You are on page 1of 15

Vol. 46 No.

SCIENCE IN CHINA (Series E)

October 2003

Columnar to equiaxed transition during alloy solidification


LIN Xin1 ( ), LI Yanmin1,2 (), WANG Meng1 ( ),
FENG Liping1 (
), CHEN Jing ( ) 1 & HUANG Weidong1 ( )
1. State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi'an 710072, China;
2. School of Materials Science and Engineering, Shanghai Jiaotong University, Shanghai 200030, China
Correspondence should be addressed to Lin Xin (email: cgs@nwpu.edu.cn)
Received September 12, 2002

Abstract Considering the local linear superposition of the species and combining the calculation
of phase diagram, the columnar and equiaxed growth behaviours are investigated systematically
during solidification of multicomponent alloys. A theoretical model is developed to describe the columnar to equiaxed transition during multicomponent alloy solidification by taking account of the
competition between nucleation and growth ahead of a dendrite array, which shows a good
agreement with the experimental results.
Keywords: columnar-to-equiaxed transition, multicomponent alloy, solidification.
DOI: 10.1360/02ye0337

Dendrite is the most prevalent microstructural pattern during the preparation and formation
of materials. The microstructural scale and pattern of the dendrites largely determine the final
properties of the materials. In general, we desire the dendrite to grow with the mode of equiaxed
growth. Polycrystalline material obtained by equiaxed growth commonly possess higher strength
and ductility, because when the polycrystalline material is forced to deform, there exists mutual
compatibility of deformation among the crystal grains, and a large number of grain boundaries
inside the material can not only increase resistance of deformation, but also hinder the development of the cracks. For example, in the Super-Steel Plan of China at present, the researchers try to
make the microstructures fully equiaxed in the steels in order to obtain super-high strength steels.
However, in aerospace field, lots of key parts are used under high temperature environment, the
high temperature creep rupture usually originates from the grain boundaries. To improve the high
temperature properties of the parts, one wishes that there do not exist grain boundaries inside the
materials of these parts, or at least, there do not exist grain boundaries in the main force direction;
that is, the dendrite can grow with the mode of columnar growth during the preparation and formation of materials. Thus controlling the columnar and equiaxed growth during the solidification
of the materials is crucial to the final properties of the materials.
In order to realize the control of the columnar and equiaxed growth, it is necessary to understand the columnar to equiaxed transition (CET) mechanism during solidification, and make clear
the CET transition condition. Fundamentally, it is necessary to have a knowledge of the competition between nucleation and growth during solidification. Qualitatively, the CET occurs more eas-

476

SCIENCE IN CHINA (Series E)

Vol. 46

ily when an alloy has a high solute concentration, low pouring temperature (for casting), low
temperature gradient, high nucleation density in the melt and vigorous melt convection. However,
a quantitative understanding of the CET requires a thorough comprehension of all physical
mechanisms involved. In 1984, Hunt[1] first developed an analytical model to describe steady-state
columnar and equiaxed growth, qualitatively revealing the effects of alloy composition, nucleation
density and cooling rate on the CET. However, he used a very simple empirical relationship to
describe the variation of the undercooling with alloy composite and solidification rate. Cockcroft
et al.[2] used a more recent growth theory for the columnar and equiaxed growth but without considering high velocity non-equilibrium effects under rapid solidification. Recently, based on
Hunts CET model, Gumann et al.[3,4] developed a more perfect model by combining KGT
model[5] for direction solidification with LKT model[6] for the undercooling melt growth, with
high velocity non-equilibrium effects taken into account. Gumann et al. succeeded in applying
their model to epitaxial laser metal forming of single crystal.
As for numerical model, Flood and Hunt[7,8] first simulated the CET in a one-dimensional
ingot by incorporating Hunts CET model into a heat flow calculation, but they did not consider
the solute transfer. It should be indicated that the constitution undercooling resulting from the solute transfer has important effect on dendritic growth. In addition, they used different governing
equations to describe the columnar and equiaxed growth zones, but they did not strictly trace the
interface separating these two zones. By fully considering the solute transfer process, Wang and
Beckermann[9], and Gandin[10] used a more complex and rigorous front tracing algorithm to further
analyze the CET process. Brown and Spittle[11], Zhu and Smith[12,13], and Rappaz and Gandin[14]
adopted a completely different approacha probabilistic methodbased on the Monte-Carlo
procedure to predict the CET process. However, these models either lack a rigorous physical basis
or make certain important simplifications on dendritic growth and nucleation process. More recently, Nastac[15,16] developed a more comprehensive stochastic/probabilistic model to simulate
the evolution of dendritic growth and CET during the solidification of binary alloy through explicitly tracing the solid/liquid interface.
It should be indicated that most of the above models focused on binary alloy, and little attention is paid to multicomponent alloy. In this paper, via local linear superposition of the different
species and the calculation of phase diagram, a general theoretical model is developed to describe
the CET process under the solidification of multicomponent alloys.
1 Formulation of the model
In order to exactly understand the CET process during the solidification of multicomponent
alloy, we must have a grasp of the nucleation and columnar and equiaxed growth behaviour during
the alloy solidification. First, we should investigate the phase equilibrium of this alloy system. The
calculation of phase diagram is an important method for this purpose.

