You are on page 1of 39

6.5.

EUTECTIC SOLIDIFICATION

201

is the temperature gradient in the liquid near the interface. This


where dT
dz zS
actual temperature is also shown in Fig. 6.29 for the case where there is a
region where:
T (z) < TL (z),
(6.15)

that is, the local temperature is lower than the liquidus temperature. This
condition is known as constitutional undercooling. In this case, a protrusion
which sticks out in front of the interface will have a lower temperature than
the local liquidus, and so will grow. Hence if the region in front of the
interface satises condition 6.15, dendrite formation will occur.
If the temperature gradient into the liquid is large enough, then condition 6.15 will never be met, and a planar interface will be stable. Mathematically, this stability condition is:

dT
dTL
>

dz zS
dz zS
Dierentiating Eqn. 6.14 we nd:

dT
v
>
(T1 T3 )
dz zS
D
The temperature range T1 T3 is known as the equilibrium freezing range
of the alloy. We see that systems with a large equilibrium freezing range
will have a greater tendency for constitutional undercooling and formation
of dendritic growth. The development of dendrites is shown in Fig. 6.30.

6.5

Eutectic Solidication

We next consider eutectic solidication, represented by:


L+
There is a diverse medley of possible two-phased microstructures which can
be produced by eutectic solidication, even in the relatively simple case of solidication along one direction at a constant velocity. These include lamella
or plate structures, rods, discontinuous structures, and irregular distribution
of phases. The behavior is determined by a variety of factors including the interface energies, diusion rate in the melt, volume fraction of the two phases,
preferred orientations for growth, and relative growth rates of the phases.

202

CHAPTER 6. SOLIDIFICATION

a)

b)

c)

d)

e)

important figure ....how


the dendrite formation
takes place...but looks
like lamellar structure

Figure 6.30: Schematic showing development of dendrites in alloy solidication. a) Planar interface. b) Protrusion develops, grows rejecting more
solute into the surrounding region. This slows growth near protuberance. c)
Dendrite grows more surrounding regions slow. d) Dendrites grow. e) Final
cellular microstructure.
For the purposes of illustration, we consider the unidirectional solidication at velocity v of a two-component eutectic liquid, resulting in a lamellar
structure comprised of alternating plates of A-rich -phase and B-rich phase. This process is illustrated in Fig. 6.31. Solidication of causes
buildup of B in the liquid in front of the growing interface. This excess of
B must diuse laterally in the liquid to the -phase solidication front. A
similar process occurs for the excess A in front of the -phase regions. The
resulting atomic uxes are shown schematically in Fig. 6.31.
One important aspect of two-phase solidication is the presence of interfaces between the phases. The energy contributions of these interfaces aect
growth morphology and rate. For example, consider the interfacial tension
equilibrium at the growing solid-liquid interface, as shown schematically in
Fig. 6.32. If we assume that the surface tension with the liquid is the same
for both and -phases, and we assume that the interfaces are in force
equilibrium, we nd:
= 2L cos
For many cases, the solid-liquid interface energy is much higher than that
between the two phases, that is L  , so that /2. If however,
the - interface energy is larger than twice the solid-liquid interface en-

6.5. EUTECTIC SOLIDIFICATION

203

ergy ( > 2L ), then interfacial tension equilibrium is not possible, and a


discontinuous microstructure will form.

JA

JB

Figure 6.31: Schematic of solidication of eutectic liquid into lamellar microstructure with lamella spacing , showing the ux of A atoms into the
growing -phase region and the ux of B atoms into the growing -phase
region.
Even when the interface force equilibrium allows formation of lamellar
structures, the growth rate and lamellar spacing reect the energy cost of
the - interface. The shaded volume element in Fig. 6.31, which has length
l, depth d and height , has interface energy 2ld , so the energy per volume
is:
2ld
2
=
ld

Hence the formation of this eutectic lamella from the melt results in a change
in free energy per volume of:
G() = Gsolid Gliquid = G() +

204

CHAPTER 6. SOLIDIFICATION

Figure 6.32: Schematic of interface


tension balance in eutectic growth.

Here G() is the bulk free energy change associated with the solidication
of the liquid to the eutectic solid, which can be approximated by the Turnbull
extrapolation:
HF T
G() =
TE
where HF is the (positive) heat of fusion for melting of the eutectic, TE is
the equilibrium eutectic temperature, and T is the undercooling given by
T = TE T . Note that at temperatures below TE , G() is negative,
but for small undercoolings G() may not be, since the energy cost of the
interfaces is a positive quantity. For a given lamellar spacing , the minimum
undercooling T1 required for eutectic solidication is found by setting
G()|T =T1 = 0
Solving for T1 we nd:

2 TE
HF
At undercoolings greater than T1 , where the temperature is below TE T1 ,
the free energy change G() is negative and eutectic solidication with a
lamellar spacing can occur.
Conversely, by similar arguments, we can nd the minimum spacing min
which is stable for a given level of undercooling. This is:
T1 =

min =

2 TE
HF T

(6.16)

This can be represented schematically in a couple of ways. If we plot the free


energy as a function of composition at a given temperature, we see that the
free energy of the solid phases is a function of the lamellar spacing. This is
shown in Fig. 6.33, where the two extremes are shown. For innite lamella
spacing, the free energy curves of the solid phases have a tangent line lying

6.5. EUTECTIC SOLIDIFICATION

205

below the liquid free energy curve by an amount G(). For the case where
= min the solid free energy curves share a tangent with the liquid free
energy curve. This raising of the solid phase free energy curves has the
eect of lowering the liquidus lines on the equilibrium phase diagram, since
solidication with a nite lamellar spacing will only occur at temperatures
below TE T1 . This is shown schematically in Fig. 6.34.

