Vous êtes sur la page 1sur 9

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elseviers archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright

Author's personal copy

Available online at www.sciencedirect.com

Acta Materialia 60 (2012) 27112718


www.elsevier.com/locate/actamat

Local non-equilibrium diusion model for solute trapping during


rapid solidication
S.L. Sobolev
Institute of Problems of Chemical Physics, Academy of Sciences of Russia, Chernogolovka, Moscow Region 142432, Russia
Received 19 January 2012; accepted 23 January 2012
Available online 1 March 2012

Abstract
A local non-equilibrium diusion model (LNDM) for rapid solidication of binary alloys has been briey reviewed and used to modify a number of solute trapping models with dierent solidliquid interface kinetics. The LNDM takes into account deviation from local
equilibrium of a solute diusion eld in bulk liquid on the basis that the exact solutions to hyperbolic diusion equations govern the
solute concentration and solute ux in bulk liquid under local non-equilibrium conditions. The LNDM leads to a velocity-dependent
eective diusion coecient in bulk liquid ahead of the solidliquid interface DLNDM
V , which goes to zero when the interface velocity
b
V ! VDb, where VDb is the bulk liquid diusion speed. The results show an abrupt transition from diusion-limited to purely thermally
controlled solidication, with the diusion coecient in bulk liquid DLNDM
V = 0 and complete solute trapping KLNDM(V) = 1 at a nite
b
interface velocity V = VDb for any type of solidliquid interface kinetics. The bulk liquid diusion speed VDb is a critical parameter for
the transition. The velocity dependence of partition coecients KLNDM(V) has been calculated for dierent types of solidliquid interface
kinetics, with allowance for local non-equilibrium diusion eects. The calculation shows that the local non-equilibrium partition coefcients KLNDM(V) reduce to the standard K(V) at low interface velocity (V  VDb) and dier substantially at high interface velocity
(V  VDb).
2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Solute trapping; Rapid solidication; Local non-equilibrium diusion; Hyperbolic diusion equation

1. Introduction
Rapid solidication phenomena occur under conditions
that are far from equilibrium [140]. The most obvious
manifestations of the deviations from local equilibrium is
solute trapping, which increases the solute concentration
in the solid phase and reduces the segregations in the liquid
side of interface. The degree of solute trapping is usually
quantied by the partition coecient K, dened as the ratio
of the concentration of solute in the solid to that in the
liquid at the interface. The rapid solidication process
under local non-equilibrium conditions helps to obtain a
very ne structure with uniform properties. Examples of
rapid solidication products accompanied by solute trap Tel.: +7 9032478152.

E-mail address: SSL55@yandex.ru

ping are powders, wires and foils which can be used in powder metallurgy or in producing higher-performance
composite materials. Therefore, development of the capability to predict solute trapping phenomena is an important
task in designing new materials and new processes.
The eect of solute trapping has been investigated both
theoretically [115,2640] and experimentally [1625]. Theoretical studies include analytical models [814,2640],
phase-eld models [5,34,37], molecular dynamics simulations [6,7] and Monte Carlo computer simulations [15].
The most important question in all these studies is how
to achieve complete solute trapping K(V) = 1 at a high
interface velocity V. The analytical models [815] take into
account deviation from equilibrium only at the solidliquid
interface using dierent approaches to the kinetic processes
which take place on the atomic scale at the interface during
crystal growth. The models t experimental data quite well

1359-6454/$36.00 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2012.01.036

Author's personal copy

2712

S.L. Sobolev / Acta Materialia 60 (2012) 27112718

at relatively low interface velocities, and predict complete


solute trapping K(V) = 1 only at V ! 1. The local nonequilibrium diusion model (LNDM) [2629], based on
the assumption of the absence of a local equilibrium during
solute diusion in bulk liquid, leads to complete solute
trapping K(V) = 1 at a nite interface velocity V = VDb,
where VDb is the diusive velocity in the bulk liquid, i.e.
the speed of atomic diusion in the liquid phase. Complete
solute trapping was also obtained in a phase-eld model [5]
and molecular dynamics simulations [6,7]. The partition
coecient K(V)  1 was also observed in the experiments
[2325]. LNDM [2629] takes into account the deviation
from a local equilibrium of the diusion eld in bulk liquid
on the basis of the exact solutions to hyperbolic diusion
equations for solute concentration and solute ux in bulk
liquid. To describe local non-equilibrium solute partitioning, the LNDM was combined with the continuous growth
model (CGM) [9,11,12], which considers deviation from
equilibrium at the solidliquid interface. Two characteristic
velocities the diusive velocity in the bulk liquid VDb
[27,28,42], and the diusive velocity at the solidliquid
interface VDi [12] were used to calculate the solute partition coecient [27,28]. The approach predicted the transition to diusionless solidication with complete solute
trapping at V = VDb and tted the K(V) data quite well
[2628]. LNDM has been successfully used to analyze
non-equilibrium dendrite growth in a bulk undercooled
alloy melt [31,35], morphological stability for a planar
interface incorporating the eect of nonlinear liquidus
and solidus [32], rapid solidication of plasma-sprayed
yttria-stabilized zirconia [33], solute drag eects for binary
alloy solidication with a planar phase interface [34], planar interface migration during solidication to a non-dilute
solution with non-straight liquidus and solidus curves [36],
modeling of overall solidication kinetics for undercooled
single-phase solid-solution alloys [38], oscillatory morphological stability for rapid directional solidication [39] and
grain renement for undercooled single-phase solid-solution alloys [40].
The main purpose of this paper is to combine the classical solute trapping models based on dierent approaches to
solidliquid interface kinetics [915] with LNDM to provide an adequate description of solute trapping for both
low (V  VDb) and high (V  VDb) values of interface
velocity.