No. 5

COLUMNAR TO EQUIAXED TRANSITION DURING ALLOY SOLIDIFICATION

477

1.1 Calculation of phase diagram(CALPHAD)


Based on the thermodynamic principle, thermodynamic equilibrium at given temperature,
pressure and amounts of the species is determined by the minimum of the Gibbs free energy
G
= 0,
xi

(1)

or chemical potentials of the species are equal in all reacting phases

i1 = i2 = = ir ,

(2)

where xi is the molar fraction of species i, and superscript is phase index. Thus, to calculate and
determine the Gibbs free energy of the system at different temperature is very important for the
calculation of phase diagram.
The Gibbs free energy of a phase is commonly expressed as
m

i =1

i =1

G = xi oGi + RgT xi ln ( xi ) + exG ,

(3)

where Rg is the gas constant, and oGi the molar Gibbs free energy of pure species i with the
same phase structure as phase . On the right-hand side of eq. (3), the first term corresponds to the
Gibbs free energy of mechanical mixture, the sum of the first and the second terms corresponds to
the Gibbs free energy of an ideal solution, and the third term is the excess Gibbs free energy, expressed as[17]
ex

G =

m 1 m

i =1 j =i +1

xi x j Lij ,

(4)

where Lij is the binary parameter describing the interaction between species i and species j in
phase . Certainly, if necessary, we can also introduce the ternary interaction parameter xi x j xk Lijk
and multicomponent interaction parameter to make further amendment.
The partial molar free energy or chemical potentials of the species can be given by
m

G
x j

i = Gi = G x j
j =1

G
.
+
xi

(5)

The phase diagram of multicomponent alloys can be calculated by simultaneously solving


eqs. (3)(5) and eq. (1) or eq. (2) and the phase equilibrium parameters can also be acquired at
the meantime, e.g. the liquidus slopes mi = T L xiL , and the equilibrium partition coefficient
L

ki = xiS xiL , where T is the liquidus temperature.

1.2

Columnar growth
For low velocity cellular growth, the tip undercooling can be approximated by[18]
T = GD V ,

(6)

where G is the temperature gradient, D the diffusion coefficient of solute in the liquid, and V the

478

SCIENCE IN CHINA (Series E)

Vol. 46

solidification velocity. That is, cells are growing close to the limit of stability and there does not
exist the constitution undercooling ahead of the cellular front. However, for dendritic growth,
there always exists a constitution undercooling ahead of the dendritic front.
Assume that the dendritic tip satisfies the Ivantsov solution, i.e. the supersaturation i of
each species is given by

i = Iv( Pei ),

(7)

with
Ci* C0i

i =

Ci* (1 ki )

Pei =

(8)

VR
,
2 Di

(9)

Iv ( Pei ) = Pei exp ( Pei ) E1 ( Pei ) ,

(10)

where Ci* is the concentration of species i in the liquid at the dendritic tip, C0i the nominal concentration of species i, R the dendritic tip radius, Di the diffusion coefficient of solute in the liquid
of species i, Pei the Peclet number for species i, and E1 the exponential integral function.
As in KGT model[5] or Rappaz et al.s treatment[19], because of multiplicity of Ivantsov solution, in order to select a unique solution for dendritc growth, the marginal stability criterion[20] is
applied to the dendritic tip, i.e. the steady state dendritic tip radius R is equal to the critical wavelength s of the solid-liquid interface at the limit of stability[21]. The wavelength of the marginally
stable plane front is given by
m 1

mvi GCiC ( Pei ) GT

i =1

*s2

(11)