G( min)

G()

GL

G( min)

G()

G()

x
Figure 6.33: Schematic of free energy representation for eutectic solid. Shown
are the free energy for the liquid and bulk solid phases as a function of
composition. For a lamellar spacing of min , the free energy cost of the
interface formation raises the free energy curves of the solid phases to have
a common tangent with the liquid free energy curve.
During actual growth the undercooling T will be greater than T1
which is the minimum required for interface formation. If we write:
T = T1 + T2
then the excess undercooling T2 is that which is available to drive the other
growth processes, such as the atomic uxes mentioned earlier and shown
schematically in Fig. 6.31. For growth at lamella spacing greater than the
minimum allowed min , the free energy curves for the solid phases will be
raised only part of the way to the point of tangency with the liquid free
energy curve, as shown in Fig. 6.35a. The free energy driving force |G()|
provides an amount 2 / for interface formation and an amount:


2
2 2
1
1
|G()|
=
= 2

min

min

206

CHAPTER 6. SOLIDIFICATION

=
<

x
Figure 6.34: Schematic phase diagram for a eutectic system showing the
lowering of the liquidus lines for nite lamellar spacing.
for other growth processes, including driving the diusion ux. By using the
Turnbull extrapolation we can associate this driving force to the undercooling
T2 (= T T1 ):


HF T2
1
1
= 2

TE
min

Solving for T2 and inserting min from Eqn. 6.16 we nd:


T2

2 TE
=
HF

min

min
= T 1

(6.17)

To deal with the complicated problem of atomic ux in the concentration


proles in front of the growing eutectic interface we make some simplifying
assumptions. We rst assume that the liquid directly in front of the interface
is in equilibrium with the growing phase, that is, the atomic fraction of B
atoms in front of the -phase is xL and that in front of the -phase is xL ,
as shown in Fig. 6.35a. These have corresponding atomic concentrations
cL and cL , which are related to the atomic fraction by a factor of the
atomic volume. Hence there is a composition change along the interface of
c = cL cL . Further, we assume that the liquidus lines near the eutectic
are linear with slopes which are equal in magnitude but opposite in sign. This
allows us to deduce a simple relationship between the undercooling T2 and
the composition dierence. If we take m to be the magnitude of the slope of

6.5. EUTECTIC SOLIDIFICATION

207

GL

G ( min )
G ()

2
min

G()

xL

xL

G()
xL

xL

<

T1
T2
xL

xL

Figure 6.35: Schematic of the a) free energy and b) phase diagram for conditions of eutectic growth .

208

CHAPTER 6. SOLIDIFICATION

the liquidus line, and VA to be the (constant) atomic volume, then we have
(referring to Fig. 6.35b):
T2 =

m L
m
x xL = VA c
2
2

Solving for c we nd:


2T2
(6.18)
mVA
Hence there exists a concentration gradient along the interface which is
approximately the concentration dierence c divided by the distance over
which it occurs. That is:
dc
c

dy
/2
where y is measured along the interface. There will be a ux of atoms along
the interface in response to this gradient with magnitude given by:
c =

Jy =




dc

DL


dy

2DL c

(6.19)

where DL is the diusivity in the liquid in front of the growing interface. For
a given undercooling, the growth rate of the eutectic phase will be determined
by this ux. During a time increment dt, the ux will carry 2Jy dt B atoms
per area away from the front of the growing -phase region. During this
time, the interface will advance a distance dz which will expel (cE cL )
from the growing region, where cL is the concentration corresponding to
xL in Fig. 6.35a and cE is the eutectic concentration. Equating these two
B atom amounts we nd:

cE cL dz = 2Jy dt

where the factor of 2 accounts for the ux to adjacent slabs of -phase on


either side of the growing -phase. Solving for the growth velocity and using
Eqn. 6.19 for Jy , Eqn. 6.18 for c, and Eqn. 6.17 for T2 we nd:
2Jy
cL
4DL c
=
(cE cL )


8DL
min
T 1
=
1
(cE cL ) mVA

v =

cE

(6.20)

6.5. EUTECTIC SOLIDIFICATION

209

We can see that, for a given undercooling, there will be a peak in growth
velocity as a function of lamellar spacing. For large the diusion is slow
since the composition dierence is spread out over a large distance so the
composition gradient is small. Conversely, as approaches min the driving
force is entirely consumed by interface formation.
By dierentiating Eqn. 6.20 we can nd the lamellar spacing which corresponds to the fastest growth, which will be the most likely lamellar spacing
to occur. We nd:

dv

=0
d =

= 2min

Inserting this into Eqn. 6.20 above we nd maximum growth velocity:


2DL T 1
cL mVA min
DL HF (T )2
=
mVA TE (cE cL )
 2
1
16DL TE
=
E
L
HF mVA (c c )

vmax =

cE

so we see that the maximum velocity is proportional to the square of the


undercooling or to the square of the reciprocal of the lamellar spacing.