@C Db @ 2 C

Db r2 C
@t V 2Db @t2

@J
Db @ 2 J
2
Db rrJ
@t V Db @t2

These hyperbolic equations predict the nite speed of


the diusive wave VDb = (Db/s)1/2, i.e. the maximum speed
with which the diusion perturbations can propagate in the
bulk liquid [41,42,45]. It should be noted that VDb limits
only the speed of diusive perturbations (diusive signal),
but the interface velocity V can be greater than VDb. To
derive the interface condition, we integrate the balance
low over an innitesimal zone that includes the interface
between the liquid and solid phases. The interface condition is given as
V C  C S J  J S

According to extended irreversible thermodynamics [41]


and other local non-equilibrium approaches [4245], the
simplest generalization of the classical Fick law for mass
transport, which includes the relaxation to local equilibrium of the diusion eld, is given as
1

where J is the solute ux, C is the solute concentration, Db


is the diusion coecient in bulk liquid and s is the relax-

where CS and JS are the solute concentration and solute


ux, respectively, in the solid at the interface.
Now let us consider the solute concentration eld ahead
of the interface moving with constant velocity V = const.
Following the usual steady-state approach, we view the
solidication from a reference frame attached to a planar
liquidsolid interface. In such a case, a one-dimensional
version of Eqs. (2) and (3) takes the form
d 2C
dC
Db 1  V 2 =V 2Db 2 V
0
5
dX
dX
dJ
Db 1  V 2 =V 2Db
VJ 0
6
dX
Exact solutions to Eqs. (5) and (6) give the solute concentration and solute ux distributions in the liquid X > 0
(origin of the reference frame is xed on the beginning of
interface X = 0) [2629,42]:
C LNDM X
(
C i  C 0 expVX =Db 1  V 2 =V 2Db  C 0 ; V < V Db

C0;
V > V Db
(
J LNDM X

2. The local non-equilibrium diusion model

J s@J =@t Db rC

ation time of J. In contrast to the Fick law, which leads to a


diusion equation of parabolic type, the evolution Eq. (1)
gives rise to hyperbolic equations for the solute concentration and solute ux [2629,42]:

7
J i expVX =Db 1  V 2 =V 2Db ;

V < V Db

0;

V > V Db

8
where C0 and Ci are the solute concentration in the liquid
far from (X ! 1) and at the interface (X = 0), respectively. The results of the LNDM (Eqs. (7) and (8)) clearly
demonstrate that the value of the diusion speed VDb decisively aects the diusion eld in the bulk liquid. When
V < VDb, there is a solute concentration gradient and solute
ux in the liquid, and the solidication is essentially controlled by the solute ux. It should be noted that, as the

