Está en la página 1de 9

Journal of Crystal Growth 395 (2014) 4654

Contents lists available at ScienceDirect

Journal of Crystal Growth


journal homepage: www.elsevier.com/locate/jcrysgro

An analytical model for solute diffusion in multicomponent


alloy solidication
S.L. Sobolev a,n, L.V. Poluyanov a, Feng Liu b
a
b

Institute of Problems of Chemical Physics, Academy of Sciences of Russia, Chernogolovka, Moscow Region 142432, Russia
State Key Laboratory of Solidication Processing, Northwestern Polytechnical University, Xi'an, Shaanxi 710072, People's Republic of China

art ic l e i nf o

a b s t r a c t

Article history:
Received 1 January 2014
Received in revised form
12 February 2014
Accepted 8 March 2014
Communicated by: M. Uwaha
Available online 18 March 2014

An analytical model describing steady-state plane front solidication of multicomponent alloys has been
developed, taking into account both the diffusive interaction between the solutes and the local
nonequilibrium diffusion effects in the bulk liquid. These effects strongly affect the solute distributions
in the liquid phase ahead of the interface. At relatively low interface velocity the strong diffusional
interaction between the species leads to a minimum or a maximum of solute concentration ahead of the
interface. In ternary systems the model predicts two-step transition to diffusionless solidication.
& 2014 Elsevier B.V. All rights reserved.

Keywords:
A1. Solidication
A1. Difusion
A1. Segregation
B1. Alloys

1. Introduction
Rapid solidication of multicomponent alloys (three or more components) is pertinent to many commercial materials and industrial
processes, while also raising challenging questions from a fundamental point of view [17]. Most of the models describing solidication in
multicomponent systems have been developed assuming independent diffusion of the solutes, which implies that the diffusion elds in
multicomponent alloys are then given by the same mathematical functions as in binary systems. Off-diagonal diffusion terms of the
diffusion matrix are usually neglected for multicomponent liquids; yet there is little justication, because in some situations the
diffusional interaction between the species in multicomponent alloys plays an important role but the effect of the off-diagonal diffusion
terms on diffusional transformation kinetics in multicomponent systems is still an open question [1]. Moreover, during rapid solidication
of alloys the solute diffusion occurs under far from local equilibrium conditions, when the solute concentration in the bulk liquid is
governed by the hyperbolic diffusion equation [813]. The local nonequilibrium effects change drastically the diffusion eld in the bulk
liquid leading to a sharp transition from diffusion controlled growth to kinetic controlled growth with complete solute trapping [813].
The transition is often accompanied by grain renement and disorder trapping phenomena.
The aim of this work is to present an analytical model for solute diffusion during rapid solidication of multicomponent alloys, taking into
account both the diffusional interaction between the species and the local nonequilibrium diffusion effects in the bulk liquid. This would lead to
the understanding of the fundamental relationships between different rapid solidication phenomena in multicomponent systems, including the
break in the velocityundercooling function, complete solute trapping, grain renement, and disorder trapping at large undercooling.

2. The model (n-component system)


In most cases, multicomponent diffusion is described by the Onsager law for an n-component system:
n1

J i  Dij C j
j1

Corresponding author.
E-mail address: sobolev@icp.ac.ru (S.L. Sobolev).

http://dx.doi.org/10.1016/j.jcrysgro.2014.03.009
0022-0248/& 2014 Elsevier B.V. All rights reserved.

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

47

where Ji is the ux of the component i, Ci is the concentration of the component i, Dij are diffusion coefcients, which constitute an
n  1n 1 diffusion matrix. In general, the diffusion coefcients are not symmetric (Dij a Dji ). The diagonal terms Dii are the so-called
main terms of the diffusion matrix, because they are commonly larger than the off-diagonal terms Dij and similar in magnitude to the
binary values. The off-diagonal terms are usually 10% or less of the on-diagonal terms.
If the solidliquid interface during alloy solidication moves sufciently fast, the solute diffusion in the bulk liquid occurs under local
nonequilibrium diffusion conditions [813] and, according to the extended irreversible thermodynamics [14], the generalization of the
classical Fick law in the absence of the diffusional interaction is given by
J i i