Chapter 7
Solid-Solid Transformations
We now turn to a discussion of solid-solid transformations. As an example, we
consider a single phase solid which is at a temperature and composition where
there would be two phases in equilibrium. An example is a precipitation
reaction  + where  is the super saturated -phase obtained from
quenching from a higher temperature, as shown in Fig. 7.1 a). A second
example would be a eutectoid transformation + as shown in Fig. 7.1
b).
The approach to equilibrium can occur by at least two mechanisms:
Nucleation and growth
Spinodal decomposition
These processes are shown schematically in Fig. 7.2. In nucleation and
growth, thermal uctuations form a nucleus of -phase with a sharp interface
between it and the surrounding  matrix. As the -phase particle grows, it
depletes the matrix of B atoms, and the growth of the -phase particle is
limited by diusion of B atoms down the concentration gradient in the surrounding matrix. In spinodal decomposition, small local uctuations develop
in the super saturated phase. These uctuations increase in amplitude as B
atoms diuse uphill toward higher B atom concentration levels. This is the
which, as you may recall, is
consequence of a negative diusion coecient D
given by:


d
ln

B
= (xA DB + xB DA ) 1 +
D
d ln xB
229

230

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

a)

+
c

b)

+
c

Figure 7.1: Phase diagrams for solid-solid transformations. a) Precipitation


reaction  + . b) Eutectoid reaction + .

231
where DA and DB are the tracer diusivities of the species A and B, and B
is the activity coecient of B in the A-B solution. For the case of a regular
solution in the quasi-chemical approximation we found:


= (xA DB + xB DA ) 1 Hmix
D
kT

where Hmix is the heat of mixing of the solution. If this is large enough,
that is, if unlike bonds are much less favorable than like bonds, then the
chemical diusion coecient can be negative, and uphill diusion can occur.
We will explore this further in future sections and problems.

a)
c

t1

t2

t3

t1

t2

t3

c0
c
b)
c
c0
c
Figure 7.2: Schematic of a) nucleation and growth and b) spinodal decomposition. In nucleation and growth the -phase particle starts out small and
grows as B atoms diuse down the concentration gradient to the - interface. In spinodal decomposition a small composition uctuation forms and
increases as B atoms diuse up the composition gradient by a process of
uphill diusion.

232

7.1

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

Chemical Concentration of the Precipitate

Consider the situation of a phase precipitate in an phase matrix as shown


in Fig. 7.3 which is a schematic of the free energy for the two phases and
as a function of the atomic fraction of B atoms, x. The equilibrium concentrations for the and phases are x and x . We are interested in
the composition of the phase precipitate which initially precipitates from
a supersaturated -phase of composition x0 . We consider the metastable
equilibrium between the the matrix of composition x1 and precipitate of
composition x1 . We allow the system to seek its lowest energy with respect to the composition of the matrix and precipitate, while xing the total
number of atoms in the precipitate, N , and matrix, N . A change in precipitate composition from x1 to x2 and the corresponding change in matrix
composition from x1 to x2 will have the change in energy:
*

G = [g (x2 ) g (x1 )] N + g (x2 ) g (x1 ) N

(7.1)

where g (x) and g (x) are the free energy per atom for and phases. For
innitesimal changes, we can write the dierence in free energies as:
dg
(x2 x1 ) =
dx
dg
g (x2 ) g (x1 ) =
(x2 x1 ) =
dx

g (x2 ) g (x1 ) =

dg
x
dx
dg
x
dx

(7.2)

We conserve the total number of atoms so that the number of B-atoms


added to the phase is equal to the number of B-atoms leaving the -phase
matrix.
NB = NB
(7.3)
Since x = NB /N we can see that:
NB = N x
NB = N x

(7.4)

(Note that we can see from Eqn. 7.4 that the change in composition in the
matrix is much smaller than that of the precipitate, since most of the atoms

7.1. CHEMICAL CONCENTRATION OF THE PRECIPITATE

233

x x x
x

x x x

Figure 7.3: Schematic of free energy of A-B binary system.


are still in the matrix, NB  NB .) Inserting Eqns. 7.2, 7.3 and 7.4 into
Eqn. 7.1, we nd:


dg dg
G =
NB
(7.5)

dx
dx
From Eqn. 7.5 we see that unless dg /dx = dg /dx, the system can always
lower its energy by changing the composition of the precipitate. For example
if:
dg
dg
>
(7.6)
dx
dx
then the system can lower its energy by transferring B atoms from the phase precipitate to the -phase matrix. This will continue until moving
down on the -phase free energy curve causes condition 7.6 to no longer be
true. Thus the composition which minimizes the energy is the one which

234

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

corresponds to the condition:


dg
dg
=
dx
dx

(7.7)

This is the so-called tangent construction.

7.2

Nucleation Energy Contributions

We consider the formation of a nucleus of -phase in a matrix of super


saturated  -phase. We have found in Section 7.1 that the concentration of
the precipitate will be found by the tangent construction. Now we turn to
consideration of the free energy change associated with formation of a nucleus
containing N atoms of this composition. For a solid-solid transformation,
the free energy can be written:
GT = N (gA + gel ) +

2/3

(7.8)

where gA is the dierence in free energy per atom between the supersaturated  phase and the phase, gel is the elastic strain energy per atom,
is the shape factor dened by:
=

A
(N )2/3

as in our treatment of nucleation in solidication, and is the interface free


energy. We rst turn to the free energy of the transformation.

7.2.1

Transformation Free Energy

The free energy dierence between the  matrix and the matrix with a phase precipitate is the driving force for the transformation. We refer to the
free energy schematic shown in Fig. 7.4 which shows the free energies of the
and phases as a function of atomic fraction x of B atoms. The total free
energy drop associated with the transformation is shown as an arrow from
the free energy of the original supersaturated  -phase at atomic fraction x0
to the common tangent line connecting the free energy curves of the and
phases at atomic fractions x and x .

7.2. NUCLEATION ENERGY CONTRIBUTIONS

235

g
A

B
A

x x
x

x x

Figure 7.4: Schematic of free energy for precipitation reaction.

