You are on page 1of 91

Solidification Processing

1 The very first step: Nucleation


2 1 The very first step: Nucleation

The process of nucleation is the formation of a critical cluster of atoms


which can eventually grow. The process is ”stochastic” because the for-
mation of the initial cluster of atoms, its size and number are statistically
determined depending on the activation barriers for the process. Physically,
there are two forces involved in the formation of the nucleus. First, is
the driving force for the formation of a crystalline state which is different
from the non-crsystalline state in the liquid. This is a positive force, for
undercoolings greater than zero. However, the formation of a nucleus
involves the generation of a surface between the parent liquid and the
daughter nuclei which forms. This is a negative contribution to the total
energetics determining the formation of the nuclei. Formally this term
which relates, to the surface of interaction between the solid and the liquid
is referred to as the capillary force. The equality of the driving force
for solidification and the negative capillary force happens at the critical
state, also called the nucleus, since any size beyond this critical cluster
will eventually grow as a result of the dominant driving force.

1.1 Thermodynamic driving force for first-order


phase transition

The driving force for phase change is expressed as a difference of the


Gibbs-free energy between the phases. For example, a pure solid has a
driving force to transform to the liquid when the temperature is below the
melting point where the free energy of the liquid is higher as compared to
the liquid.

𝑔(𝑇 )
𝑇 ̸= 𝑇𝑚
Δ𝑔 = 𝑔𝑙 − 𝑔𝑠 =Driving force
(𝑔𝑠 = 𝑔𝑙 ) 𝑔𝑠
𝑔𝑙
𝑇
(𝑇𝑚 )
1.1 Thermodynamic driving force for first-order phase transition 3

In general however, the relation of the driving force which is mechanically


the pressure difference across the interface is related to the difference of
the relevant thermodynamic potential in the phases under consideration.
For a pure material, the first law for a system of N particles reads

𝑑𝑈 = 𝑇 𝑑𝑆 − 𝑃 𝑑𝑉 + 𝜇𝑑𝑁 (1.1)
𝑑𝑈 = 𝑑(𝑇 𝑆) − 𝑆𝑑𝑇 − 𝑃 𝑑𝑉 + 𝜇𝑑𝑁 (1.2)
𝑑(𝑈 − 𝑇 𝑆) = −𝑆𝑑𝑇 − 𝑃 𝑑𝑉 + 𝜇𝑑𝑁. (1.3)

𝑈 is the internal energy, 𝑆 is the entropy 𝑇 is the temperature, 𝑃 is the


pressure and 𝑉 the volume. We note that 𝑈 − 𝑇 𝑆 is the definition of the
Helmholtz free energy 𝐹 . The following relation follows from the first law
of thermodynamics which writes,

𝑑𝐹 = −𝑆𝑑𝑇 − 𝑃 𝑑𝑉 + 𝜇𝑑𝑁. (1.4)

However, from the integral form of the first law, it follows,

𝑈 = 𝑇 𝑆 − 𝑃 𝑉 + 𝜇𝑁 (1.5)
𝐹 = 𝑈 − 𝑇 𝑆 = −𝑃 𝑉 + 𝜇𝑁. (1.6)

Inserting into the previous relation we have,

−𝑉 𝑑𝑃 = −𝑆𝑑𝑇 (1.7)
𝑑𝑃 = 𝑠𝑑𝑇 + (𝑁/𝑉 )𝑑𝜇 (1.8)

where 𝑠 = 𝑆/𝑉 , is the entropy density and (𝑁/𝑉 ) is the inverse of the
atomic volume. This relation essentially relates the deviation of pressure
to the change in temperature and the chemical potential. Performing
another transform 𝐹 + 𝑃 𝑉 = 𝐺 = 𝜇𝑁 gives,
4 1 The very first step: Nucleation

𝑑(𝐹 + 𝑃 𝑉 ) = 𝑑𝐺 = 𝑉 𝑑𝑃 − 𝑆𝑑𝑇 + 𝜇𝑑𝑁. (1.9)