J i
 Dii C i
t

where i is the relaxation time of Ji to local equilibrium. The modied Fick law, Eq. (2), takes into account a ballistic (or propagative) mode
of mass transfer process at short time scale, when the process occurs under far-from-equilibrium conditions [13,14]. It implies that the
local nonequilibrium diffusion has inertia properties the diffusion ux reacts on the concentration gradient not at the same instant of
time t as in local equilibrium, but later, at the time t i, i.e. J i t i  Dii C i t. The diffusion relaxation time i is the average time
between two successive jumps of particles of solute i. After expansion of J i t i in a Taylor series with allowance for the rst two terms
one can obtain Eq. (2).
In the case of multicomponent diffusion the modied Fick law (2) can be expressed as
J i i

n1
J i
 Dij C j
t
j1

Combining Eq. (3) with the mass conservation law C i =t  J i , we obtain the local nonequilibrium solute diffusion equations for the
multicomponent system in the form

2 C i C i n  1
Dij 2 C j

t
t 2
j1

For the sake of simplicity let us consider a ternary system with two solutes, i.e. i1,2. The corresponding diffusion equations take the
form

2 C 1 C 1
C 2
C 2
D11 12 D12 22

2
t
t
X
X

2 C 2 C 2
C 2
C 2
D22 22 D21 12

t
t 2
X
X

For a steady-state planer growth in a reference frame moving with constant interface velocity V in the X-direction Eqs. (4) and (5)
result in
D11 1  V 2 =V 2D1

D22 1  V 2 =V 2D2

C 21
X

C 22
X

D12

D21

C 22
X 2
C 21
X 2

dC 1
0
dX

dC 2
0
dX

where V Di Dii =i 1=2 is the characteristic diffusion velocity for solute i. The following boundary conditions will be used (the solidliquid
interface is placed at X0):
C i X-1 C 1
i
C i X 0 C 0i

C1
i

C 0i

are the solutes concentrations in the bulk liquid far from the interface (initial concentrations) and
are the solutes
where
concentrations in the bulk liquid at the interface. Introducing velocity dependent on-diagonal terms of the diffusional matrix in the form
Dnu Du 1  V 2 =V 2Di

Eqs. (6) and (7) can be rewritten as


Dn11

Dn22

C 21
X

C 22
X

D12

D21

C 22
X 2
C 21
X 2

dC 1
0
dX

10

dC 2
0
dX

11

These equations coincide in the form with the diffusion equation for the local equilibrium case [6] but with principal difference: on-diagonal
terms of the local equilibrium diffusion matrix Dii are always positive and do not depend on the interface velocity but the on-diagonal terms of the
local nonequilibrium diffusion matrix Dnii are velocity-dependent and can be both positive and negative. As demonstrated below the fact leads to
some qualitatively new properties of the solutions to the multi-component concentration eld under local nonequilibrium conditions.

48

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

3. Results
The diffusion elds for the solutes are given by linear combinations of the binary solutions using the eigenvalues and eigenvectors of
the diffusion matrix instead of the diffusion coefcients. The eigenvalues for Eqs. (10) and (11) are given as

1;2 Dn11 Dn22 =2 7 Dn11 Dn22 2 =4 D12 D21 1=2

12

Note that the eigenvalues of the diffusive matrix for the local nonequilibrium case 1,2, Eq. (12), are velocity-dependent due to the velocitydependent on-diagonal terms of the diffusion matrix Dnii V.
When D12 D21 D, the solutions to Eqs. (10) and (11) take the form (see Appendix A)

2 a2 sin
exp V X=1
exp  VX=2
V
V
1 a1 sin
2 a2 cos
exp  VX=1 
exp  VX=2
C 2 X C 1
2
V
V