One obtains

*
R = m 1

mvi GCi C ( Pei ) GT


i =1

12

(12)

with
ki kvi 1 ln ( kvi ki )
mvi = mi 1 +
,
1 ki

GCi =

C ( Pei ) = 1

(1 kvi )VCi*
Di

(14)

2kvi

12

1 + 1 * Pe 2
i

(13)

,
1 + 2kvi

(15)

No. 5

COLUMNAR TO EQUIAXED TRANSITION DURING ALLOY SOLIDIFICATION

kvi =

ki + a0V Di
,
1 + a0V Di

479

(16)

where * is the stability constant given by 1 4 2 , the Gibbs-Thomson coefficient, GCi the
concentration gradient of species i in the liquid at the dendritic tip, mvi the velocity dependent liquidus slope of species i, kvi the velocity dependent partition coefficient of species i, a0 the characteristic length in the order of the atomic distance, and GT the effective temperature gradient. For
columnar growth, generally the latent heat effect is negligible. Considering the thermal conductivities and thermal diffusivities in the solid and in the liquid phases are the same, one has
L

GT = G L , where G is the liquid temperature gradient.

By simultaneously solving eq. (8) and eqs. (12)(16), one can obtain the variation in the
columnar dendritic tip radius and the concentration of each species in the liquid at the dendritic tip
against the processing parameters (V, GT).
The tip undercooling for columnar dendrites can be expressed by the sum of different
contributions
T = TC + Tr + Tk ,

(17)

where TC, Tr and Tk are the constitution undercooling, the curvature undercooling and the attachment kinetic undercooling respectively. They are expressed as
TC =

m 1

i =1

mi C0i mvi Ci* ,

(18)

2
,
R

(19)

, Tr =
Tk =

(20)

where k the linear kinetic coefficient. The dendritic tip temperature can be obtained by
T * = T0L T ,

(21)

where T0L , the equilibrium liquidus temperature at nominal concentration, can be obtained by
CALPHAD.
Fig. 1 gives the variation in the dendriteic tip radius and temperature with the solidification
velocity for the columnar growth of 316L stainless steel. For thermodynamic parameters and
thermophysical parameters of 316L stainless steel, please refer to tables 1[22] and 2[23]). It can be
seen that the dendritic tip radius and the tip temperature first decrease gradually with increasing
solidification velocity until they reach a minimum when the solidification velocity is close to
0.1m/s, then the tip radius increases sharply, and the tip temperature also starts to increase again.
Fig. 2 shows the relationship between the liquid concentration at the dendritic tip of Cr, Ni and the
solidification velocity for the columnar growth of 316L stainless steel. It can be known that the

480

SCIENCE IN CHINA (Series E)

Vol. 46

liquid concentrations of Cr, Ni at the dendritic tip increase with increasing solidification velocity.
At high velocity, close to 0.1m/s, the liquid concentration tends to increase sharply. It may be inferred that solidification system has approached the absolute stability when solidification velocity
is close to 0.1m/s.

Fig. 1. Variation of dendrite tip radius and tip temperature


with the growth velocity during unidirectional solidification

Fig. 2. Variation of liquid composition at the dendrite tip


with the growth velocity during unidirectional solidification

of 316L stainless steel (Fe17Cr12Ni), GT = 1 103 K/m.

of 316L stainless steel (Fe17Cr12Ni), GT = 1 103 K/m.

Table 1 Thermodynamic data of FeCrNi system(J/mol)[22]