Author's personal copy

S.L. Sobolev / Acta Materialia 60 (2012) 27112718

interface velocity approaches zero, the relaxation model


approaches the classical formulation. However, as the
velocity increases, the solute boundary layer shrinks more
rapidly than expected from the classical mass transport theory, and its thickness dLNDM, dened as (see Eqs. (7) and
(8))
(
Db 1  V 2 =V 2Db =V ; V < V Db
LNDM
d
V =V Db
9
0;
V > V Db
approaches zero at V = VDb. When V > VDb, the LNDM
(Eqs. (7) and (8)) implies that C(X) = C0, J(X) = 0 and
d(V) = 0 at X > 0. This means that the solute concentration
eld ahead of the interface is undisturbed. The result has a
clear physical meaning: a source of perturbations (i.e. the
interface) moving with a velocity greater than the maximum speed of perturbations cannot disturb the medium
ahead of itself [42]. In this case, the solute atoms do not
have enough velocity to escape the solidliquid interface.
Thus, there is no diusion of solute in the bulk liquid at
V > VDb and, consequently, the solidication cannot be
controlled by diusion. Hence, the solidication mechanism changes qualitatively when the interface velocity V
passes through the critical point V = VDb. At this point,
a sharp transition from diusion-controlled to purely thermally controlled growth occurs.
The comparison of exact solutions to the LNDM, Eqs.
(7)(9), with classical solute concentration and solute ux
eld (see e.g. Ref. [9]), allows us to introduce the eective
diusion coecient in bulk liquid DLNDM
[2629] as
b
(
Db 1  V 2 =V 2Db ; V < V Db
DLNDM
V =V Db
10
b
0;
V > V Db
If V  VDb, the eective diusion coecient DLNDM
reduces
b
to the classical diusion coecient Db. However, when V is
of the order of VDb, DLNDM
predicts less solute ux
b
J DLNDM
rC
than
expected
from
the classical Fick law
b
J = Db$C. If V > VDb, then DLNDM
0, which diers
b
qualitatively from the prediction of the classical approach.
It implies the absence of solute diusion ahead of the interface. The eective diusion coecient (Eq. (10)) can be
used to modify some results of the classical rapid solidication theory based on Fick law by substituting DLNDM
for Db
b
at least for steady-state regimes.
3. Solute partitioning
The results of LNDM Eqs. (7)(10) clearly demonstrate
that when V P VDb there is no diusion of solute ahead of
the interface X > 0 (C(X) = C0 and J(X) = 0). This implies
that in such a case the partition coecient KLNDM  1. It
should be stressed that the result of the LNDM,
KLNDM  1 at V P VDb, does not depend on the interface
kinetics it is purely a diusive eect [2629]. This means
that for any interface kinetics the limit KLNDM(V) ! 1 will
be achieved at V ! VDb as long as no other solute
transport process occurs in bulk liquid.

2713

3.1. Continuous growth model, dilute limit


The continuous growth model (CGM) [11,12] treats the
case where the interface is atomically rough enough for
growth and redistribution to occur simultaneously as
strictly steady-state processes, even on the microscopic
scale of the crystal lattice. It predicts a velocity dependence
of the partition coecient given by
K CGM V K E V =V Di =1 V =V Di 

11

where KE is the equilibrium partition coecient and VDi is


the interface diusion speed. At V = VDi, Eq. (11) describes
the mid-transition between KE and unity, and predicts
complete solute trapping KE = 1 only at V ! 1.
The diusive speed in the LNDM, VDb, and the interface
diusion speed, VDi, in Eq. (11) require further discussion.
The CGM denes the VDi as the ratio of solute diusivity
through the interface Di to the atomic jump distance at
the interface ki. In other words, VDi is the average diusive
speed over the interface region, and has been called the
interface diusive speed VDi. The interface diusive speed
is a kinetic rate parameter for solute redistribution at the
interface. The LNDM (Eqs. (1)(10)) denes the diusive
speed VDb as the speed of propagation of a diusive wave
in the bulk liquid [2629,42]. The diusive speed VDb is a
diusive parameter for solute diusion in bulk liquid under
local non-equilibrium conditions and it does not depend on
the interface kinetics. Taking into account that the relaxation time s in the LNDM can be rewritten as s = kb/VDb,
where kb is the atomic jump distance in the bulk liquid,
the diusive speed in the LNDM takes the form
VDb = Db/kb, where Db is the solute diusion coecient in
the bulk liquid. Assuming that Di  Db and ki  k, one is
led to VDi  VDb. This assumption was considered in Refs.
[26,29]. In the general case, these two velocities are dierent
(VDi < VDb), and the partitioning of solute depends on both
VDb and VDi [27,28]. The interface diusive speed VDi governs the solute partitioning at a relatively low interface
velocity, when the interface kinetics is a limiting stage of
solute redistribution. At high interface velocity (V  VDb).
the limiting stage of solute redistribution is the solute diusion in the bulk liquid, with VDb being the characteristic
parameter.
Thus, introduction of DLNDM
V (Eq. (10)) into the
b
expression for K(V) (Eq. (11)) leads to the generalized
partition coecient KLNDM(V), which takes into account
both interfacial kinetic eects according to the CGM (the
interface redistribution parameter VDi) and local nonequilibrium diusion eects according to the LNDM (the
bulk liquid diusion speed VDb) [27,28]:
( K 1V 2 =V 2 V =V
E
Di
Db
; V < V Db
LNDM
1V 2 =V 2Db V =V Di
K
V
12
1;
V > V Db
Evolutions of solute partition coecients as functions of
the non-dimensional interface velocity, V/VDb (Eqs. (11)
and (12)), are presented in Fig. 1, with KE = 0.1 and