At equilibrium, the free energies of both the phases are equal, i.e. 𝐺𝛼 = 𝐺𝛽 .
This is thermodynamic potential which is most accessible in experiments
since T, P are variables which are easy to control. However, calculations
might sometime be more elegant with other potentials.
Also, it is easy to note that in conditions of constant pressure and volume
of the total system, the potentials 𝑑𝐹 = 𝑑𝐺 = 𝑑(𝜇𝑁 ).
Another interesting potential which is sometimes elegant to use, is the
”grand-potential” with T, V and 𝜇 as its natural variables. This is arrived
at, by performing the Legendre Transform of the Helmholtz free energy of
the system which writes,

𝑑Ψ = 𝑑(𝑈 − 𝑇 𝑆 − 𝜇𝑁 ) = 𝑑(−𝑃 𝑉 ) = −𝑆𝑑𝑇 − 𝑃 𝑑𝑉 − 𝑁 𝑑𝜇, (1.10)

where Ψ = (𝐹 − 𝜇𝑁 ) is the grand-potential of the system. It is convenient


to use this potential for certain calculations of effective driving forces
since it has T, V and 𝜇 as its state variables which are involved in the
assumption of local thermodynamic equilibrium at the interface, i.e the
phases are at the same chemical potential and temperature. It is also
interesting to note that written as a grand-potential density, it is equal
to the pressure −𝑃 = 𝑓 − 𝜇𝜌, where 𝜌 is the number density and 𝑓 is the
Helmholtz free energy density and 𝜇 is the chemical potential.

1.1.1 Calculating the driving force

The driving force at an interface arises due to a mechanical imbalance


of forces at the interface i.e a pressure difference. Consider two phases
initially at equilibrium, i.e same chemical potential 𝜇, same 𝑇 and pressure
𝑃 . Let, us say we perturb the system by changing the temperature 𝑇 just
a little bit say by Δ𝑇 . This would disturb the equilibrium among the
phases, and we wish to know the effective driving force induced due to this
1.1 Thermodynamic driving force for first-order phase transition 5

1
change in temperature. Denoting 𝑁/𝑉 = , Ω is the atomic volume and
Ω
assuming conditions of local thermodynamic equilibrium (same chemical
potential 𝜇𝛼 = 𝜇𝛽 ) and temperature 𝑇 . From Eqn. 1.8, we derive for a
pure material,

1
𝑑𝜇𝛼 = 𝑑𝑃 𝛼 − 𝑠𝛼 𝑑𝑇 (1.11)
Ω
1
𝑑𝜇𝛽 = 𝑑𝑃 𝛽 − 𝑠𝛽 𝑑𝑇. (1.12)
Ω

Since, the perturbation is from a equilibrium state, and we assume local


thermodynamic equilibrium, we derive the relation,

𝑑 𝑃 𝛼 − 𝑃 𝛽 = 𝑠𝛼 − 𝑠𝛽 𝑑𝑇.
(︀ )︀ (︀ )︀
(1.13)

Integrating, we get the difference of pressures and the effective driving


force as,

𝑃 𝛼 − 𝑃 𝛽 = Δ𝑠𝛼𝛽 Δ𝑇. (1.14)

Note, since we are very close to equilibrium, the change in entropy can be
𝐿𝑓
assumed to have the equilibrium value which is nothing but , where
𝑉𝑚 𝑇𝑚
for simplicity, we have assumed that the molar volumes of both phases
are equal to 𝑉𝑚 , 𝐿𝑓 is the latent heat of fusion and 𝑇𝑚 is the melting
temperature. Thus Δ𝑠𝛼𝛽 Δ𝑇 = 𝐿𝑓 − 𝑇 Δ𝑠𝛼𝛽 , which is the Gibbs-free
energy density, Δ𝑔 𝛼𝛽 . Together with what we have previously derived,
that the grand-potential density is the pressure, gives us the relation:
Δ𝑃 = Δ𝑔 = Δ𝜓.
6 1 The very first step: Nucleation

1.2 Gibbs-Thomson effect: quantification of the


capillarity effect

Consider a system consisting of pure solid and liquid with a planar interface
between them. The condition that the system will continue to remain in
equilibrium, occurs when the system temperature is at the melting point.
Now consider, a curved spherical solid nucleus, which is in equilibrium
with the liquid. Can this equilibrium be attained at the melting point
of the liquid? The answer is no. The reason can be understood though
an analogous case of an air bubble which can sustain its shape given the
pressure difference confirms to certain relation with the surface tension
and the radius of the bubble. Consider a small sector of the bubble in
two dimensions, with small finite depth of 𝑑𝑙. Then, due to symmetry, the
horizontal forces would be equal, and the balance of the vertical forces
would require

