2/s. This value is

quite reasonable and we shall use it for further calculations.

4.3 WAVELENGTH SELECTION.

-

Our experimental

values for the wavelength versus the pulling speed

are shown in figure 5, for two different temperature gradients. (The samples considered were thicker than 50 Rm to allow comparison with the thresholds

predicted by relation (1) - see explanation below).

We have drawn on the same figure the neutral stability curve deduced from the linear analysis

described above [7, 9]. Periodic deformations of

wavelengths within the band delimited by the curve

will not a priori relax, whereas those outside the band will, but no clear selection mechanism within the band has been described theoretically. Experi- mentally a wavelength is selected, to within a few

tens of percent (the shaded area on the curve shows

the maximum experimental dispersion). The value

Fig. 5.

-

Wavelength versus pulling speed for two

different temperature gradients. The full line is the calculated neutral stability curve ; the points are experi-

mental. The shaded strip gives the experimental disper-

sion. (e > 50 03BCm; (a) and full circles : G= 120 7cm;

(b) and empty circles : G

=

70*/cm).

(7)

2100

of the critical wavelength A ms predicted by the linear analysis is shown for both values of the temperature gradient. It is 2 to 3 times greater than the threshold value A c observed experimentally (180 pm versus 75-

80 pm for G

=

70*/cm, 125 R versus 50-55 pm for G = 120 7cm). This tends to confirm the existence of a subcritical bifurcation since the prediction made

under the hypothesis of infinitesimal deformations does not hold. The fact that we observe a critical

wavelength A c - A ms/2 may mean that we are

indeed observing the second bifurcation predicted by Ungar and Brown [16, 20].

The wavelength easily adjusts by birth or death of

cells (Fig. 6) in a manner very similar to that found

numerically by Kerszberg [15].

Above the threshold A varies roughly as V-’/2 (we find a functional dependence of V - o-’ ± 0. 1). This

dependence is consistent with predictions by

Karma [21]. For a symmetric model the system should evolve at constant

where 1

=

2 DL/V is the diffusion length and do

=

’Y T M/ Lco m ( 1- k ) is the capillary length. In the

present case, one can write

u

=

6.6 x 10- 4/A 2 V (A in J.Lm, V in Rm/s) .

We find u - 5 - 6 X 10-3; Karma predicts

at

=

3.1 x 10-2 (symetric model, k = 1).

Experimentally, a , monotonous decrease in À ( V ) as V-1/2 or V-1/4 has often been reported

in literature, generally for three-dimensional metal

alloy samples (see reviews by Hunt [29] and Klaren

et al. [30]). It is not always clear at what distance above the threshold the experiments are performed.

For speeds above = 3 Vc’ Jin and Purdy [31] report a

Fig. 6.

-

Wavelength adjustment by birth and death of cells. Time increases from left to right.

h oc V - 1,,14 behaviour in a Fe-8 wt % Ni alloy. An exception to this monotonous variation was found in Pb alloys [30, 32] where a very slight increase was

first observed, followed by a decrease at higher pulling speeds. Trivedi [5] has reported a pronoun-

ced maximum in À ( V ) for a succinonitrile-acetone system, but the region studied is far above the threshold predicted by the constitutional supercoo-

ling criterion. The question of how the A ( v ) curves

near the cellular threshold and in the dendritic

region join each other is quite open, both theoreti-

cally and experimentally.

Relation (4) predicts no direct dependence of A on

the temperature gradient because the latter is not yet introduced into the model. We find A slightly lower

for the lower temperature gradient at a given speed (Fig. 5) but the shift is barely greater than our experimental dispersion. This tendency is reported

in the literature but at speeds well above the threshold. Various dependences on the temperature gradient G have been reported for metal alloys,

from G- 1/3 for Pb alloys [30, 32] to G-1 in Al alloys [33] (see Ref. [29] and [30] for a review of literature).

According to Mason et al. [32] the variables V and T

are not separable, and the exponent of G varies with V.

More work, nearer the threshold, and concerning

various systems, is clearly necessary to confirm the

experimental picture. We insist once again on the importance of doing experiments sufficiently slowly : if a pulling speed above the critical value is

applied suddenly a transient pattern of wavelength

smaller than the steady state wavelength is observed.

This effect, which has been mentioned by various

authors [27, 33-35], can be a source of error in

measurements of À, particularly in the case of high

response times, i.e. low pulling speeds.

4.4 EFFECT OF SAMPLE THICKNESS.

-

Our most

unexpected result is the effect of the thickness of the

sample. Samples thicknesses are of the order of tens of microns, compared to lengths and widths of

several centimeters, and the problem is usually

treated theoretically as two dimensional. In fact, sample thickness is an essential parameter : the solid liquid interface becomes less and less stable when the sample thickness decreases. Figure 7 gives the

dependence we observe for the critical pulling speed V; versus sample thickness, for two temperature gradients (70 and 120 °Icm). We have observed a

similar effect with succinonitrile.