Author's personal copy

2714

S.L. Sobolev / Acta Materialia 60 (2012) 27112718

a monolayer, including any solute atoms in the layer. Solute diuses back into the liquid during the period before
the passage of the next step, at which point any remaining
solute is assumed to be permanently trapped in the solid.
The predicted velocity dependence of K for this mechanism
is [9,13]

1,0
LNDM-Jackson
LNDM-CGM

Partition coeficietnt

0,8
LNDM-SGM
0,6
SGM
0,4

K SGM V K E 1  K E expV Di =V

CGM

Taking into account local non-equilibrium diusion


eects in bulk liquid according to the LNDM approach,
Eq. (15) can be rewritten as

Jackson
0,2

0,0
0,0

0,2

0,4

0,6

0,8

15

1,0

1,2

1,4

Nondimensional velocity

Fig. 1. Partition coecient K as a function of interface velocity scaled to


the bulk diusion speed VDb with KE = 0.1. Upper solid curve: LNDM,
Eq. (12); lower solid curve: CGM, Eq. (11), with KE = 0.1 and VDb/
VDi = 2; upper dash-dotted curve: LNDM, Eq. (16); lower dash-dotted
curve: SGM, Eq. (15), with KE = 0.1 and VDb/VDi = 2; upper dashed
curve: LNDM, Eq. (20); lower dashed curve: Jackson et al. model, Eq.
(19), with A0 = 0.8 s m1 [15] and VDb = 15 m s1.

VDb/VDi = 2. The LNDM, Eq. (12) the upper solid curve


in Fig. 1 demonstrates that the transition to complete solute trapping KLNDM(V) = 1 occurs at a nite interface
velocity V = VDb, while the CGM, Eq. (11) the lower
solid curve in Fig. 1 predicts the complete solute trapping
only at V ! 1.
3.2. Continuous growth model, non-dilute limit

K LNDM V
(
K E 1  K E exp1  V 2 =V 2Db V Di =V ; V < V Db

1;
V > V Db
16
The expression for KLNDM(V), Eq. (16), takes into
account both the departure from the local equilibrium of
the diusion process in bulk liquid according to the
LNDM and the interface kinetic conditions assumed by
the SGM. A comparison between the SGM and the
LNDM is also presented in Fig. 1, with KE = 0.1 and
VDb/VDi = 2: the lower dash-dotted curve being the
SGM, Eq. (15), and upper dash-doted curve being the
LNDM, Eq. (16). The LNDM predicts complete solute
trapping KLNDM(V) = 1 at V = VDb, whereas SGM at predicts it at V ! 1.
3.4. Aperiodic stepwise growth model (ASGM)

According to the CGM, the partition coecient in nondilute limit takes the form [12]
K CGM V ; C 0 K E 1  K E C 0 V =V Di =1 V =V Di 
13
This equation can also be modied to the local
non-equilibrium diusion case by using the concept of
eective diusion coecient DLNDM
V , Eq. (10). After
b
some algebra, we have
K LNDM V ; C 0
8
2
2
2
2
>
< 1  V =V Db K E 1  K E C 0 V =V Di =1  V =V Db
V =V Di ;
V < V Db
>
:
1;
V > V Db

The ASGM [13,23] treats the same case as the SGM,


except that the passage of steps is assumed to occur randomly, rather than periodically, in time. It predicts variation of K with interface orientation. The ASGM assumes
that at any orientation the interface is broken into (1 1 1)
terraces of single double-layer height and random width,
and that solidication proceeds by the lateral passage of
these steps at random intervals. Some (though not all, as
in the SGM) of the solute atoms in the monolayer of liquid
adjacent to the interface are trapped as a step passes. The
ASGM expression for K(V, h) is given by [13,23]
K ASGM V ; h

14

As expected, the partition coecient KLNDM(V, C0) ! 1


when V ! VDb due to local non-equilibrium diusion
eects. The same expression, Eq. (14), was also obtained
in Ref. [30] using the LNDM calculation of a liquid diusion eld [2629] and the CGM approach to solidliquid
interface kinetics [12].
3.3. Stepwise growth model (SGM)
The SGM treats the case in which an atomically smooth,
sharp interface advances by the periodic lateral passage of

K E bT K E bL =1 bL 
1 bT

17

where bT V =V TD cos h; bL V =V LD cos h; V LD is the


diusive speed at the ledge, V TD is the diusive speed at
the terrace and h is the angle of inclination from (1 1 1).
Using the eective diusion coecient (10), the LNDM
leads to
K LNDM V ; h
( 1V 2 =V 2

DB

1;