𝑇 𝜃 = Δ𝑃 𝑟𝜃𝑑𝑙
𝜎𝑑𝑙𝜃 = Δ𝑃 𝑟𝜃𝑑𝑙
1
𝜎 = Δ𝑃,
𝑟

This is the fundamental Gibbs-Thomson equation i.e 𝜎𝜅 = Δ𝑃 , 𝜅 being


the interfacial curvature. One can verify that for the case of a sphere a
similar analysis would give the same result with 𝜅 being 2/𝑅, which is the
sum of the two principal curvatures.

In general, the notion of curvature for small incremental volumes can be


visualized as the ratio of the incremental area to an incremental change in
Δ𝐴
volume i.e. 𝜅 = . Therefore, the work to create a surface 𝜎Δ𝐴, can
Δ𝑉
be equivalently written as 𝜎𝜅Δ𝑉 , which at equilibrium would equal to the
pressure work done by the system i.e. Δ𝑃 Δ𝑉 . Therefore, equating the
two would consequently give us the Gibbs-Thomson relation which reads
𝜎𝜅 = Δ𝑃 .
1.3 Homogeneous Nucleation 7

1.3 Homogeneous Nucleation

A given cluster of atoms of a phase 𝛼 is critical when it is in equilibrium


with another phase 𝛽 which is the parent phase. As we have just seen
equilibrium between two phases, exists when the the chemical potential
and temperature are equal such that there is no mass transfer between the
phases. Additionally, we have equilibrium of the interface which should
have mechanical equilibrium given by the Gibbs-Thomson equation. An
event of nucleation is called as a homogeneous nucleation event when the
cluster of atoms forms independently/unsupported in the melt. Assume,
that we have a cluster which is reasonably close to a sphere, such that the
mechanical equilibrium condition for this phase to exist would be,

2𝜎
Δ𝑃 = 𝜎𝜅 = . (1.15)
𝑅

Also, assume that we are presently considering a single component in the


system.
One can now use the relation in Eqn. 1.14, to derive the critical un-
dercooling at which a given size of radius 𝑅 is critical. The relation
writes,

2𝜎
Δ𝑇 = . (1.16)
𝑅Δ𝑠𝛼𝛽

One can also derive the above by writing down the total energy of the
system, which writes,

4
Δ𝐺 = − 𝜋𝑅3 Δ𝑔 + 4𝜋𝑅2 𝜎. (1.17)
3

At equilibrium, the total force of the system should go to zero. From


mechanics, we can derive that the total force on the system is nothing
8 1 The very first step: Nucleation

𝜕Δ𝐺
but . One can work out that we indeed land-up at the same Gibbs-
𝜕𝑟
𝜎
Thomson condition as before which reads Δ𝑔 = Δ𝑃 = 2 . Substracting
𝑅
the energy of critical nucleus from the energy of the uniform phase 𝛽 from
which the nucleation took place, gives us the activation barrier or the
barrier to forming a nucleus of 𝛼 homogeneously in 𝛽.
It can be shown that the barrier to nucleation, in terms of 𝜎 and Δ𝑔 and
eliminating 𝑅*, the critical radius derives as,

16𝜋𝜎 3 16𝜋𝜎 3 𝑇𝑚
2 2
𝑉𝑚
Δ𝐺* = 2 = 2 (1.18)
3 (Δ𝑔) 3𝐿2𝑓 (Δ𝑇 )
2𝜎𝑉𝑚 𝑇𝑚
𝑅* = . (1.19)
𝐿𝑓 Δ𝑇

Therefore, the barrier to nucleation reduces for larger undercoolings while


the critical radius varies inversely with undercooling with smaller and
smaller clusters becoming critical with increase in undercooling.