Liquid
LFe,CrLiquid = (17737 + 7.997 T) + 1331(xFexCr)
LFe,NiLiquid = (16911 + 5.162 T) + (10180 4.147 T) (xFexNi)
LCr,NiLiquid = (318 7.332 T) + (16941 6.37 T) (xCrxNi)
LFe,Cr,NiLiquid = (100000 50 T)vFe + (140000 50 T)vCr + (80000 50 T)vNi
Bcc
LFe,CrBcc = 20500 9.68 T
LFe,NiBcc = (957 1.287 T) + (1789 1.929 T) (xFe xNi)
LCr,NiBcc = (17170 11.82 T) + (34418 11.858 T) (xCrxNi)
LFe,Cr,NiBcc = (18500 + 10 T)vFe + (6000 + 10 T)vCr + (27000 + 10 T)vNi
TcBcc = 1043xFe 311xCr + 575xNi + xFexCr(1650 550(xFe xCr)) + xCrxNi (2376 + 617(xCr xNi))
Bcc = 2.22xFe 0.008xCr + 0.85xNi 0.85 xFexCr + 4xCrxNi
Fcc
LFe,CrFcc = (10833 7.477 T) 1410(xFe xCr)
LFe,NiFcc = (12054 + 3.274 T) + (11082 4.45 T) (xFe xNi) 726(xFe xNi)2
LCr,NiFcc = (8083 12.88 T) + (33080 16.036 T) (xCr xNi)
LFe,Cr,NiFcc = 6500vFe + (10000 10 T)vCr + 48000vNi
TcFcc = 201xFe 1109xCr + 633xNi+ xFexNi(2133 682(xFe xNi)) 3605xCrxNi
Fcc = 2.1xFe 2.46xCr + 0.52xNi + xFexNi(9.55 + 7.23(xFe xNi) +5.93(xFe xNi)2 + 6.18(xFe xNi)3) 1.91xCrxNi

In order to obtain the undercooling profile ahead of the solidification interface, we must acquire the solute redistribution in front of the dendritic tip. By Ivantsov solution, one obtains

Ci ( z ) = C0i + Ci* C0i E1 Pei ( 2 z + R ) R E1 ( Pei ) ,

(22)

No. 5

COLUMNAR TO EQUIAXED TRANSITION DURING ALLOY SOLIDIFICATION

481

Table 2 The physical parameters of 316L stained steel used for calculation[23]
TL
CCr
CNi
kCr
kNi
mCr
mNi

0
DCr
0
DNi

QCr
QNi
a0
I0

liquidus temperature of Fe-17Cr-12Ni


concentration of chromium
concentration of nickel
partition coefficient for chromium
partition coefficient for nickel
slope of the liquidus surface with respect to chromium concentration
slope of the liquidus surface with respect to nickel concentration
Gibbs-Thomson coefficient
pre-exponential diffusion coefficient for chromium

1736.7K a)
17 wt.%
12 wt.%
0.89 a)
0.98 a)
3.00K/wt.% a)
1.08K/wt.% a)
1.9 107 Km
2.67 107 m2 s1

pre-exponential diffusion coefficient for nickel


pre-exponential diffusion coefficient for chromium
pre-exponential diffusion coefficient for nickel
length scale for solute trapping
nucleation pre-exponential factor
linear kinetic coefficient
thermal diffusion coefficient

4.92 107 m2 s1
6.69 104 J mole1
6.77 104 J mole1
5 109 m
1 1012 s1
2.33 K sm1
3.89 10 6 m2 s1

a) obtained from CALPHAD.

where z is the distance in the liquid from the dendritic tip parallel to the dendrite axis. By CALPHAD, one can obtain the equilibrium liquidus temperature distribution corresponding to the solute redistribution in front of the solidification interface. Assume a linear phase diagram. Then the
equilibrium liquidus temperature distribution corresponding to the solute redistribution in front of
the solidification interface can be given by

( ( )

T ( z ) = T0L + T L Ci* T0L E1 Pei ( 2 z + R ) R E1 ( Pei ) ,

(23)

where T L (Ci* ) is the equilibrium liquidus temperature at the concentration Ci* , to be determined directly from CALPHAD.
Considering a linear temperature field, the actual temperature distribution in the liquid ahead
of the solidification interface can be expressed by
Tq ( z ) = T * + GT z.

(24)

By combining eq. (23) with eq. (24), the


local undercooling profile ahead of the solidification interface can be obtained
T ( z ) = T ( z ) Tq ( z ).

(25)

Nucleation may happen if the maximum


local undercooling ahead of the interface is larger than the nucleation undercooling. Fig. 3
shows the undercooled zone ahead of the moving
solidification interface for 316L stainless steel. It
can be seen that, the undercooling of phase first

Fig. 3. Undercooled region ahead of the moving solidification interface for 316L stainless steel. V = 50 m/s, GT = 1
103 K/m.