K E bT 1V 2 =V 2Db K E bL =1V 2 =V 2Db bL


;
1V 2 =V 2Db bT

V < V Db
V > V Db
18

Author's personal copy

S.L. Sobolev / Acta Materialia 60 (2012) 27112718

Eq. (18) generalizes Eq. (17) for the local non-equilibrium


case and predicts complete solute trapping at V ! VDb, in
contrast to Eq. (17), which gives K(V, h) ! 1 only at V ! 1.
3.5. The Jackson et al. model
Another theory for solute trapping due to Jackson et al.
[14,15] is derived using reaction rate theory to describe the
rate of atom attachments to the active sites of a sharp
solidliquid interface. The model is based on the standard
quasi-equilibrium model for crystallization, which has been
extended to include the non-equilibrium eects found in the
Monte Carlo simulations of alloy solidication [15]. The
parameter in the model, b, is clearly identied as the important parameter for solute trapping. This parameter embodies the relationship between the growth and diusion
parameters found in the Monte Carlo simulations. This
parameter is interpreted as the ratio of the distance the
interface advances to the distance which an atom can move
by diusion during the time that it is at the interface. If s is
the average time it takes for an atom to join the crystal, the
distance which the interface advances during this time is
Vs. The distance which an atom can move by diusion during this time is (Dbs)1/2, where Db is the diusion coecient
of B atoms in the liquid. The ratio of these two distances is
the dimensionless parameter b = Vs/(Dbs)1/2 [14,15].
Taking into account that VDb = (Db/s)1/2 and the fact that
the relaxation time in the LNDM corresponds to the
average time it takes for an atom to join the crystal in
the Jackson et al. model, the parameter b transforms to
the ratio V/VDb. The parameter V/VDb is also the critical
parameter in the LNDM, Eqs. (1)(10), and in some other
models [42], but in the LNDM V/VDb governs solute diusion in bulk liquid and, hence, solute partitioning at
V ! VDb, and in the Jackson et al. model parameter b
governs non-equilibrium segregation at the solidliquid
interface. Parameter (Dbs)1/2 is the distance which an atom
can move by diusion during time s in bulk liquid far from
the interface when the interface does not inuence the diffusion process in the liquid. However, in some cases the
assumption Di = Db is a reasonable approximation. Thus,
for rough interfaces and a low concentration of solute,
the theory of Jackson et al. yields [14,15]
1=1A0 V

K J V  K E

19

where A0 depends on the square root of the diusion coefcient. From Eqs. (10) and (19), one can calculate
8 pp

<
Db 1V 2 =V 2Db
Db 1V 2 =V 2Db A0 V
; V < V Db
K LNDM V K E
: 1;
V > V Db
20

The partition coecient, Eq. (20), incorporates both the


interface kinetic model of Jackson et al. and the local
non-equilibrium diusion eects according to the LNDM.

2715

The results of Eqs. (19) and (20) are shown in Fig. 1, using
KE = 0.1, A0 = 0.8 s m1 obtained from the Monte Carlo
simulation results [15] and VDb = 15 m s1 [1,17,19]: the
lower dashed curve shows the Jackson et al. model; the
upper dashed curve, the LNDM. As expected, one can
observe the partitionless solidication KLNDM(V) ! 1 predicted by the LNDM at V ! VDb, in contrast to Jackson
et al. model, which leads to complete solute trapping only
at V ! 1.
3.6. The Burton et al. model
The Burton et al. model [10] examines the incorporation
of solute elements into single crystals of germanium grown
from the melt. It takes into account the contribution of solute transport in the melt, owing to diusion and uid
motion, to the overall process of impurity incorporation
during steady-state crystallization. The crystallization is
viewed as a heterogeneous reaction. The chemical process
consists of the incorporation of impurity at the solidliquid
interface. The analysis is extended to treat the transient
inclusion of solute that results when the composition of
the melt is changed abruptly. For the important case in
which equilibrium prevails at the interface virtually independently of growth rate, this leads to the following expression for partition coecient [10]:
K BPS V K E =K E 1  K E expV d=Db

21

where d is a parameter that depends on certain physical


properties of the liquid phase and the crystallization
conditions. Using the expression for the eective diusion
coecient, Eq. (10), after some algebra we have
K LNDM V
(
K E =K E 1  K E expV d=Db 1  V 2 =V 2Db ; V < V Db

1;
V > V Db
22

The modied partition coecient, Eq. (22), also predicts


complete solute trapping KLNDM(V) ! 1 when V ! VDb,
whereas Eq. (21) leads to complete solute trapping only
at V ! 1.
3.7. The Elliott and Peppin model
In their recent work, Elliott and Peppin [8] used the
Boltzmann velocity distribution of a particle to derive an
expression for the non-equilibrium segregation coecient
K(V) at an icecolloidal suspension interface as a function
of the freezing interface velocity V and particle radius R.
The model incorporates two characteristic velocities: the
critical engulfment velocity Vc and the molecular [8]
velocity of the particles VK = (2kBT/mp)1/2, where mp is
the mass of a particle, kB is Boltzmanns constant and T
is temperature. In this case, the partition coecient takes
the form [8]