C 1 X C 1
1

1 a1 cos

13

where constants ai are determined by the boundary conditions (8), and tan 2 2D=Dn11  Dn22 .
3.1. No diffusional interaction D12 D21 0
Let us rst consider the case when D12 D21 0, i.e. the diffusional interaction between the species is ignored. Without diffusional
interaction the diffusion of each component is unaffected by the presence of the other and the solutes diffuse independently. In this case
the solute concentrations are governed by the independent hyperbolic diffusion equations

2 C i C i
2 C
Dii 2i

2
t
t
X

14

It should be noted that from a mathematical point of view the transfer equation of hyperbolic type predicts that high-frequency
perturbations propagate in the system with a nite velocity [14]. Regarding solute diffusion during rapid solidication it implies that the
diffusional perturbations in the bulk liquid propagate with a nite velocity V Di Dii =i 1=2 , which is usually called diffusive velocity
[813], i.e. the maximum speed with which concentration perturbations can propagate in the bulk liquid. In other words VDi describes
how fast a solute atom can travel in the liquid. Thus the hyperbolic diffusion equation incorporates both the diffusive (dissipative) and
wave (ballistic) features of mass transfer under local nonequilibrium diffusion conditions.
At D12 D21 0 the eigenvalues of the diffusive matrix i reduce to

i Dnii
Fig. 1 shows the eigenvalues of the diffusive matrix
abscissa axis i(V)0 at the points VD1 and VD2.

i as functions of interface velocity V (solid lines). The eigenvalues i (V) cross the

1. If V oVD2 oVD1, then both i(V)4 0


In such a case the both components diffuse ahead of the interface and their concentrations decrease exponentially with increasing X:
0
1
n
C i X C 1
i C i  C i exp  VX=Dii

15

These equations dene the solute boundary layers

i as

i Dii 1  V 2 =V 2Di =V

16

Nondimensional lambda

-1

Nondimensional interface velocity


Fig. 1. Nondimensional eigenvalues of the diffusive matrix i/D22 as functions of nondimensional interface velocity V/VD2 at VD1/VD2 D11/D22 2. Solid lines no diffusion
interaction between the species D12 D21 0, dashed lines with diffusion interaction D12 D21 4 0. Two upper lines 1/D22, two lower lines 2/D22.

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

49

Note that the thicknesses of the solute boundary layers i differ one from another due to the different values of the on-diagonal terms
Dii and the different values the diffusive velocities VDi.
2. If VD2 oV oVD1, then 2(V)o0 and 1(V)40
In this case the solute concentration eld ahead of the interface is described as
0
1
2
2
C 1 X C 1
1 C 1  C 1 exp  VX=D11 1 V =V D1

C 2 X

17

C1
2

18

At such interface velocity only solute 1 diffuses ahead of the interface but solute 2 does not. The result has a clear physical meaning and arises
due to the wave nature of the local nonequilibrium diffusion when the solute concentrations are governed by the hyperbolic diffusion
equations a source of perturbations (i.e. the solidliquid interface) moving with a velocity greater than the maximum speed of perturbations
cannot disturb the medium ahead of itself. This effect is similar to the supersonic propagation of the source of sound. In this case the
solidication mechanism does not depend on diffusion of solute 2 but depends on diffusion of solute 1. In other words, the growth can be
treated as diffusionless regarding to solute 2 but diffusion controlled by solute 1.
3. If V 4VD1 4VD2 then i(V) o0
At such high velocity the solute concentrations eld ahead of the interface cannot be disturbed due to the wave nature of solute
diffusion under local nonequilibrium conditions and the growth occurs in fully diffusionless regime:
C 1 X C 1
1
C 2 X C 1
2
Thus, in ternary alloys when diffusional interaction between the species is ignored the local nonequilibrium diffusion effects lead to the
two-step transition to the diffusionless solidication at the critical points V VD2 and V VD1.
4. The behavior of the local nonequilibrium diffusion eld in case of the independent solute diffusion allows us to introduce the effective
(velocity-dependent) diffusion coefcients as [8,9]
(
Dii 1  V 2 =V 2Di ; V oV Di
ef f
Dii V
19
0;
V 4 V Di
f
Fig. 2 shows the nondimensional effective diffusion coefcients Def
ii V=Dii (solid lines) as functions of nondimensional interface
velocity V/VD2 at VD1/VD2 2. The effective diffusion coefcients allow us to represent the solute diffusion eld, discussed in points 1 to
3 in the form identical to the classical (local equilibrium) solute diffusion eld for 0 oVo 1
ef f
0
1
C i X C 1
i C i  C i exp  VX=Dii

20

Thus the effective diffusion coefcients can be used to extend some results of the local equilibrium theory to the local nonequilibrium case not only for binary alloys [811] but also for the multicomponent alloys if diffusional interaction can be
ignored.