Convection in the liquid could have been a

possible explanation for this effect. We have checked that none takes place, at least on a time

scale comparable to that of the phenomena under study, by introducing small (5 pm) particles of silica (of roughly the same density as CBr4). The particles

are definitely mobile since in one case we were able to observe their movement induced by Marangoni

effect on the liquid-gas interface of a bubble.

We believe that the solid-liquid meniscus is the

cause of this strong dependence of the critical pulling

speed on the sample thickness. As drawn in figure 8,

(8)

Fig. 7.

-

Critical pulling speed V+c versus sample

thickness for two different temperature gradients : a) G =120°/cm ; b) G

=

707cm.

Fig. 8.

-

A schematic representation of the solid-liquid

interface.

the interface is curved to assure mechanical equili-

brium at the point of contact with the glass. The

contact angle depends on the pulling speed because

the solute distribution along the interface varies.

The shape of the interface adjusts to assure local thermodynamic equilibrium (Gibbs-Thomson rela-

tion ; see Ref. [7, 8 or 9]). The meniscus is not easily

visible in a true sample, but figure 9 shows the one

formed by the solid-liquid interface of succinonitrile

on the side of an optical cell (on a polished glass surface). The meniscus affects a distance of about

50 tL which is of the order of the sample thickness

range over which the threshold pulling speed is seen

to vary.

The presence of the meniscus, which deforms the interface in a manner that depends on the pulling speed, should have an effect similar to that of grain

boundaries described by Voorhees et al. [28] and by Ungar and Brown [18], i.e. the threshold is shifted to lower values. The description of the interface is in itself a non-linear problem ; one has then to treat

Fig. 9.

-

Solid-liquid meniscus in succinonitrile in an

optical cell (full picture height 260 J.Lm).

the coupling of deformations in perpendicular pla-

nes. Caroli, Caroli and Roulet [37] have calculated

the meniscus analytically to the third order and then

performed a linear stability analysis of the solid- liquid interface (in a symmetric model). This calcula- tion is valid only. for a very weakly curved meniscus,

but it shows qualitatively that the effect of the curvature is to decrease the critical pulling speed for

thicknesses smaller than the cellular wavelength ;

the shift increases as thickness decreases.

A similar effect may well be responsible for the

unusual tip velocity

-

tip curvature relation obser- ved by Sawada and co-workers [38] for dentritic growth in very thin (5 J.Lm) samples.

Because of this dependence of the critical pulling speed on the sample thickness, one must be cautious in comparing with the value given by the « constitu- tional supercooling criterion » (Eq. (1)). The rela-

tion is expected to apply to thick samples for which

the effect of the meniscus is weak. It is for this

reason that we used the value of V+c at which V+c (e) levels off to calculate DL and that we drew figure 5 for samples with thicknesses in the plateau region.

Note that the wavelength of the cellular deforma- tion of the solidification front gives the length scale

to which the thickness of the sample should be compared. Since the wavelength varies with the

pulling speed, a sample is « thick » or « thin » depending on the speed at which it is being pulled.

In the calculation performed by Caroli et al. [37] of

the effect of the sample thickness on the critical pulling speed, ÀMs is the scale which‘ appears. In fact the experimentally observed value Àc is probably the appropriate scale. For example the curve of V’

versus e given in figure 7a intersects the curve of A versus V (given in a log-log diagram in Fig. 5) at

about e

=

A

=

50 pm. It is in that same region that

it saturates.

We have found that the samples become definitely

« thick », and present several layers of cells, once

the wavelength becomes smaller than 1/2 to 1/3 of

(9)

2102

Fig. 10.

-

Multiple layers of cells in a thick sample (e = 170 ttm, k

=

35 fJ.m, V

=

45 Rm/s). a) Focusing on the top of the sample ; b) Focusing on the bottom of the sample. A vertical temperature gradient staggers the layers. Note the

effect of cristal orientation observed at high pulling speeds.

the sample thickness. At that point, several layers of

cells are observed (Fig. 10) and A is larger than expected according to the relation A 2 V = constant.

Apparently, the wavelength is not thickness dependent. This is somewhat surprising since one

would expet the meniscus to increase in some way the « effective » curvature of the interface and

perhaps cause A to decrease. If such an effect exist, it

is weak. To a first approximation, Caroli et al.

predict theoretically that A MS will not be affected.

5. Conclusion.

In conclusion, we have studied the nature of the bifurcation from a planar to a cellularly deformed

interface and have found it to be subcritical, as

predicted by the non-linear analysis of the problem.

This result is confirmed by the values found for the critical wavelength. Response times are shown to

become very long at low pulling speeds and to be a possible source of experimental error. We have

discovered that the critical pulling speed is strongly dependent on the sample thickness. This last result should be stressed : the sample cannot be considered two-dimensional.

Acknowledgments.

It is a pleasure to thank B. Caroli, C. Caroli, B.

Roulet and C. Misbah for frequent and stimulating

discussions.

References

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

[2] See for example the special issue of Materials Science and Engineering on Solidification Microstructure 65 (1984).

[3] JACKSON, K. R. and HUNT, J. D., Trans. Met. Soc.

AIME 236 (1966) 1929.