236

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

We consider the formation of a -phase nucleus having N total atoms,


where the atomic fraction x1 in the precipitate satises the condition:

dg
dg


=
dx x1
dx x1
where x1 is the atomic fraction in the phase matrix after formation of
the -phase precipitate. This is the tangent construction which determines
the atomic fraction of the precipitate which minimizes the free energy. The
transformation free energy associated with the nucleus is:
Gtr N gA
=

N +N

(7.9)
g (x0 ) N g (x1 ) N g (x1 )

= N g (x0 ) g (x1 ) + N [g (x0 ) g (x1 )]


*

dg
(7.10)
dx
where g (x) and g (x) are the free energy of the and phases as a function
of atomic fraction of B atoms, and N is the number of atoms remaining in
the -phase matrix after forming the -phase precipitate. Consider the last
term in Eqn. 7.10. In order to identify gA , we would like to transform this
into a product of N and some other quantity. We can do this by observing
that:
N x0 = (N N )x0 = NB N x0
N g (x0 ) g (x1 ) + N (x0 x1 )

and
N x1 = NB N x1
since the number of B atoms in the -phase plus the number of B atoms
in the -phase must equal the total number NB of B atoms. Using these
relationships we nd:
N (x0 x1 ) = N (x1 x0 )
Inserting this into Eqn. 7.10 we nd:


N gA = N

dg
g (x0 ) g (x1 ) + (x1 x0 )
dx

Thus we identify gA as the quantity in brackets. This is shown schematically in Fig. 7.4. Hence, given the free energy functions g and g , we can
nd the composition of the precipitate, and the free energy per precipitate
atom associated with the transformation.

7.2. NUCLEATION ENERGY CONTRIBUTIONS

7.2.2

237

Strain Energy

One aspect of a solid-solid nucleation which is dierent from liquid-solid nucleation, is the possibility of elastic strain resulting from a change in volume
or shape during the  + transition. We consider a simple treatment
of the strain energy gel in Eqn. 7.8, and see that both nucleation rate and
nuclei shape can be aected.
The strain energy associated with this transformation can be examined
by the following sequence of imaginary operations.
Remove a small volume of and transform it to .
Apply surface tractions to this phase particle to allow it to return to
its original size and shape.
Stick it back in its original location and weld surfaces together.
Allow entire assembly to relax.
This places the sample of matrix and nuclei into a state of self stress, and
the strain energy must be taken into account in the free energy associated
with formation of this transformed region.
Nabarro has shown that if the interfaces are incoherent and all the strains
are in the matrix, the strain energy for a spheroidal precipitate with semi
axes c parallel to the rotation axis and b perpendicular, is given by1 :
gel =

2 V 2
(c/b)
3 V

where is the shear modulus of the matrix phase , V is the dierence in


atomic volumes of the two phases, V = V V , and (c/b) is a function
of the shape of the particle which has the following limits:

1
if c/b = 1 (sphere)
if c/b  1 (rod)
(c/b) = 3/4

3c/4b if c/b 1 (disk)


This is shown schematically in Fig. 7.5.
1

F.R.N. Nabarro, Proc. Roy. Soc. A 1175, 519, (1940)

238

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

sphere
(c/b)
1
0.75
rod

plate

c/b

Figure 7.5: Schematic of function (c/b).

7.3

Nucleation Barrier

To see how this can aect nucleation, lets examine a simple case of diskshaped precipitates with large semiaxis b, and small semiaxis c. The volume
and surface area are then:
4 2
b c
3
A 2b2

The shape factor is found by:


A
(N )2/3


V
= A
V

= 2b

2/3

V
4b2 c/3

= 2
where s = c/b.

1/3

3V
4s

2/3

2/3

7.4. GROWTH IN SOLID-SOLID TRANSFORMATIONS

239

The free energy associated with formation of a transformed region can


then be written as:


GT = N

3V
(V )2
gA +
s + 2 1/3
2V
4s

By setting:

2/3

(N )2/3

GT

=0
N N =N

we can nd the number of atoms in a critical nuclei, as before, but now we


must also minimize the energy with respect to the shape of the particle, that
is, we must set:

GT

=0
s s=s
and solve for s . After some algebra we nd:
N

s
G

32 3 (V )4 3 2
3gA5
gA V
=
(V )2
8 3 3 2 (V )4
=
3
gA4
=

(7.11)

Eqn. 7.11 shows that now the nucleation barrier has gA to the fourth power
in the denominator, which gives a much more sudden onset of nucleation
than the (gA )2 in the expression for homogeneous nucleation with no strain
energy eect. We also see that a large V will favor at disk-shaped nuclei
which can minimize the strain energy.

7.4

Growth in Solid-Solid Transformations

Growth in the solid phase has many characteristics of growth of solids from a
liquid. We still must have atoms crossing the interface and attaching to the
growing phase, so the atom attachment mechanisms discussed previously are
important. Transformation kinetics are usually limited by the relatively slow
atomic jump rates; it is rare for heat removal to be the limiting factor as it is
in some solidication processes. Therefore, we discuss isothermal solid-solid
transformations.