Partition coefficient, K
Nondimensional effective diffusion coefficient

1.2

1.0
solute 2

0.8

solute1

0.6

0.4

0.2

diffusion controlled
by solute 1
(no diffusion of solute 2)

diffusion controlled
solidification

diffusionless
solidification

0.0

-0.2

VD2
0

VD1
2

Nondimensional interface velocity, V


f
Def
ii V =Dii

Fig. 2. Nondimensional effective diffusion coefcients


(solid lines) and partition coefcients Ki(V) (dashed lines) as functions of nondimensional interface velocity
V/VD2 at VD1/VD2 2 for independent diffusion between the species (D 0).

50

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

3.2. Diffusional interaction (off-diagonal terms D a0)


When D a 0 the eigenvalues of the diffusive matrix i(V), Eq. (12), cross the abscissa axis i(V) 0 at the points V2 oVD2 and V1 4VD1
(see the dashed lines in Fig. 1). Thus the diffusion process in the ternary system with diffusional interaction between the species
determines two characteristic velocities Vi, which differ from VDi.
1. When VoV2, the eigenvalues of the diffusive matrix
(see Appendices A and B)

i 40 and the solute concentrations ahead of the interface are governed by

n
n
n
n 2
2 1=2
DC 02  C 1
C 01  C 1
2 D11  D22 D11  D22 4D
1 =2

exp  VX=1
)
n
n
n
n 2
2 1=2
DC 02  C 1
C 01  C 1
0
1
2 D11  D22 D11  D22 4D
1 =2
C1  C1 
exp  VX=2
Dn11  Dn22 2 4D2 1=2

C 1 X C 1
1
(

Dn11  Dn22 2 4D2 1=2

n
n
n
n 2
2 1=2
DC 01  C 1
C 02  C 1
1  D11  D22  D11  D22 4D
2 =2

exp  VX=1
)
n
n
n
n 2
2 1=2
DC 01  C 1
C 02  C 1
0
1
1  D11  D22  D11  D22 4D
2 =2
C2  C2 
exp  VX=2
Dn11  Dn22 2 4D2 1=2

C 2 X C 1
2
(

21

Dn11  Dn22 2 4D2 1=2

22

Usually it is assumed that the effect of diffusive interaction is small and the off-diagonal diffusion coefcients are smaller than the
diagonal coefcients by more than typically an order of magnitude. It is also reasonable to assume that for the ternary alloy the
difference between the main diffusional terms (i.e. between the on-diagonal terms) is considerable. Otherwise the diffusion of solute
1 and solute 2 would be identical and the alloy can be treated as a binary one. Under these assumptions jDj o o jD11 D22 j and Eqs. (21)
and (22) reduce to
0
1
n
C 1 X C 1
1 C 1  C 1 exp V X=D11

DC 02  C 1
2
exp  VX=Dn11  exp  VX=Dn22 
Dn11  Dn22

23

0
1
n
C 2 X C 1
2 C 2  C 2 exp V X=D22

DC 01  C 1
1
exp  VX=Dn11  exp  VX=Dn22 
Dn11  Dn22

24

As expected for this approximation the concentration of each solute represents a sum of the solute distribution without diffusional
interaction, Eq. (20), and the diffusional interaction term, which is proportional to D (last term in Eqs. (23) and (24)). In this case both

Concentration (at%)

Concentration (at%)

Nondimensional distance, X
Fig. 3. Solute concentration proles in the liquid ahead of the interface (C1) with strong diffusional interaction between the species at V o V2: a) Do 0, b) D 40. The strong
diffusional interaction between the species leads to a minimum (a) or a maximum (b) of solute concentration C1 depending on the off-diagonal diffusion coefcient. Dashed
lines no diffusion interaction between the species, solid lines strong diffusion interaction between the species.