Author's personal copy

2716

S.L. Sobolev / Acta Materialia 60 (2012) 27112718


2 !
VK
VC V
p exp 
K V
VK
2V p


1
VC V
erfc
2
VK
EP

23

Using the eective diusion coecient, Eq. (10), and


taking into account that VK is proportional to D1/2 and
plays the same role in the model of Elliott and Peppin as
VDb in the LNDM, the partition coecient, Eq. (23), can
be rewritten as
K LNDM V

V K 1  V 2 =V 2K
p
2V p
0
 exp @

1=2

VC V

V K 1  V 2 =V 2K 1=2
!
VC V

1
erfc
1=2
2
V K 1  V 2 =V 2K

!2 1
A

24

When V ! VK the modied partition coecient, Eq.


(24), predicts complete solute trapping KLNDM(V) ! 1
due to local non-equilibrium eects described by the eective diusion coecient (Eq. (10)). We expect that the local
non-equilibrium eects at high interface velocity can be
also incorporated into the Elliott and Peppin model on
the basis of kinetic theory using, for example, the relaxation-time approximation [41], but the problem requires special consideration.
4. Hierarchy of deviation from equilibrium
The solidication of undercooled melts occurs at a wide
range of velocities, up to hundreds of meters per second,
and includes various physical phenomena, such as solute
diusion, heat conduction and the interface kinetics of
solidication. Each of these processes has its own characteristic time and length scales (in a steady state characteristic velocities). This implies that, while the interface
velocity increases the solute diusion, the interface kinetics
and heat conduction deviate from equilibrium at dierent
values of interface velocity. This allows us to introduce a
hierarchy of deviations from equilibrium which is followed
with increasing solidication velocity.
1. V = 0; full equilibrium. Here, there is no chemical potential gradient (the composition of the phases are uniform)
and no temperature gradients, K = KE.
2. V  VDb; local equilibrium. Here, there are concentration and temperature gradients near the interface, i.e.
there is no full equilibrium, but there is local equilibrium
both in the bulk liquid and at the interface. The partition coecient is equal to its equilibrium value KE. Diffusion and temperature elds are described by a local
equilibrium transfer equation of the parabolic type.
3. V < VDb; non-equilibrium interface kinetics. Here, there
is no local equilibrium at the interface and the partition

coecient depends on the interface velocity V (Eqs. 11,


13, 15, 17, 19, and 21). In this case (as well as in case 2),
the solute concentration and temperature elds are governed by the classical (local equilibrium) parabolic-type
transport equation.
4. V 6 VDb; non-equilibrium diusion eld. In this case,
there is no equilibrium at the interface but local equilibrium in the bulk liquid. Hence, according to the LNDM,
the solute concentration and solute ux elds in the bulk
liquid are governed by the hyperbolic-type mass transport equations, Eqs. (2) and (3), and take the form of
Eqs. (5) and (6) for the steady-state regime [2629,42].
The temperature eld is still at a local equilibrium due
to V  VDb  VT and it can still be described by the
classical parabolic-type heat conduction equation. Partition coecients KLNDM(V) are given by Eqs. 12, 14, 16,
18, 20, and 22 for dierent types of interface kinetic
model. The choice of the model depends on the characteristics of a particular rapid solidication process at the
interface and the conditions of the experiments. When
V 6 VDb, solute partitioning is governed by both the
interface kinetics and the solute diusion in the bulk
liquid. However, when the interface velocity increases,
the dependence of KLNDM(V) on the interface kinetics
is weaker and solidication depends mainly on VDb,
demonstrating a transition to the diusionless regime
DLNDM
V ! 0, with complete solute trapping
b
KLNDM(V) ! 1 at V ! VDb for all interface kinetic
models.
5. V P VDb; diusionless solidication. According to the
LNDM, at this interface velocity KLNDM(V)  1 and
DLNDM
V  0. This implies thermally controlled diub
sionless solidication with complete solute trapping.
Ahead of the interface, C(X) = C0 and J(X) = 0. The
result does not depend on the interface kinetics but is
a consequence of deviation from local equilibrium in
the bulk liquid, with VDb being the critical parameter
[2629].
6. V  VT; local non-equilibrium temperature eld. At such
high velocities there is also local equilibrium for heat
transport processes. The temperature eld is governed
by hyperbolic transport equations [2629]. If the solid
liquid interface only propagates due to undercooling
eects, the interface velocity V is limited by VT the
speed of the heat wave [42].