482

SCIENCE IN CHINA (Series E)

Vol. 46

increases sharply to a maximum with increasing distance from the solidification interface, and
then smoothly decreases gradually.
1.3

Equiaxed growth
For equiaxed growth, the latent heat is dissipated through the liquid, leading to a thermal un-

dercooling Tt ahead of the solidification interface


Tt =

H
cpL

Iv ( Pet ),

(26)

where H is latent heat, cpL the specific heat of the undercooling liquid, Pet the thermal Peclet
number, similar to solute Peclet number, expressed by
Pet =

VR
2 L

(27)

L being the thermal diffusivity. Thus, the tip undercooling for equiaxed dendriteic growth can be
expressed as
T = Tt + TC + Tr + Tk .

(28)

The dendritic tip radius for equiaxed growth also can be calculated by eq. (12), with the temperature gradient GT positive for columnar growth, and negative for equiaxed growth. At the same
time, the controlling parameters are solidification velocity and the temperature gradient for columnar, and the melt undercooling for equiaxed growth. Therefore, before using eq. (12), the
temperature gradient ahead of the interface should be determined for equiaxed growth.
According to the interface heat flux equilibrium condition,
H V = K L G L ,

(29)

where K L = L CPL , is the thermal conductivity of the liquid, one obtains


GT =

GL
V H
= L L.
2
2 cp

(30)

Substituting eq. (30) into eq. (12), one obtains


R=

*
.
m 1 2 Pe m C
H
i vi 0i (1 kvi )
C ( Pei )
Pet
cpL
i =1 1 (1 kvi ) Iv ( Pei )

(31)

Fig. 4 shows the variation of the dendritic tip velocity and tip radius with the melt undercooling for equiaxed growth of 316L stainless steel. It can be seen that, the tip velocity increases
with the undercooling, and the tip radius first decreases sharply, passes a minimum and a local
maximum, and then decreases smoothly with increasing undercooling. Fig. 5 further shows each
partial contribution to the total undercooling with increasing melt undercooling. The constitution
undercooling is found to dominate in low undercooling range. However, with increasing melt undercooling, the ratio of thermal undercooling to the total undercooling increases gradually. And

No. 5

COLUMNAR TO EQUIAXED TRANSITION DURING ALLOY SOLIDIFICATION

483

finally thermal undercooling prevails in high undercooling range. These behaviours of different
undercoolings also explain why the tip radius first decreases sharply, passes a minimum and a local maximum, and then decreases gradually with increasing melt undercooling. Hence one can see
that, compared with other undercooling, the thermal undercooling can be neglected under low
melt undercooling, so it is possible to use the same model to describe the columnar and equiaxed
growth, in which only the object of solution is different, i.e. for equiaxed growth, the controlling
parameter is the undercooling, and solidification velocity and dendriteic tip radius are the objects
to be solved.

Fig. 4. Variation of solidification velocity and dendrite tip


radius with bath undercooling for 316L stainless steel.

Fig. 5.

Normalized thermal Tt/T, constitution TC/T,

Gibbs-Thomson Tr/T, kinetic Tk/T contributions vs total


bath undercooling T for phase solidification of 316L
stainless steel. 1, Tt; 2, Tc; 3, Tr; 4, TK.

1.4

Columnar to equiaxed transition (CET)


In general, if the maximum local undercooling ahead of the interface is larger than the nucleation undercooling, it may lead to nucleation and equiaxed crystal growth; however, if the
volume fraction of equiaxed crystals is too small, the equiaxed crystals will be incorporated into
the columnar crystal. Only when the volume fraction of the equiaxed crystals is greater than a certain value, can the equiaxed crystals block the columnar growth, and fully equiaxed growth be
present. Hunt[1] first suggested that, fully equiaxed growth may occur if the volume fraction of the
equiaxed crystals is greater than 0.49, whereas the structure is assumed to be fully columnar if it is
lower than 0.49%. This criterion was later confirmed by Brown and Spittle[11]. Therefore, in order
to ascertain the critical condition of CET, one should first determine the volume fraction of the
equiaxed crystals before they are wrapped up by the columnar front.
Based on the above consideration, one should first calculate the size the equiaxed crystal can
reach before it is wrapped by the columnar front. Fig. 6 shows the undercooled region ahead of the
columnar dendritic front and the possible equiaxed crystal growth zone, where the undercooling is
equal to the nucleation undercooling. The maximum radius re of the equiaxed crystal can be calculated out by[3]