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

51

compositions ahead of the interface change monotonically.


Now let us consider the case of relatively strong diffusional interaction between the species, i.e. when the on-diagonal and off-diagonal
diffusion coefcients are of the same order of magnitude. Fig. 3 shows examples of the most unusual behavior of the diffusion eld
calculated using Eqs. (21) and (22). It is observed that in the case of attractive diffusive interaction (Do0) there may be a minimum of
composition of solute 2 ahead of the interface but in case of repulsive interaction (D4 0) there may be a maximum of composition of
solute 1 ahead of the interface. However, it should be specied that in cases of weak diffusion interaction, both compositions only
change monotonically (see Eqs. (23) and (24)) and that a maximum or a minimum, like shown in the examples, can occur only in the
presence of relatively strong diffusional interactions between the species. The unusual behavior of the diffusion eld has been
discussed for local equilibrium case in more detail by Hunziker [6]. However, it should be noted that the results of [6] are only valid for
relatively small interface velocity when the solute diffusion occurs under local equilibrium conditions. At higher interface velocity the
effective coefcients of diffusion matrix Dnii are velocity dependent, which leads to qualitatively new behavior of the diffusion eld.
2. When V2 o VoV1, the eigenvalues of the diffusive matrix 1 40 and 2 o0
In this case the solute concentration ahead of the interface are given as
C 1 X C 1
1

C 2 X C 1
2

n
n
n
n 2
2 1=2
DC 02  C 1
C 01 C 1
2 D11  D22 D11  D22 4D
1 =2

Dn11  Dn22 2 4D2 1=2


n
n
n
n 2
2 1=2
DC 01  C 1
C 02 C 1
1  D11  D22  D11  D22 4D
2 =2

Dn11  Dn22 2 4D2 1=2

exp  VX=1

25

exp  VX=1

26

when diffusional interaction is ignored, i.e. D 0, these equations reduce to the solute distributions described in point 2 (Eqs. (17) and
(18)). This is to be expected since, without diffusional interaction, each diffusion eld is unaffected by the presence of the other and the
solutes diffuse independently. For relatively weak diffusion interaction when jDjo ojDn11  Dn22 j, Eqs. (25) and (26) reduce to
0
1
n
C 1 X C 1
1 C 1  C 1 exp  VX=D11