1.4 Alloys
For the case of alloys, the thermodynamic driving force can be calculated
similar to what was performed in the case of the pure components. We can
start by writing the Gibbs-Duhem relation for the case of alloys where for
simiplicity, we start with the case of a binary alloy composed of components
A and B, which reads as,

𝑉 𝑑𝑃 = 𝑆𝑑𝑇 + 𝑁𝐴 𝑑𝜇𝐴 + 𝑁𝐵 𝑑𝜇𝐵 , (1.20)

where 𝑁𝐴 is the number of particles of A, 𝑁𝐵 is the number of particles


in B, with the total number of particles being 𝑁 . 𝜇𝐴 and 𝜇𝐵 are the
𝜕𝐺 𝜕𝐺
respective chemical potentials which are defined as and , with
𝜕𝑁𝐴 𝜕𝑁𝐵
𝐺 being the Gibbs-free energy. The relation can also be reformulated as,
1.4 Alloys 9

𝑉 𝑑𝑃 = 𝑆𝑑𝑇 + 𝑁𝐴 (𝑑𝜇𝐴 − 𝑑𝜇𝐵 ) + 𝑁 𝑑𝜇𝐵 . (1.21)

We will define 𝜇 = 𝜇𝐴 − 𝜇𝐵 as the diffusion potential, as the gradient of


this term, determines the mass flux of A atoms. Also, at equilibrium the
value of the potential is uniquely defined 𝜇𝑒𝑞 (𝑇, 𝑃 ) for every temperature
given a pressure of the system.

1.4.1 Calculating the driving force for the case of alloys

Consider two binary alloys 𝛼 and 𝛽 in thermodynamic equilibrium, which


𝛽 𝛽
implies that the system is such that 𝜇𝛼 𝛼 𝛼 𝛽
𝐴 = 𝜇𝐴 , 𝜇𝐵 = 𝜇𝐵 , 𝑇 = 𝑇 and
𝛼 𝛽
𝑃 = 𝑃 . As in the case of the pure materials, let us now perturb the
system a little bit such that we change the chemical potential 𝜇𝑒𝑞 by
Δ𝜇. We would like to derive the effective change in the pressure due to
this change in the chemical potential, while assuming conditions of local
𝛽
thermodynamic equilibrium 𝜇𝛼 𝑖 = 𝜇𝑖 , 𝑖 = (𝐴, 𝐵) and constant T. Using
the Gibbs-Duhem relation, one can write the deviation of pressure in each
phase as,

𝑉 𝛼 𝑑𝑃 𝛼 = 𝑁𝐴𝛼 (𝑑𝜇𝐴 − 𝑑𝜇𝐵 ) + 𝑁 𝛼 𝑑𝜇𝐵 (1.22)


𝑉 𝛽 𝑑𝑃 𝛽 = 𝑁𝐴𝛽 (𝑑𝜇𝐴 − 𝑑𝜇𝐵 ) + 𝑁 𝛽 𝑑𝜇𝐵 . (1.23)

Dividing throughout the first relation by 𝑁 𝛼 and the second by 𝑁 𝛽 , and


𝑉𝛼 𝑉𝛽
noting that 𝛼 = Ω𝛼 and 𝛽 = Ω𝛽 which are the atomic volumes in each
𝑁 𝑁
phase respectively and substracting the latter from the first we have,

(︁ )︁
𝛽
Ω𝛼 𝑑𝑃 𝛼 − Ω𝛽 𝑑𝑃 𝛽 = 𝑐𝛼𝐴 − 𝑐𝐴 𝑑𝜇, (1.24)

𝛽
with 𝑐𝛼
𝐴 and 𝑐𝐴 as the mole fraction of A atoms. The above differential
equation can be understood as a difference in pressure work in phase 𝛼
10 1 The very first step: Nucleation

and that in phase 𝛽. The above can be integrated to derive a relation


between the 𝑃 𝛼 and 𝑃 𝛽 . The pressure in one of the phases needs to be
known, such that the pressure in the resulting phase can be fixed with
respect to this chosen pressure reference. The integrated equation derives
as,

(︁ )︁
𝛽
Ω𝛼 𝑃 𝛼 − Ω𝛽 𝑃 𝛽 − Ω𝛼 − Ω𝛽 𝑃𝑒𝑞 = 𝑐𝛼
(︀ )︀
𝐴 − 𝑐𝐴 (𝜇 − 𝜇𝑒𝑞 ) . (1.25)