484

SCIENCE IN CHINA (Series E)

Vol. 46

Fig. 6. Schematic representation of the undercooled region ahead of the moving solidification interface. zn is the distance from
the solid/liquid interface in the liquid.

re =

zn Ve
0

[ z ] dz,

(32)

where Ve, the equiaxed growth velocity, can be calculated by using local concentration and undercooling as determined with eqs. (22) and (25), and solving the above equiaxed growth model.
Since the nucleation sites are arranged randomly, introducing the concept of the extended volume
fraction, the actual volume fraction is obtained using Avarami equation[24]

= 1 exp [ e ] ,

(33)

where is the actual volume fraction, and e the extended volume fraction. Assuming that equiaxed growth goes on with a spherical mode, because the number of the total nucleation sites will
rapidly reach the number of heterogeneous nucleation sites once the nucleation undercooling is
reached, the extended volume fraction can be given by

e =

4 re3 N 0
,
3

(34)

where N0 is the total number of the heterogeneous nucleation sites per unit volume. According to
the volume fraction of the equiaxed crystals ahead of the columnar front under different growth
conditions, by combining Hunt criterion, the critical temperature gradient and solidification velocity of CET can be obtained.
2

Discussion

Fig. 7 shows the CET curves for different nucleation undercooling. It can be seen that the
nucleation undercooling has remarkable effect on the CET at low temperature gradient, but has
little effect on the CET at high temperature gradient. The lower the nucleation undercooling, the
lower the critical solidification velocity of CET, and the more easily the CET happens, when the

No. 5

COLUMNAR TO EQUIAXED TRANSITION DURING ALLOY SOLIDIFICATION

485

nucleation undercooling is zero, the CET curve becomes approximately a line. This is due to the
fact that under low temperature gradient and low solidification velocity, the undercooling ahead of
the interface is very low, and the nucleation rate is usually very high so long as the nucleation
undercooling is reached. Thus, the nucleation undercooling has a decisive effect on the CET. The
CET curves for different nucleation sites are given in fig. 8. It can be seen that, different from the
effect of the nucleation undercooling, the number of the nucleation sites has little effect on the
CET at low temperature gradient, but has remarkable effect on the CET at high temperature gradient, and with the increasing nucleation sites, the critical solidification velocity of CET decreases
rapidly. This is mainly because the number of nucleation sites determines the number of equiaxed
crystals which can grow when the nucleation undercooling is reached. The more the nucleation
sites, the more the equiaxed crystals, the larger the volume fraction of equiaxed crystal, and the
more easily the CET happens.

Fig. 7. Influence of the nucleation undercooling on the CET,


N0 = 109/m3.

Fig. 8. Influence of the number of nucleation sites on the


CET, Tn = 2.5 K.

Fig. 9 gives a comparison of the CET profile of 316L stainless predicted by the present
model with the experimental results from directional solidification by Poole and Weinberg[25]
and from laser rapid forming. The model shows
a good agreement with the experimental results.
It should be indicated that, during laser rapid
forming, from the bottom of the melt pool to the
top, the temperature gradient decreases continuously, and the solidification velocity increases gradually, so the solidification parameters of laser rapid forming present a continuous
range, as shown by shaded zone in fig. 9. In the
experimental parameter range of laser rapid

Fig. 9. The CET profile of 316L stainless steel, Tn = 2.5 K,


N0 = 1012/m3. , experimental result from ref. [25]; , results
from the laser rapid formation.

486

SCIENCE IN CHINA (Series E)

Vol. 46

Fig. 10. Solidification microstructure of 316L stainless during laser rapid forming. (a) Longitude cross section; (b) top cross
section at the arrow in (a).

forming, the solidification morphology all lie on


the columnar growth zone, the CET does not
happen. Fig. 10 illustrates the solidification microstructures at the middle and top of the laser
forming part of 316L stainless steel, both of
which do not show the equiaxed growth. Fig. 11
gives a comparison of the CET profile of Rene
95 superalloy predicted by the present model
with the experimental results on laser rapid
forming. The parameters used in the calculation
are listed in tables 3 and 4. The present model
Fig. 11. CET profile of Rene95 superalloy, Tn = 2.5 K, N0
also shows a reasonable agreement with the ex- = 1012/m3. Shadow area shows results from the laser rapid
perimental result of Rene95 superalloy, i.e. from forming.
the bottom of the melt pool to the top, with the decreasing temperature gradient and the increasing
the solidification velocity, there is a columnar to equiaxed transition (fig. 12).
3 Conclusions
(1) Based on the local linear superposition of the species and the calculation of phase diagram, the columnar/equiaxed growth behaviour of multicomponent alloy is systematically studied.
A theoretical model is developed to describe the columnar to equiaxed transition during the solidi-