C 2 X C 1
2

DC 01  C 1
1
exp  VX=Dn11
jDn11  Dn22 j

DC 02  C 1
2
exp  VX=Dn11
jDn11  Dn22 j

27

28

In this case the concentration of solute 1 changes due to both the diffusion of solute 1 (i.e. its own diffusion), described by the two rst
term on the right hand side of Eq. (27) and the diffusional interaction with solute 2, which is proportional to D (the third term on the
right hand side of Eq. (27)). The concentration of solute 2 changes only due to diffusional interaction with solute 1 and, hence, the
diffusion layer of solute 2 in this case is determined by the diffusion layer of solute 1 (see Fig.4). It contrasts to solute concentrations
behavior at VoV2 when the diffusion layer of solute 1 is mainly proportional to Dn11 =V and the diffusion layer of solute 2 is mainly
proportional to Dn22 =V.
3. When V 4V2 4V1, the eigenvalues of the diffusive matrix i o0 and solute concentration eld ahead of the interface is described by
C 1 X C 1
1
C 2 X C 1
2
It implies that if the interface velocity is higher than the diffusive velocities, the solutes diffusion eld ahead of the interface remains
undisturbed and growth occurs in the diffusionless regime regarding to both solute 1 and solute 2.
4. Discussion
4.1. Solute redistribution at the interface
Solute partitioning at the solidliquid interface in the multi-component system has been discussed in [5] ignoring the effects of solute
diffusion in the bulk liquid. Such approximation is only valid for relatively low interface velocity when the solute diffusion occurs under
local equilibrium conditions. As demonstrated for binary alloy solidication [812], the local nonequilibrium diffusion effects in the bulk
liquid play an important role in solute redistribution at the interface leading to diffusionless solidication with complete solute trapping at
a nite interface velocity. We discuss here qualitatively the inuence of the diffusional interaction between the species on solute
partitioning in ternary system. This can be done more accurately by solving the model equations together with the kinetic rate equations
for phase transformation in multi-component system and this will be presented elsewhere.
In the case of independent diffusion the solute partitioning at the interface can be described by the following expression for the
velocity-dependent partition coefcient [12]
(
K Ei 1=1 Ai V=V Di  V ; V o V Di
K i V
1;
V 4 V Di
where KEi is the equilibrium partition coefcient of solute i, Ai is a constant such as VDi/(Ai 1) represents the characteristic velocities above
which Ki(V) deviates strongly from KEi [12]. The partition coefcients Ki(V) as a function of the interface velocity V is shown in Fig. 2.
Dashed lines represent the partition coefcients Ki(V) in case of independent diffusion D 0. Complete solute trapping Ki(V)1 is reached
for each solute independently at the nite interface velocities V VDi. The result, i.e. the complete solute trapping reached at a nite
interface velocity, was earlier obtained for binary alloys [812] and is consistent with experimental data (see [9,10,17]. Molecular dynamic
simulations also showed partitionless crystal growth at a nite interface velocity [15,16]. Yang et al. simulated a LeonardJones binary and

52

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

10

Concentration (at%)

4
2a
2
2b
0

10

12

14

Nondimensional distance, X
Fig. 4. Solute concentration proles in the liquid ahead of the interface at V2 oV o V1. Solid line 1 concentration of solute 1, solid line 2a concentration of solute 2 at
D 40, solid line 2b concentration of solute 2 at Do 0, dashed line concentration of solute 2 with no diffusion interaction and local nonequilibrium effects (shown for
comparison). Concentrations of solute 2 (solid lines 2) change only due to the diffusional interaction with solute 1 therefore the diffusional layer of solute 2 is equal to the
diffusional layer of solute 1 (solid line1).

a CuNi EAM model to show that the results hold for different systems [16]. Recently, Humadi et al. [17] have incorporated two time scales
into the phase-eld-crystal model of a binary alloy to explore different solute trapping properties as a function of the interface velocity.
The authors demonstrated that the introduction of wavelike dynamics in both density and concentration elds leads to complete solute
trapping at a nite interface velocity. The transition to complete solute trapping for non-planar interface has been considered in [18].
Let us consider solute partitioning at V2 oV oV1. In case of independent diffusion there is no diffusion of solute 2 ahead of the interface
1
and C 2 X C 02 C 1
2 (see point 2). For steady state planar growth the initial solute concentration in the bulk liquid C 2 is equal to the
solute concentration in the solid at the interface C 0S2 , which implies the complete solute trapping with the partition coefcient
K 2 C 0S2 =C 02 1. But if the diffusional interaction cannot be ignored (see Fig. 4), the diffusion eld of solute 1 affects the diffusion eld of
solute 2 and its concentration (as well as the concentration of solute 1) depends on the distance from the interface X and the solidication
cannot be considered as diffusionless regarding to solute 2. The concentration of solute 2 at the interface C 02 differs from that in the bulk
liquid far from the interface C 1
2 and, hence, K2 a1 even at V 4V2. It contrasts to the case of independent diffusion D 0 when the partition
coefcient K2(V)-1 at V-V2 and K2(V)1 at V4V2. Thus the diffusional interaction between the species changes qualitatively the solute
diffusion eld in multicomponent systems and affects the solute redistribution at the solidliquid interface.
4.2. Concentration dependence of diffusion coefcients
Most of the assumptions used in the present paper are the same as those used for the binary alloys. However, the assumption of
diffusion coefcients independent of composition seems to be more controversial for multicomponent alloys [1,4,6] and needs to be
discussed here in more detail. Coates et al. [4] and Hunziker [6] argued that whereas it is usually found that on-diagonal diffusion
coefcients approach constant values for dilute solutions, the off-diagonal coefcients show sensitive concentration dependence and must
vanish with the appropriate concentrations.
Therefore they suggest that an analytical description of the diffusion elds applies only to cases where the range of composition in
which diffusion occurs (between the far eld and the interface composition) is rather small. This occurs in particular in systems with
rather little partitioning between the liquid and the solid. If partition coefcients are sufciently close to unity (say within 0.8 to 1.2) then
there will be less than a 20% variation in the solute concentrations from the interface to the bulk liquid and a correspondingly small
variation in the off-diagonal diffusion coefcients [4,6]. In this situation the appropriate average values of the off-diagonal diffusion
coefcients can be used without introducing too much error. It is just the case of particular interest of the present work because the local
nonequilibrium effects play the most important role at high interface velocity V-VDi when Ki(V)-1 and solute partitioning at the
interface is relatively weak. A change in velocity only inuences the scale of the boundary layer in the X direction, but leaves the overall
shape of the composition elds identical. An open question remains regarding the impact of different diffusive speeds for different solutes
in multicomponent solute trapping models of rapid solidication.