240

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

Under the condition of constant undercooling, there are two possible processes which can dominate the growth rate: interface attachment kinetics and
diusion. In interface attachment limited growth, the growth rate is limited
by the rate at which atoms can jump from the matrix to the growing phase.
The rate at which atoms are supplied to the interface region is suciently
fast so that this process does not limit growth. In this regime, the interface
moves with a constant velocity given by:


0 gT
GM
v=a
exp
kB T
kB T
for uniform growth or:

a2 xk 0 gT
GM
v=
exp
ys kB T
kB T

for nonuniform growth, where a is the distance grown per atomic layer, 0 is
the vibration frequency, gT is the driving force per atom for atomic jumps,
GM is the jump (migration) activation energy, xk is the fraction of step
sites which are kinks, and ys is the spacing between steps. These equations
are the same as those for solid growth into a liquid.
Interface attachment is likely to limit growth in phase transitions where
there is no composition change, as in crystallization of a metallic glass into
a compound of the same composition. However, in the more common case,
where the growing phase has a composition which diers from the matrix,
diusion of atomic species is likely to limit the growth rate. The growing
phase can grow only as fast as allowed by the ux of atoms to the interface.
Earlier in the course, we examined an example of this diusion limited growth
as a homework problem which dealt with a compound forming at the interface
of a diusion couple. We found that the width of the growing phase was linear
with the square root of time. This dependence is typical of diusion limited
growth and contrasts with interface attachment limited growth where the
interface position moves linearly with time (constant interface velocity).
As a second example of diusion limited growth, we consider growth of a
-phase spherical nuclei of composition c growing into an -phase matrix
of composition c0 (Fig. 7.6). The composition in the -phase at the interface
is c , and the interface movement is governed by the rate that the ux in the
-phase can provide atoms to drive the composition change at the interface:

c
dR (c c ) = J dt = D
dt
r

7.5. ISOTHERMAL TRANSFORMATION KINETICS

241

where R is the radius of the precipitate particle, D is the diusion coecient


in the -phase, and c is the composition in the -phase as a function of radial
coordinate r. Hence, the rate of change of the particle radius is:
v=

(c /r)|
dR
= D
dt
c c

With the spherical symmetry of this problem it is natural to work in spherical


coordinates. For this symmetric case, the diusion equation in spherical
coordinates is:


2
c
2

c
c

+
= D 2 c = D
t
r2
r r
and for a steady-state with the present boundary conditions, we nd:
c = c 0
Hence we nd:

R
(c0 c )
r

c
c0 c
=
r
R

and the interface velocity becomes:


v=
Integrating we nd:
R=

dR
D (c0 c )
=
dt
R(c c )
.
.
/2D

c0 c
t
c c

Thus for diusion limited growth, we again nd that the interface position
is proportional to the square root of time.

7.5

Isothermal Transformation Kinetics

Having considered separately the processes of nucleation and growth, we now


wish to describe the kinetics for a complete transformation which occurs by
a process of isothermal nucleation and growth. We consider a transformation
from an undercooled phase to phase. During this process, phase nuclei
will form in regions of phase. These phase particles will grow, consuming

242

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

c
c
c
0

10

r/R
Figure 7.6: Composition as a function of position for spherical precipitate.

the phase matrix. The number of phase particles which nucleate between
time t = and t = + d is given by:
IV d
where I is the nucleation rate per unit volume and V is the volume of
untransformed phase at time . If we assume that the growth is interface
limited, the volume, V , of a phase region which nucleated at time t =
is:

4 3
v (t )3 if t >
3
V =
0
if t <
where v is the growth rate. Actually, this assumes that the particle we are
considering does not impinge on other growing particles, since this would
halt new growth at the boundary between the two impinging particles. This
will only be the case at short times, where the average size of the growing
particle is small compared to the interparticle spacing. Within this short
time assumption, we can nd the total volume, Vst , of transformed region by
integrating over the nucleation times, .
4V  t 3
Iv (t )3 d
3 0
where we have assumed that the volume of the transformed region is small,
so that V = V , the entire sample volume. Assuming a constant nucleation
Vst =

7.5. ISOTHERMAL TRANSFORMATION KINETICS

243

rate, we nd for the transformed fraction, F :


F

Vst

= Iv 3 t4
V
3

(7.12)

This is good for short times only. As the particles grow they impinge on
each other, resulting in mutual interference. This presents a messy geometrical problem. In addition, as the transformation continues, the volume of
untransformed region is decreased, so that the region where nucleation can
occur is less than the sample volume, V .
The volume which we calculate by ignoring these eects is known as the
extended volume, Ve . It is larger than the actual transformed volume since
it includes:
Nuclei which form in already transformed region - these are known as
phantom nuclei.
Growth occurring in previously transformed regions, as growing particles impinge on each other.
The extended volume is what we could calculate to be the transformed volume if we:
Removed each nuclei from the sample as soon as it nucleates, lling the
hole left behind with untransformed .
Place the nucleated phase particle in an innite chunk of untransformed phase where no nucleation is occurring, thus letting it grow
unimpeded by other phase particles.
The expression for the extended volume is the same as that for the short
time volume found previously:
Ve =

4V  t 3
v I(t )3 d
3 0

(7.13)

As time goes on, the extended volume can get to be larger than the sample
volume!
How does the ability to calculate this extended volume help us? Well,
we can relate the extended volume to the actual transformed volume. In
a time interval the extended volume will increase by dVe while the actual

244

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

transformed volume will increase by dV . If we assume randomly located


nucleation sites, a fraction, V /V , of the increase in extended volume will
occur in previously transformed material, while a fraction, (1V /V ), occurs
in untransformed material. Only the fraction which occurs in untransformed
material will contribute to dV , hence:


V
dV = 1
V

Integrating we nd:

dVe

Ve

V
= V ln 1
V

So that for the transformed fraction we nd:

V
V
= 1 exp e
F =
V
V

For the case considered previously, where the nucleation rate is constant and
we have interface limited growth:


Iv 3 t4
F = 1 exp
3

(7.14)

Note that for short times, we recover Eqn. 7.12.