If the 𝛼 phase is in contact with the surroundings such that the pressure
in the 𝛼 phase is known and fixed, we can derive the driving force for the
phase transformation as,

1 (︁ 𝛽
)︁
𝑃 𝛼 − 𝑃 𝛽 = 𝛽 𝑐𝛼
(︀ )︀
𝐴 − 𝑐 𝐴 (𝜇 − 𝜇𝑒𝑞 ) . (1.26)
Ω

𝜇𝛽𝐵 (𝑃 𝛼 ) − 𝜇𝛼 𝛼
𝐵 (𝑃 )
The right-hand side of this equation equals 𝛽
=
Ω
𝜇𝛽𝐴 (𝑃 𝛼 ) − 𝜇𝛼 𝛼
𝐴 (𝑃 )
. Thus a deviation of the diffusion potential (𝜇 − 𝜇𝑒𝑞 )
Ω𝛽
would result in a deviation of pressure under local thermodynamic equilib-
rium which is the driving force for phase transformation, the magnitude
of which is the deviation in the chemical potential of either component
computed at the same pressure divided by the atomic volume.

1.4.2 Gibbs-Thomson condition: Deviation of the


equilibrium chemical potential

Equating the difference in pressures to 𝜎𝜅, will be enable us to derive the


deviation in the diffusion potential 𝜇, resulting from a pressure difference
of Δ𝑃 , which balances the capillarity surface term. The change in 𝜇
derives as,
1.4 Alloys 11

𝜎𝜅Ω𝛽
Δ𝜇 = . (1.27)
𝑐𝛼
𝐴 − 𝑐𝛽𝐴

The corresponding change in compositions can be derived from the ther-


𝜕𝜇𝛼
modynamic variation of . It can be shown that the deviation in
𝜕𝑁𝐴
compositions of the phases are respectively,

𝜎𝜅
Δ𝑐𝛼,𝛽 = 2 )︁ . (1.28)
𝜕 𝑔𝛼,𝛽 (︁
𝛽
𝑐𝛼
𝐴 − 𝑐𝐴
𝜕𝑐2𝐴

The Gibbs-Thomson equation can be visualized as follows


solid solid 𝜇˜𝐵 (˜
𝜇)
𝑔(𝑐) 𝑔(𝑐) ˜ ̸= 𝜇
𝜇 ˜𝑒𝑞
liquid liquid
𝜇𝑒𝑞
Δ˜ 𝐴 = 𝜎𝜅
(𝜇𝑒𝑞
𝐵)

𝑐 𝑐 𝜇
˜
𝜇𝑒𝑞 )

𝜇𝑒𝑞
(˜ 𝐴) 𝜇𝑒𝑞
(˜ 𝐴)
𝑇 𝑇 Δ𝜇˜𝐴 = (𝑐𝑠 − 𝑐𝑙 )Δ𝜇 = 𝜎𝜅
𝜎𝜅
Δ𝑐𝑙 = 2 𝑙
𝜕 𝑓
2
(𝑐𝑠 − 𝑐𝑙 )
⏟𝜕𝑐⏞
𝑚 𝜎
⏟ ⏞𝑙
𝑐𝑠 𝑐𝑙 𝑐 𝑐𝑠 𝑐𝑙 𝑐 Γ𝑠𝑙 =
𝜕2𝑓 𝑙
2
(𝑐𝑠 − 𝑐𝑙 )
Gibbs-Thomson effect ⏟𝜕𝑐⏞
12 1 The very first step: Nucleation

1.5 Probability of nucleation


Nucleation is a stochastic event. The probability of a given cluster of
critical size 𝑟* to form is related to its activation barrier or energetic
barrier to form that size Δ𝐺* . The number of clusters a given size 𝑟 can
be assumed to vary as,