[4] GLICKSMAN, M. E., SHAEFER, R. J. and RYERS, J.

D., Metall. Trans. A 7 (1976) 1747.

[5] SOMBOONSUK, K., MASON, J. T., TRIVEDI, R., Metall. Trans. 15A (1984) 967 ; TRIVEDI, R., Metall. Trans. 15A (1984) 977.

[6] ESAKA, H. and KURZ, W., J. Crystal Growth 72 (1985) 578.

[7] MULLINS, W. W. and SEKERKA, R. F., J. Appl. Phys.

35 (1964) 444.

[8] WOLLKIND, D. and SEGEL, L., Philos. Trans. R. Soc.

268 (1970) 351.

[9] CAROL, B., CAROLI, C. and ROULET, B., J. Physique

43 (1982) 1767.

[10] LANGER, J. S. and TURSKI, L. A., Acta Metall. 25

(1977) 1113.

[11] LANGER, J. S., Acta Metall. 25 (1977) 1121.

[12] DEE, G., Physica 12D (1984) 270.

[13] KERSZBERG, M., Phys. Rev. B 27 (1983) 3909.

[14] KERSZBERG, M., Phys. Rev. B 27 (1983) 6796.

[15] KERSZBERG, M., Physica 12D (1984) 262.

[16] FADDEN, G. B. and CORIELL, S. R., Physica 12D (1984) 253.

[17] UNGAR, L. H. and BROWN, R. A., Phys. Rev. B 29 (1984) 1367.

[18] UNGAR, L. H. and BROWN, R. A., Phys. Rev. B 30 (1984) 3993.

[19] UNGAR, L. H. and BROWN, R. A., Phys. Rev. B 31 (1985) 5923.

[20] UNGAR, L. H. and BROWN, R. A., Phys. Rev. B 31 (1985) 5931.

[21] KARMA, A. to be published.

[22] TILLER, W. A., JACKSON, K. A., RUTTER, J. W., CHALMERS, B., Acta Metall. 1 (1953) 428.

[23] SATO, T., OHIRA, G., J. Crystal Growth 40 (1977)

78.

[24] SATO, T., ITO, K. and OHIRA, G., Trans: JIM 21

(1980) 441 and 449.

(10)

[25] DE CHEVEIGNÉ, S., GUTHMANN, C., LEBRUN, M.

M., J. Crystal Growth 73 (1985) 242.

[26] BILTZ, W. and MEINECKE, E., Z. phys. Chem. 131 (1923) 1.

[27] SOMBOONSUK, K. and TRIVEDI, R., Acta Metall. 33

(1985) 1051.

[28] VOORHEES, P. W., CORIELL, S. R., MCFADDEN, G.

B. and SEKERKA, R. F., J. Crystal Growth 67 (1984) 425.

[29] HUNT, J. D. in Solidification and Casting of Metals (The Metals Society, London) 1979.

[30] KLAREN, C. M., VERHOEVEN, J. D., TRIVEDI, R., Metall. Trans. 11A (1980) 1853.

[31] JIN, I. and PURDY, G. R., J. Crystal Growth 23 (1974) 37.

[32] MASON, J. T., VERHOEVEN, J. D. and TRIVEDI, R.,

J. Crystal Growth 59 (1982) 516-524.

[33] MIYATA, Y., SUZUKI, T., UNO, J.-I., MIYATA, Y., SUZUKI, T., Metall. Trans. 16A (1985) 1807.

[34] SHARP, R. M. and HELLAWELL, A., J. Crystal

Growth 6 (1970) 334-340.

[35] SHARP, R. M. and HELLAWELL, R., J. Crystal

Growth 11 (1971) 77-91.

[36] MORRIS, L. R., WINEGARD, W. C., J. Crystal

Growth 5 (1969) 361.

[37] CAROLI, B., CAROLI, C., ROULET, B., J. Crystal

Growth 76 (1986) 31.

[38] HONJO, H., OHTA, S. and SAWADA, Y., Phys. Rev.

Lett. 55 (1985) 841.

[39] FREDERICK, K. J. and HILDEBRAND, H. J., J.A. C. S.

61 (1939) 1555.

[40] KAUKLER, W. F. and RUTTER, J. W., Mat. Sci. Eng.

65 (1984) L1.

[41] ANDERSSON, P. and Ross, R. G., Mol. Phys. 39

(1980) 1359.

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

L’interaction entre EB1 et CLIP-170 (et Bim1p et Bik1p) est intéressante. Nos données nous font penser que cette interaction n’est pas importante aux bouts plus des microtubules.

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

Ces remarques nous ont conduit a` ´etudier les processus ponctuels marqu´es qui permettent d’allier les atouts des approches stochastiques robustesse au bruit, prise en compte du

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

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

La mobilité et la progression des dépôts sableux se font surtout d’ouest vers l’est avec des directions secondaires (ouest nord ouest, est nord est, ect…), en corrélation avec

RJC : Revue de la Jurisprudence Commerciale SEC : Securities and exchange Commission.. امك ت دأ تامزلأا هذه ةمكاترلما لىإ لإا تلا باحسن