Lets consider a dierent case, where there are a xed number of randomly
distributed heterogeneous nucleation sites. In this case, the nucleation rate
is not constant. If we say that NN is the number of nucleation sites per unit
volume, then:
dNN = NN N dt
where N is the nucleation rate of a given site. We can then nd that:
NN = NN 0 exp (N t)
where NN 0 is the initial number of nucleation sites. The nucleation rate, I,
will be given by:
dNN
I=
= NN 0 N eN t
dt
Inserting this expression for nucleation rate at a given time into Eqn. 7.13
we nd:
 t
4 3
Ve
eN (t )3 d
=
v NN 0 N
V
3
0

7.5. ISOTHERMAL TRANSFORMATION KINETICS

245

Integration by parts three times yields:




N2 t2 N3 t3
8v 3 NN 0 N t
Ve
e

1
+

=
+
N
V
N3
2
6

(7.15)

Lets consider some limiting cases:


Slow Nucleation Rate (N t 1) We rst note that in this case:
I = NN 0 N eN t NN 0 N = constant
so that we should get the same result as for the constant nucleation
case, i.e. Eqn. 7.14. Expanding the exponential in the right hand side
of Eqn. 7.15 we get:
eN t 1 N t +

N2 t2 N3 t3 N4 t4

+
2
6
24

(7.16)

The rst four terms of Eqn. 7.16 are canceled by the other terms in the
brackets of Eqn. 7.15, so that we have:
Ve
V

so that:

8v 3 NN 0 N4 t4
=
N3
24
3 4
Iv t
=
3


Iv 3 t4
F = 1 exp
3

as we found previously in Eqn. 7.14 as we expected.


Rapid Nucleation Rate (N t  1) In this case the nuclei will be used
up quickly. In the brackets of Eqn. 7.15, we keep only the term which
is highest order in N t and nd:
8NN 0 v 3
Ve
=
V
N3
so that:

N3 t3
6

= NN 0

4 3 3
v t
3

4NN 0 3 3
v t
F = 1 exp
3

246

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

In this case all the nucleation occurs very early and so the volume transformed
is just due to the growing precipitates.
Avrami has proposed that we use as a general expression:
F = 1 exp ((kt)n )

(7.17)

where 3 n 4, with n = 3 corresponding to a nucleation rate which


decreases with time, and n = 4 corresponding to a constant nucleation rate.
As nal examples we consider 2-dimensional and 1-dimensional growth.
If the particles of the phase grow as disks of thickness d, we have for the
volume of a particle nucleated at time t = :
V = dv 2 (t )2
so that for the transformed fraction we have the same expression as in
Eqn. 7.17 but with:
2n3
For particles which grow only in one direction we nd the transformed fraction again given by Eqn. 7.17, but with:
1n2
Eqn. 7.17 is a fairly general expression for the transformed fraction as a
function of time for a nucleation and growth process. A typical example is
shown in Fig. 7.7.

7.6

Time-Temperature-Transformation Diagrams

We have a general expression for the volume transformed as a funcition of


time:


Ve
= 1 exp[(kt)n ]
F = 1 exp
V
where the rate constant k and Avarmi exponent n depend on the exact
transformation mechanism. For example for a constant nucleation rate and
interface limited growth, we have:
n = 4

k =

Iv 3
3

1/4

7.6. TIME-TEMPERATURE-TRANSFORMATION DIAGRAMS

247

1.0

0.6

1
2
3
4
5

0.8

0.4
0.2
0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

tk
Figure 7.7: Fraction transformed as predicted by isothermal transformation
kinetics where the fraction transformed is given by F = 1 exp[(kt)n ].
Shown are behaviors for several dierent values of n.

where I is the nucleation rate and v is the constant interface velocity. Similar
rate constants can be found for other nucleation and growth situations. We
have seen that the nucleation rate rises and then falls as a function of the
undercooling, being dominated rst by its increasing ability to overcome the
nucleation barrier resulting from the surface energy, and then by its decreasing ability to overcome the activation energy for atomic motion. Similarly,
the growth velocity v rst increases with increased uncercooling, as the driving force increases, and then decreases as atomic jumps atomic jumps are
frozen out. Hence, the reaction constant k will increase and then decrease
as a function of undercooling. At temperatures close to the transition temperature, the driving force will be small and rate of transformation will be
low. As the temperature decreases, the driving force and the transformation
rate will increase. As the temperature decreases further, the transformation
rate will again decrease as atomic jumps become unlikely due to the lack of
thermal energy.
This behavior can be summarized in a time-temperature-transformation
(TTT) graph which plots the temperature required for a given transformed
fraction as a function of the time required. This is shown schematically in
Fig. 7.8, where the boundary representing a given transformed fraction is
plotted on temperature-time plot. A given thermal history can be traced

248

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

on this transformation, for example an anneal where the temperature is increased at a steady rate and then held constant for a given amount of time
and then cooled rapidly to an temperature where the transformation is inactive.

T
Te

ln t
Figure 7.8: Time-temperature-transformation (TTT) diagram for a solidsolid transformation showing annealing paths.

7.7
7.7.1

Spinodal Decomposition
Spinodal Instability

So far we have considered transformations which occur by nucleation and


growth. This type of transition always occurs when the two phases have a
dierence in symmetry and even sometimes when the two phases dier only
by composition. However, there is a second type of transformation which can
occur if the two phases dier only in composition. This is known as spinodal
decomposition. Shown in Fig. 7.9 is the free energy versus composition for
a system where the two terminal phases have the same structure, but there
is a large positive heat of mixing between the constituents. Note that the
equilibrium conguration as given by the common tangent rule will be a two
phase mixture of distinct phases with compositions c1 and c2 . Between these
compositions there is a one phase mixture which is unstable with respect to
decomposition.
Consider the region between the compositions marked cs1 and cs2 . Within

7.7. SPINODAL DECOMPOSITION

249

A)
g

B)

Spinodal
Boundary

Figure 7.9: A) Schematic of free energy per volume versus composition for
a system where the terminal phases are the same structure, but there is a
large positive heat of mixing. B) Schematic of phase diagram exhibiting a
spinodal region.