No. 5

COLUMNAR TO EQUIAXED TRANSITION DURING ALLOY SOLIDIFICATION

487

fication of multicomponent alloy by taking account of the competition between nucleation and
growth ahead of a dendrite array.
Table 3 Thermodynamic data of CoCrNi system(J/mol)[22,26,27]
Liquid
LCo,CrLiquid = (11900 + 2.4 T) + (6500 + 0.1 T) (xCo xCr)
LCo,NiLiquid = 1331
LCr,NiLiquid = (318 7.332 T) + (16941 6.37 T)(xCr xNi)
Fcc
LCo,CrFcc = ( 23080 + 8.34 T) + (12370 12.08 T) (xCo xCr)
LCo,NiFcc = (800 + 1.2629 T)
LCr,NiFcc = (8083 12.88 T) + (33080 16.036 T) (xCr xNi)
Table 4 The physical parameters of Rene95 superalloy used for calculation[28]
1709.1 Ka)

TL

liquidus temperature of Ni-13Cr-8Ni

CCr

concentration of chromium

13 wt.%

CCo

concentration of cobalt

8 wt.%

kCr

partition coefficient for chromium

0.92 a)

kCo

partition coefficient for cobalt

1.03 a)

mCr

slope of the liquidus surface with respect to chromium concentration

mCo

slope of the liquidus surface with respect to cobalt concentration

1.75 K/wt.% a)
0.3 K/wt.% a)

DCr

Gibbs-Thomson coefficient

1.9 107 Km

diffusion coefficient for chromium

2.67 107 m2/s

DCo

diffusion coefficient for cobalt

5.34 107 m2/s

a0

length scale for solute trapping

5 109 m

I0

nucleation pre-exponential factor

linear kinetic coefficient

1 1012 l/s
2.88 K/(sm)

thermal diffusion coefficient

3.89 106 m2/s

a) obtained from CALPHAD

Fig. 12. Solidification microstructure of Rene95 superalloy during laser rapid forming. (a) Microstructure near the interface
between the substrate and the cladding layer; (b) microstructure of the outer layer of forming sample.

(2) The nucleation undercooling has a remarkable effect on the CET at low temperature gra-

488

SCIENCE IN CHINA (Series E)

Vol. 46

dient and low solidification velocity. The critical velocity of CET decreases with the decreasing
nucleation undercooling. The nucleation at cooling has little effect on the CET under high temperature gradient.
(3) The nucleation site has a remarkable effect on the CET at high temperature gradient and
high solidification velocity, the velocity of CET decreases with an increase in the nucleation sites.
The nucleation site has little effect on the CET at low temperature gradient.
(4) The present model is in good agreement with the experimental results.
Acknowledgements This work was supported by the National High Technology Research and Development Program of
China (Grant Nos. 2001AA337020 and 2002AA336050), the National Natural Science Foundation of China (Grant No.
50201012) and National Science Fund for Distinguished Young Scholars of China (Grant No. 59825108).

References
1.

Hunt, J. D., Steady state columnar and equiaxed growth of dendrites and eutectic, Mater. Sci. Eng., 1984, 65: 7583.

2.

Cockcroft, S. L., Rappaz, M., Mitchell, A. et al., An examination of some of the manufacturing problems of large single-crystal turbine blades for use in land-based gas turbines, Materials for Advanced Power Engineering (eds. Cout-

3.

Gumann, M., Trivedi, R., Kurz, W., Nucleation ahead of the advancing interface in directional solidification, Mater. Sci.

souradis, J. et al.), New York: Kluwer Inc., 1994, 11611175.


Eng., 1997; A226228: 763769.
4.

Gumann, M., Bezenon, C., Canalis, P. et al., Single-crystal laser deposition of superalloys: processing-microstructure
maps, Acta Mater., 2001, 49: 10511062.

5.

Kurz, W., Giovanola, B., Trivedi, R., Theory of microstructural development during rapid solidification, Acta Metall. Ma-

6.