5. Discussion
The LNDM clearly demonstrates that the solidication
mechanism changes qualitatively when the interface velocity V passes through the critical point V = VDb. At this
point a sharp transition from mostly diusion-controlled
to purely thermally controlled regimes occurs. When
V < VDb, there is a solute concentration gradient and solute
ux ahead of the interface and the solidication is governed
by both the redistribution of heat and the solute, whereas

Author's personal copy

S.L. Sobolev / Acta Materialia 60 (2012) 27112718

1,0

Partition coeficietnt

0,9
0,8
0,7
0,6
0,5
0,4
0,3
0,5

1,0

1,5

2,0

2,5

Interface velocity, m/s

Fig. 2. Partition coecient K as a function of interface velocity for SiAs


alloys. Data points are from Refs. [23,25]. The dashed curve represents the
CGM, Eq. (11), with KE = 0.3 and VDi = 0.46 m s1; the solid curve
represents the LNDM, Eq. (12), with the same KE, and VDb = 2.45 m s1
and VDi = 0.65 m s1.

1,0

Partition coefficient

0,9

0,8

0,7

0,6

0,5
1

Interface velocity, m/s

Fig. 3. Partition coecient K as a function of interface velocity for GeSi


alloys. Data points are from Ref. [24]. The dashed curve represents the
CGM, Eq. (11), with KE = 0.4 and VDi = 2.03 m s1; the solid curve
represents the LNDM, Eq. (12), with the same KE, and VDi and
VDb = 4.47 m s1.

at V > VDb the solute concentration gradient and solute


ux are both zero in the bulk liquid and solidication is
purely thermally controlled. It should be stressed that the
transition occurs purely as a diusion eect at V = VDb
for any type of interface kinetics (see Fig. 1).
Such a sharp transition from diusion-controlled to
purely thermally controlled growth has been observed in
CuNi alloys [1618], NiB alloys [19], AgCu alloys
[20,21] and TiNi alloys [22]. Moreover, the investigations
[16] show that it is not a critical undercooling that initiates
the transition but, rather, a critical solidication velocity,
which approximately equals the diusive speed VDb. Walder and Ryder [20,21] have shown that the sharp transition
from diusion-controlled to purely thermally controlled
growth for AgCu alloys corresponds to T0 temperature

2717

and occurs at a nite interface velocity. The same results


have been obtained by Walder for TiNi alloys [22]. Taking
into account the LNDM predictions that the temperature
T0 corresponds V = VDb [28,29], the experimental results
on TiNi alloys agree with the experimental results for
CuNi alloys [1618] and NiB alloys [19] that the transition to diusionless solidication occurs at V = VDb. Thus,
the experiments [1622] give strong support to the LNDM
results that the local non-equilibrium diusion in bulk
liquid plays an important role in rapid solidication and
governs the transition to diusionless and partitionless
solidication, which occurs at a nite interface velocity
V = VDb.
The complete solute trapping K(V) = 1 is predicted by all
the classical models [815] only at V ! 1, in contrast to the
LNDM, which gives KLNDM(V) = 1 at the nite interface
velocity V = VDb for any type of interface kinetics (see
Fig. 1). A molecular dynamics study by Cook and Clancy
[6] for a LennardJones system has also shown complete
solute trapping for unstrained growth on (1 0 0) when the
interface velocity attained its steady-state regrowth value
of 4 m s1. The most recent molecular dynamics simulation
[7] provides a complete characterization of solidliquid
kinetic properties, including the nature of crystalline anisotropies and the details of the velocity or driving-force relations. The approach is applied to two model systems with
atomically rough solidliquid interfaces, which both crystallize into face-centered cubic crystal structures, but with
diering degrees of equilibrium solute partitioning. The
results [7] also show an abrupt transition to complete solute
trapping K(V) = 1 at a nite growth velocity.
The velocity dependence of the partition coecient was
measured for rapid solidication of polycrystalline SiAs
alloys induced by pulsed laser melting [23,25]. The experimental results are compared with predictions of the
CGM, Eq. (11), and LNDM, Eq. (12), in Fig. 2. The
CGM ts the data well only at a relatively low interface
velocity, while the LNDM ts the experimental results better at all values of interface velocity. The K vs. V experimental data obtained by pulsed laser melting of GeSi
alloys [24] are shown in Fig. 3, together with the predictions of both the CGM and LNDM. The CGM accurately
ts the experimental data at a relatively low interface velocity, whereas the LNDM describes the data very well at both
low and high interface velocities.
6. Conclusion
The local non-equilibrium diusion model based on
hyperbolic diusion equations predicts an abrupt transition to diusionless solidication with complete solute
trapping KLNDM(V) = 1 at a nite interface velocity
V = VDb.
The concept of the eective diusion coecient DLNDM
,
b
Eq. (10), allows one to modify the results for the solute
partition coecient K(V) obtained for dierent types of
interface kinetics in local equilibrium diusion cases.