(︂ )︂
Δ𝐺𝑟
𝑛𝑟 = 𝑛𝑜 exp , (1.29)
𝑘𝑇

where Δ𝐺𝑟 is the total energy of the cluster of atoms of size 𝑟. Above
𝑇𝑚 , the relation is valid for all 𝑟, however for temperatures below 𝑇𝑚 , the
relationship is valid for only 𝑟 < 𝑟*, since beyond 𝑟*, the cluster is part of
the solid. The number of clusters is indicative of the probability for the
cluster to form. Small clusters are formed in plenty and the probability of
formation of larger cluster reduces with the radius.
Having chosen an acceptable probability for the formation of a cluster, one
can then determine the maximum radius 𝑟𝑚𝑎𝑥 which can form with this
chosen probability, in the liquid. If 𝑟𝑚𝑎𝑥 is lower, than 𝑟* , no nucleation
event of acceptable frequency would occur. For 𝑟𝑚𝑎𝑥 greater than 𝑟*,
there will be nucleation. With increase in undercooling, we have seen
that the critical 𝑟* decreases while the possibility to form a larger cluster
𝑟𝑚𝑎𝑥 increases. As a result, there is a critical undercooling beyond which
the possibility to form clusters greater than 𝑟* exists. The value for this
critical undercooling is generally Δ𝑇 = 0.2𝑇𝑚 for metals, which is about
200K.
One can also argue from the viewpoint of energetics. The probability of
achieving a critical energy Δ𝐺* required )︂for the formation of nucleus at a
Δ𝐺*
(︂
given undercooling varies as exp . Again, choosing an acceptable
𝑘𝑇
probability, gives a critical value of the energy Δ𝐺* . As we have seen that
most metals have a critical undercooling of 0.2𝑇𝑚 , this can be inserted
in the expression to compute the barrier to nucleation Δ𝐺* and can be
shown to be approximately 78kT.
2 Transport processes in
solidification
14 2 Transport processes in solidification

2.1 Diffusion and convection

In a conventional casting apparatus, all modes of heat transfer are operating.


Firstly, there is conductive heat transfer in the bulk of solid and the liquid.
The heat transfer from the mold to the environment, is a combination of
conduction, convection and radiative heat transfer, which are dependent
on the contact properties of the mold, the thermophysical properties of the
fluid environment in contact with the mold, and the temperature difference
between the mold and the environment. An empirical way to account for
all of these, is to decribe a quantity called the heat transfer coefficient that
will need to be calibrated every time we change the materials in contact
or the surrounding environment. The heat flux, given a temperature
difference between two materials on account of the different heat transfer
processes can be written as,

𝑞 = ℎ (𝑇𝑤 − 𝑇∞ ) (2.1)

where 𝑇𝑤 is the wall temperature and 𝑇∞ is the temperature of the envi-


ronment in which it is contact.

The diffusion of heat inside the bulk-solid and liquid are described by the
conservation equations. To derive these laws consider a infinitesimally
small control volume element of volume Δ𝑉 = 𝛿𝑥𝛿𝑦𝛿𝑧. With, a density
of 𝜌 and a specific heat capacity of 𝐶𝑝 , the increase in heat of this small
element can be derived from the balance of the influx of heat - outflux of
heat. The relation writes as

[︁
Δ (𝜌𝐶𝑝 𝑇 ) 𝛿𝑥𝛿𝑦𝛿𝑧 = (𝐽𝑥 − 𝐽𝑥+𝑑𝑥 ) 𝛿𝑦𝛿𝑧 + (𝐽𝑦 − 𝐽𝑦+𝑑𝑦 ) 𝛿𝑥𝛿𝑧+
]︁
(𝐽𝑧 − 𝐽𝑧+𝑑𝑧 ) 𝛿𝑥𝛿𝑦 𝛿𝑡 (2.2)
[︂ ]︂
Δ (𝜌𝐶𝑝 𝑇 ) (𝐽𝑥+𝑑𝑥 − 𝐽𝑥 ) (𝐽𝑦+𝑑𝑦 − 𝐽𝑦 ) (𝐽𝑧+𝑑𝑧 − 𝐽𝑧 )
=− + + ,
𝛿𝑡 𝛿𝑥 𝛿𝑦 𝛿𝑧
(2.3)
2.1 Diffusion and convection 15

where 𝐽𝑖 , represents the flux in the 𝑖th direction at the plan situated
distance 𝑖 in the direction 𝑖. In the limit 𝛿𝑥, 𝛿𝑦, 𝛿𝑧 → 0 and 𝛿𝑡 → 0, we
have