250

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

this region, the curvature of the free energy is negative, that is:
2g
<0
c2
where g is the free energy per volume, and c is the composition. The
compositions,cs1 and cs2 , delineate this region, so, at these compositions:

2 g
2 g


=
=0
c2 c=cs
c2 c=cs
1

As shown schematically in Fig. 7.9, within this region, a single phase


sample is unstable with respect to small uctuations in composition. That
is, inside this region, a small uctuation in composition will lower the free
energy, while outside this region a small uctuation in composition will raise
the free energy. This region, where the curvature of the free energy is negative, is known as the spinodal region. A phase diagram is obtained from
making free energy curves like that shown in Fig. 7.9 for many temperatures,
and plotting the locust of points c1 (T ), c2 (T ), cs1 (T ), and cs2 (T ). Depending
on the relative position of the free energy of the liquid and solid, this can
result in either a solubility gap or a eutectic phase diagram. In either case,
the locust of points cs1 (T ), and cs2 (T ) dene the spinodal region of the phase
diagram. This is shown schematically in Fig. 7.9.
Up to this point, we have always described the free energy in terms of
variables such as the composition, temperature, pressure and so forth, and
in the case of multiphase samples we have included terms in the free energy
corresponding to surfaces or interfaces. Now however, we have the possibility
of having composition uctuations with no discrete interfaces. We must
include the eect on the free energy of composition gradients. Ignoring for
the time the eects of changes in the molar volume with composition, and
expanding the dependence of the free energy on the composition gradients,
we write for the free energy per volume:

g
(c)2 2 g
g(c, c) = g(c, 0) + c


+
+
c c=0
2 (c)2 c=0

(7.18)

where g(c, 0) is the free energy per volume of a homogeneous system of composition c. We know that the free energy of the system cannot depend on the
sign of the composition gradient, so the second term on the right of Eqn. 7.18
must be zero, and writing g  (c) = g(c, 0) and k = ( 2 g/(c)2 )/2 we have:
g = g  + k(c)2

(7.19)

7.7. SPINODAL DECOMPOSITION

251

The free energy, G, of the entire sample is found by integrating Eqn. 7.19
over the sample volume, V :
G=

 *

g  + k(c)2 dV

(7.20)

We can further expand g  (c) about some composition c0 to nd:


dg 
(c c0 )2 d2 g 


+
+
g  (c) = g  (c0 ) + (c c0 )
dc c=c0
2
dc2 c=c0

(7.21)

We assume a composition of the form:


c c0 = A cos z

(7.22)

where A is the composition wave amplitude, and is the spatial frequency of


the composition wave. This is a general assumption, since any composition
wave can be represented by a series of sinusoidal waves such as Eqn. 7.22. We
then plug Eqn. 7.22 into Eqn. 7.21 and perform the integration of Eqn. 7.20.
We nd for the free energy dierence between a system with composition
wave given by Eqn. 7.22 and a homogeneous system:


A2 d2 g 
G = V
+ 2k 2
4 dc2

(7.23)

If G < 0 then the system is unstable with respect to sinusoidal uctuations


with wavelength of 2/. In fact, whenever

d2 g 

<0
dc2 c=c0
the system is unstable with respect to sinusoidal composition uctuations
with a wavelength greater than some critical wavelength, c , found by setting
the term in the brackets in Eqn. 7.23 equal to 0.
2
c =
=
c

8 2 k
d2 g  /dc2

1/2

(7.24)

As can be seen from Eqn. 7.24, c as the spinodal boundary is approached.

252

7.7.2

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

Estimation of the Gradient Energy Term

Before proceeding it is interesting to estimate the relative size of the terms


involved in spinodal decomposition. One can show that in a simple nearest
neighbor bond counting model a coherent interface between two phases with
atomic fraction dierence x across the interface has an enthalpy per area
given by
H s = zFBB
(x)2
where z is the number of nearest neighbors, FBB is the fraction which are
across the given interface, is the atomic density at the interface (atoms/area)
and
is the bond enthalpy dierence

= HAB

1
(HAA + HBB )
2

where Hij is the ij bond enthalpy (assumed to be positive for a stable bond).
The enthalpy per volume is then just
H =

H s
d

where d is the interface width, or atomic plane spacing across the interface.
Recognizing that
1

=
d
VA
where VA is the atomic volume, we nd
zFBB
(x)2
H =
= zFBB
(c)2 VA
VA
where c is the concentration in atoms per volume and is related to x through
c = x/VA .
For a material with a composition gradient c, the composition dierence
across an interface will be given by
c dc
Hence a composition gradient will have associated with it an enthalpy per
volume of
H = zFBB
VA d2 (c)2

7.7. SPINODAL DECOMPOSITION

253

This allows us to identify the gradient energy parameter in the Cahn-Hilliard


spinodal decomposition theory as
k = zFBB
VA d2
Materials with a positive heat of mixing will have weaker unlike bonds and
hence a negative
and positive gradient energy term.
It is interesting to compare this with the term d2 g  /dc2 in the expression
A2
G = V
4

d2 g 
+ 2k 2
dc2

which is the free energy change associated with a sinusoidal composition


modulation of magnitude A and wavenumber . Using the regular solution
model for g  we nd


d2 g 
= VA 2z
+ kB T
dc2

1
1
+
1x x

!