5. Conclusion
Analytical solutions for the solute diffusion elds during steady-state plane front growth in multicomponent alloys have been
developed, taking into account both the diffusive interaction between the species and the local nonequilibrium diffusion effects in the bulk
liquid. Calculations for a hypothetical ternary system show some interesting effects of the diffusive interaction and diffusive relaxation on
the diffusion elds. If diffusional interaction between the species can be ignored, the transition to diffusionless solidication occurs in two
steps independently for solute 2 and solute 1 at V VD2 and VVD1, respectively. At relatively week diffusional interaction the solute
concentrations ahead of the interface change monotonically with the distance. But the strong diffusional interaction, when on-diagonal
and off-diagonal diffusion coefcients are of the same order, may lead to a maximum or a minimum of the solute concentration ahead of

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

53

the interface. Moreover the diffusional interaction changes the condition for solute trapping at the interface. For the time being however,
use of these solutions is limited by the lack of measured diffusion coefcients and methods to determine the off-diagonal terms. A
microscopic approach to this problem is needed.

Acknowledgments
The reported study was partially supported by the Russian Foundation for Basic Research, Research Project no. 13-02-91156 and
Research Project no.13-03-00726, the National Basic Research Program of China (973 Program, No. 2011CB610403), China National Funds
for Distinguished Young Scientists (No. 51125002), and the NSFC-RFBR Collaboration Project (NSFC no. 512111059).
Appendix A. (D12 D21 D)
Eq. (11) can be rewritten in the matrix form
!
!
d
d
Dn11 dX
V;
D12 dX
C n1
0
d
d
C n2
D21 dX
;
Dn22 dX
V
^
where C ni dC i =dX. Below we transform the matrix at d=dX to the diagonal form by orthogonal transformation W
(
)
!
n
D
D11
d
^
V H
D Dn22 dX
^  1H
^ W
^ W
^  1!
W
C 10
n

where
^
W

cos
sin

 sin

A:1

A:2

A:3
!

cos

and
tan 2 2D=Dn11  Dn22
 n

 D 
D 
 11

0
n
 D
D22  
The eigenvalues are given as Eq. (12). For new unknown vector-function we have two decoupled equations and the problem can be
solved analytically.
8
nn
< dC 1 C nn
10 exp  VX=1 cos  C 20 exp  VX=2 sin
dX
A:4
nn
nn
dC
: dX2 C 10 exp  VX=1 sin C 20 exp  VX=2 cos
Thus when D12 D21 D, the solution to Eqs. (10) and (11) takes the form

2 C nn
20 sin
exp  VX=1
exp  VX=2
V
V
nn
nn
1 C 10 sin
2 C 20 cos
exp  VX=1 
exp  VX=2
C 2 X C 1
2 
V
V
C 1 X C 1
1 

1 C nn
10 cos

A:5

nn
where C nn
10 and C 20 are coefcients, which depend on the boundary conditions.