Lipton, J., Kurz, W., Trividi, R., Rapid dendrite growth in undercooled alloys, Acta Metall., 1987, 35: 957964.

7.

Flood, S. C., Hunt, J. D., Columnar and equiaxed growth I. A model of a columnar front with a temperature dependent ve-

ter., 1986, 34: 823830.

locity, J. Crystal. Growth, 1987, 82: 543551.


8.

Flood, S. C., Hunt, J. D., Columnar and equiaxed growth II. Equiaxed growth ahead of a columnar front, J. Crystal.
Growth, 1987, 82: 552560.

9.

Wang, C.Y., Beckermann, C., Prediction of columnar to equiaxed transition during diffusion-controlled dendritic alloy solidification, Metall. Mater. Trans. A, 1994, 25: 10811093.

10.

Gandin, C. A., From Constrained to unconstrained growth during directional solidification, Acta Mater., 2000; 48: 2483

11.

2501.
Brown, S. G. R., Spittle, J. A., Computer simulation of grain growth and macrostructure development during solidification,

12.

Zhu, P., Smith, R. W., Dynamic simulation of crystal growth by Monte Carlo method I. Model description and kinetics,

Mater. Sci. Technol., 1989, 5: 362368.


Acta Metall., 1992, 40: 683692.
13.

Zhu, P., Smith, R. W., Dynamic simulation of crystal growth by Monte Carlo method II. Ingot microstructures, Acta Metall., 1992, 40: 33693379.

14.

Rappaz, M., Gandin, C. A., Probabilistic modeling of microstructure formation in solidification process, Acta Metall.,

15.

Nastac, L., Stefanescu, D. M., Stochastic modelling of microstructure formation in solidification processes, Model. Simul.

1993, 41: 345360.


Mater. Sci. Engng., 1997, 5: 391420.
16.

Nastac, L., Numerical modelling of solidification morphologies and segregation patterns in cast dendritic alloys, Acta Ma-

17.

Chen, S. L., Oldfield, W., Chang, Y. A. et al., Modeling solidification of turbine blades using theoretical phase relation-

ter., 1999, 47: 42534262.


ships, Metall. Mater. Trans. A, 1994, 25: 15251533.

No. 5

COLUMNAR TO EQUIAXED TRANSITION DURING ALLOY SOLIDIFICATION

489

18.

Hunt, J. D., Lu, S. Z., Numerical modeling of cellular/dendritic array growth: spacing and structure predictions, Metall.

19.

Rappaz, M., David, S. A., Vitek, J. M. et al., Analysis of solidification microstructures in Fe-Ni-Cr single-crystal welds,

20.

Langer, J. S., Muller-krumbhaar, H., Theory of dendritic growth I. Elements of a stability analysis, Acta Metall., 1978, 26:

Mater. Trans. A, 1996, 27: 611623.


Metall. Trans. A, 1990, 21: 17671782.
16811687.
21.

Ivantsov, G. P., Temperature field around a spherical, cylindrical and acircular crystal growth in a supercooled melt, Dokl.

22.

Miettinen, J., Thermodynamic reassessment of Fe-Cr-Ni system with emphasis on the iron-rich corner, Calphad, 1999, 23:

Akad. Nauk, SSSR, 1947, 58: 567569.


231248.
23.

Takeuchi, S., The Properties of Liquid Metals, London: Taylor and Francis, 1973, 343347.

24.
25.

Porter, D. A., Easterling, K. E., Phase Transformations in Metal and Alloys, 2nd ed., London: Chapman and Hall, 1992.
Poole, W. J., Weinberg, F., Observations of columnar-toequiaxed transition in stainless steels, Metall. Mater. A, 1998, 29:

26.

Guillermet, A.F., Assessment of the thermodynamic properties of the Ni-Co system, Z. Metallkde, 1987, 78: 639647.

27.

Oikawa, K., Qing, G. W., Ikeshoji, T. et al., Thermodynamic calculations of phase equilibria of Co-Cr-Pt ternary and magnetically induced phase separation in the FCC and HCP phase, Journal of Magnetism and Magnetic Materials, 2001, 236:

28.

Brandes, E.A., Smithells Metalls Reference Book, 6th ed. Bodmin: Butterworths, 1983.

855861.

220233.

You might also like