Hence


d2 g 
+ 2k 2 = VA 2z
(1 FBB d2 2 ) + kB T
dc2

1
1
+
1x x

!

This leads to a critical wavelength given by


-

c .
4 2 FBB
.

=/
d
1 + kB T 1 + 1
2z

1x

Note that this has a minimum value of




min = 2d FBB
The behavior as a function of scaled bond parameter
is shown in Figure 7.10.

7.7.3

Strain Eect on Spinodal Instability

If the molar volume is a function of composition, the composition wave induces strain. This strain has an associated elastic energy, and thus raises
the free energy cost of the composition uctuation. This acts to stabilize the
homogeneous system, and to shrink the spinodal region.

254

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS


10

c/min

8
6
4
2
0
1

-z
/2kBT
Figure 7.10: Regular solution prediction of critical wavelength as a
function
of scaled bond parameter. The critical wavelength is scaled by 2d FBB .
We dene
to be the linear expansion per unit composition change, and
we can write the molar volume as:
V (c) = V0 [1 + 3
(c c0 )] = V0 (1 + 3A
cos z)
Lets examine the components of the stress free strain:

fxx =
fyy =
fzz = A
cos z
If we impose coherency on the system, then there can be no change in the
lattice parameter measured perpendicular to the composition wave, as a function of distance along the composition wave. The total strain,
t is given by:

tii =
fii +
eii
where
e is the elastic strain. In the case where we impose coherency, and
the composition wave is along the z direction, we have that:

txx =
tyy = 0
so that:

exx =
eyy = A
cos z

7.7. SPINODAL DECOMPOSITION

255

From isotropic elasticity analysis we nd:

exx = [xx (yy + zz )] /E

eyy = [yy (xx + zz )] /E

ezz = [zz (yy + xx )] /E


where is Poissons ration and E is Youngs modulus. Simple manipulation
yields:
2
cos z
1
A
cos z
=
E
1
A
cos z
E
=
1
= 0

ezz = A

xx
yy
zz

Having all the elastic strain and stress components allows us to nd the local
strain energy, E e (x):
E e (x) =

1
A2
2 E
ii
eii =
cos2 z
2 i
1

The average elastic energy per volume, E e is then:


1  1
ii
eii dV
V V 2 i

E e =

A2
2 E
2(1 )

which is independent of wavelength.


If we include this term in the free energy we nd:
 

G=
V

and:

2 E
g (c) +
(c c0 )2 + k(c)2 dV
1


G
2
2 E
A2 d2 g 
+
=
+ 2k 2
2
V
4 dc
1

256

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

Now the region of instability is given by:


2
2 E
d2 g 
+
=0
dc2
1
A larger negative curvature is necessary to make the homogeneous sample
unstable with respect to composition uctuations. In other words, strain stabilizes the homogeneous solution. The reduced spinodal region is sometimes
called the strain spinodal (Fig. 7.11).

Spinodal
Boundary

Solubility
Limit
Strain
Spinodal
c

Figure 7.11: Schematic of strain stabilized spinodal.


The critical wavelength is now given by:


c =

8 2 k
d2 g  /dc2 + 2
2 E/(1 )

1/2

(7.25)

So that we see that strain causes the smallest stable wavelength to become
larger.

7.7.4

Spinodal Decomposition Rate

Consider the change, G, in free energy associated with a small composition


change, c:
 

G =
V

dg 
dk
2
2 E
+
(c c0 ) + (c)2 c + 2kc(c) dV
dc
1
dc

7.7. SPINODAL DECOMPOSITION

257

Integration by parts yields:


 

G =
V

dk
dg 
2
2 E
+
(c c0 ) (c)2 2k2 c cdV
dc
1
dc

The term in the brackets is the chemical potential, and its gradient will drive
a diusive ux:


dg 
dk
2
2 E
2
2
J = M
+
(c c0 ) (c) 2k c
dc
1
dc
where M is the atomic mobility. Applying the continuity equation:
c
= J
t
and keeping only terms which are linear in c or its gradients, we nd:
c
=M
t

2
2 E
d2 g 
+
2 c 2M k4 c
dc2
1

(7.26)

We try a solution to Eqn. 7.26 of the form:


c c0 = A(, t) cos z
and nd:

A
2
2 E
d2 g 
+
2 A 2M k 4 A
= M
t
dc2
1
We now try a solution for A of the form:
A = A(, 0) exp [R()t]

and we nd for the spinodal rate, R:




R() = M 2

2
2 E
d2 g 
2
+
2
k
+
dc2
1

(7.27)


The behavior of a system can be divided into two regimes:


If R < 0 then composition uctuations decay exponentially with time.
This is the case if < c , and we have a homogeneous system which
is outside the strain spinodal and hence stable relative to composition uctuations. If the composition is still between the c1 and c2 in
Fig. 7.9, then the two phased system is lower in free energy than the homogeneous alloy, but the approach to equilibrium cannon take place by
spinodal decomposition, and must take place by a process of nucleation
and growth.

258

CHAPTER 7. SOLID-SOLID TRANSFORMATIONS

If R() > 0 then composition uctuations grow exponentially with


time. This will be the case inside the strain spinodal, where > c . A
composition uctuation will grow and eventually lead to a two phase
sample. The fastest growing composition uctuation wavelength, max ,
can be found by setting:

R()

=0
=max
and we nd:
max =

2c

that is, the fastest growing wavelength is


wavelength.

2 times the shortest stable

References
J.W. Cahn, Acta Met. 9, 795-801 (1961).
J.W. Cahn, Acta Met. 10, 179-83 (1962). boundary

You might also like