Appendix B. (D12 aD21 )


In the case D12 a D21 we denote
(
2D D12 D21
2b D21  D12
^ takes the form
The operator H
(
)
!


n
D
D11
0 b d
d
^

V H
D Dn22 dX
b
0 dX

B:1

B:2

and equations for C can be written as


^!
H
C i 0
nn

To construct solutions to Eq. (B.3), we make preliminary orthogonal transformation of matrix at d=dX to the triangular form
)
(
!
1  2b d
!nn
V C 0
2 dX
0

B:3

B:4

54

S.L. Sobolev et al. / Journal of Crystal Growth 395 (2014) 4654

where
^ 1
W

Dn11

Db

D b

Dn22

!
^
W

1
0

 2b

!
B:5

and
21;2 Dn11 Dn22 7 Dn11  Dn22 2 4D2 b 1=2
2

!nn
As a rst step we solve Eq. (B.3) for C 2 , which is the rst order homogeneous differential equation. The second step is the solution of
!nn
!n
!nn
the rst order non-homogeneous differential equation for C 1 . And the nal step is the orthogonal transformation from C 1 to C 1
nn
!n
^!
C 1 W
C1

B:6

where
^
W
and

cos

 sin

sin

!
B:7

cos

8
< sin 2 20  Dn

11

: tan 20 n
D

2d
 Dn22 2 4D2 1=2

2D
 Dn11

22

Finally we have

2 C nn
20 2d cos

 sin exp  VX=2


V
V
1  2
1 C nn
2 C nn
10 sin
20 2d sin
exp  VX=1 

cos exp  VX=2


C 2 X C 1
2 
V
V
1  2
C 1 X C 1
1 

1 C nn
10 cos

exp  VX=1 

nn
where C nn
10 and C 20 are coefcients, which depend on the boundary conditions. If b-0, Eqs.(B.8) reduce to Eq. (A.5) with -.

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

M. Asta, C. Beckermann, A. Karma, W. Kurz, R. Napolitano, M. Plapp, G. Purdy, M. Rappaz, R. Trivedi, Acta Mater. 57 (2009) 941971.
F. Liu, G.C. Yang, Int. Mater. Rev. 51 (2006) 145.
U. Hecht, L. Granasy, T. Pusztai, B. Bottger, M. Apel, V. Witusiewicz, et al., Mater. Sci. Eng. R 46 (2004) 149.
D.E. Coates, S.V. Subramanian, G.R. Purdy, Trans. Metall. Soc. AIME 242 (1968) 800809.
A. Ludwig, Physica D 124 (1998) 271284.
O. Hunziker, Acta Mater. 49 (2001) 41914203.
K. Wang, H. Wang, F. Liu, H. Zhai, Acta Mater. 61 (2013) 13591372.
SL. Sobolev, Phys. Lett. A 199 (1995) 383386.
S.L. Sobolev, Phys. Rev. E 55 (1997) 68456854.
S.L. Sobolev, Acta Mater. 60 (2012) 27112718.
S.L. Sobolev, Phys. Lett. A 376 (2012) 35633566.
S.L. Sobolev, Mater. Lett. 89 (2012) 191194.
G.X. Wang, V. Prasad, Mater. Sci. Eng. A 292 (2000) 142148.
D. Jou, J. Casas-Vazquez, G. Lebon, Extended Irreversible Thermodynamics, Springer, Berlin, 2001.
Q. Yu, P. Clancy, J. Cryst. Growth 149 (1995) 4548.
Y. Yang, H. Humadi, D. Buta, B.B. Laird, D. Sun, J.J. Hoyt, et al., Phys. Rev. Lett. 107 (2011) 025505.
H. Humadi, J.J. Hoyt, N. Provatas, Phys. Rev. E 87 (2013) 022404.
S. Li, S.L. Sobolev, J. Cryst. Growth 380 (2013) 6871.

B:8

También podría gustarte