Author's personal copy

2718

S.L. Sobolev / Acta Materialia 60 (2012) 27112718

The transition to complete solute trapping is controlled


solely by solute diusion in the bulk liquid and occurs
for any type of interface kinetics at V ! VDb.
The predictions of the LNDM are in agreement with
experimental data, molecular dynamics simulations
and phase-eld modeling.

[19]
[20]
[21]
[22]
[23]
[24]
[25]

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]

Herlach DM. Mater Sci Eng 1994;R12:177.


Trivedi R, Kurz W. Int Mater Rev 1994;39:49.
Liu F, Yang GC. Int Mater Rev 2006;51:145.
Asta M, Beckermann C, Karma A, Kurz W, Napolitano R, Plapp M,
et al. Acta Mater 2009;57:941.
Furtado HS, Bernardes AT, Machado RF, Silva CA. Mater Res
2009;12:345.
Cook SJ, Clancy P. J Chem Phys 1993;99:2175.
Yang Y, Humadi H, Buta D, Laird BB, Sun D, Hoyt JJ, et al. Phys
Rev Lett 2011;107:025505.
Elliott JAW, Peppin SSL. Phys Rev Lett 2011:168301.
Hall RN. Phys Rev 1953;88:139;
Hall RN. J Phys Chem 1953;57:836.
Burton JA, Prim RC, Slichter WP. J Chem Phys 1953;21:1987.
Aziz MJ. J Appl Phys 1982;53:1158.
Aziz MJ, Kaplan T. Acta Metall 1988;36:2335.
Reitano R, Smith PM, Aziz MJ. J Appl Phys 1994;76:1518.
Jackson KA, Kirk M, Beatty KM, Gudgel KA. J Cryst Growth
2004;271:481.
Beatty KM, Jackson KA. J Cryst Growth 2004;271:495.
Willnecker R, Herlach DM, Feuerbacher B. Phys Rev Lett
1989;62:2707.
Willnecker R, Herlach DM, Feuerbacher B. Appl Phys Lett
1990;56:324.
Eckler K, Cochrane RF, Herlach DM, Feuerbacher B, Jurisch M.
Phys Rev B 1992;45:5019.

[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]

Herlach DM. Mater Sci Eng 1994;A179/A180:147.


Walder S, Ryder PL. J Appl Phys 1993;74:6100.
Walder S, Ryder PL. Acta Metall Mater 1995;43:4007.
Walder S. Mater Sci Eng 1997;229:156.
Kittl JA, Aziz MJ, Brunco DP, Thompson MO. J Cryst Growth
1995;148:172.
Brunco DP, Thompson MO, Hoglung DE, Aziz MJ, Gossmann HJ. J
Appl Phys 1995;78:1575.
Kittl JA, Sanders PG, Aziz MJ, Brunco DP, Thompson MO. Acta
Mater 2000;48:4797.
Sobolev SL. Phys Lett A 1995;199:383.
Sobolev SL. Phys Status Solidi (a) 1996;156:293.
Sobolev SL. Phys Rev E 1997;55:6845.
Sobolev SL. J Int Non-Equilib Proc 1997;10:49.
Galenko P. Phys Rev E 2007;76:031606.
Wang H, Liu F, Chen Z, Yang G, Zhou Y. Acta Mater 2007;55:497.
Wang H, Liu F, Yang W, Chen Z, Yang G, Zhou Y. Acta Mater
2008;56:2592.
Liu H, Salmi Jazi HR, Bussmann M, Mostaghimi J. Acta Mater
2009;57:6013.
Li S, Zhang J, Wu P. Scripta Mater 2010;62:716.
Chen Z, Wang H, Liu F, Yang W. Trans Nonferr Met Soc China
2010;20:490.
Li S, Zhang J, Wu P. J Cryst Growth 2010;312:982.
Wang H, Liu F, Chen Z, Yang W. Trans Nonferr Met Soc China
2010;20:877.
Wang H, Liu F, Yang G, Zhou Y. Acta Mater 2010;58:5402.
Wang H, Liu F, Wang K, Zhai H. Acta Mater 2011;59:5859.
Wang H, Liu F, Tan Y. Acta Mater 2011;59:4787.
Jou D, Casas-Vazquez J, Lebon G. Extended irreversible thermodynamics. Berlin: Springer; 1996.
Sobolev SL. Sov Phys Usp 1991;34:217.
Nettleton RE, Sobolev SL. J Non-Equilib Thermodynam
1995;20:205.
Nettleton RE, Sobolev SL. J Non-Equilib Thermodynam
1995;20:297.
Nettleton RE, Sobolev SL. J Non-Equilib Thermodynam 1995;21:1.

Vous aimerez peut-être aussi