𝑑 (𝜌𝐶𝑝 𝑇 )
= −∇ · J. (2.4)
𝑑𝑡

Here J = (𝐽𝑥 , 𝐽𝑦 , 𝐽𝑧 ) and 𝐽𝑥 , 𝐽𝑦 and 𝐽𝑧 are the currents in the respective


directions given by the indices. The relation between the heat currents J
and the state variable 𝑇 , are derived through a constitutive relation. In
general the flux of a given quantity 𝐽 is proportional linear combination
of the gradients of all the thermodynamic potentials in the system, i.e,

𝜕𝑆 𝜕𝑆
J𝑒 ∝ 𝑀𝑒𝑒 ∇ + 𝑀𝑒𝑐 ... (2.5)
𝜕𝑒 𝜕𝑐

Fourier law of heat conduction is one such cause effect relationship, wherein,
the flux of heat is proportional to the gradient in temperature, with the
constant of proportionality being the thermal conductivity in the system.
Therefore, the flux of heat writes as,

J𝑒 = −𝐾∇𝑇, (2.6)

where 𝐾 is the thermal conductivity. A similar cause effect relationship


for the case of mass transport writes as,

∑︁
J𝑐𝑖 = − 𝑀𝑖𝑗 ∇𝜇𝑗 (2.7)
𝑗

where 𝜇 is the chemical potential and 𝑀𝑖𝑗 is an element from the mobility
matrix.
Substituting, Fourier’s law into the balance equation for heat we derive,
16 2 Transport processes in solidification

𝑑 (𝜌𝐶𝑝 𝑇 )
= ∇ · (𝐾∇𝑇 ) . (2.8)
𝑑𝑡

Similarly, a balance law for the mass transport can also be written down
for the case of multi-component alloy solidification, which writes as,

⎛ ⎞
𝜕𝑐𝑖 ∑︁
=∇·⎝ 𝑀𝑖𝑗 ∇𝜇𝑗 ⎠ . (2.9)
𝜕𝑡 𝑗

Along, with heat and mass transfer processes there are natural convective
currents which could arise in the liquid ahead of the solidification front.
The motion of the liquid leads to advection of heat a mass along with the
fluid motion. The motion of the liquid itself, can be described by solving
the momentum balance equations, which are the Navier-Stokes equations
which writes as,

𝜕𝜌v
+ v · ∇ (𝜌v) = −∇𝑝 + ∇ · T + 𝑓, (2.10)
𝜕𝑡

where T is the stress tensor, 𝑝 is the local pressure, 𝑓 are the body forces.
For newtonian, incompressible fluids the equations of motion for the the
different velocity vectors can be simplified into,

𝜕𝜌v
+ v · ∇ (𝜌v) = −∇𝑝 + 𝜇∇2 v + 𝑓, (2.11)
𝜕𝑡

𝜇 being the viscosity in the melt. Including, advection in the diffusion


equations for heat and mass transport are modified as,
2.2 Stefan boundary condition 17

𝜕 (𝜌𝐶𝑝 𝑇 )
+ v · ∇ (𝜌𝐶𝑝 𝑇 ) = ∇ · (𝐾∇𝑇 ) (2.12)
𝜕𝑡 ⎛ ⎞
𝜕𝑐𝑖 ∑︁
+ v · ∇ (𝑐𝑖 ) = ∇ · ⎝ 𝑀𝑖𝑗 ∇𝜇𝑗 ⎠ (2.13)
𝜕𝑡 𝑗

2.2 Stefan boundary condition


In addition, to all the global boundary conditions solidification problems
involve heat and mass transfer at the interface between solid and the liquid
because of phase change. The heat evolved due to the phase transformation
of solid to liquid per unit mass of the solid equals the latent heat. During,
solidification the interface between the solid and the liquid assumes a
temperature, given by the Gibbs-Thomson condition. For this condition to
hold for all times of solidification, we must have a heat balance between
the amount of heat generated due to the latent heat and the amount of
heat carrier away by diffusion into the solid and the liquid. This balance,
equationn writes down as,

(︂ )︂ (︂ )︂
𝜕𝑇 𝜕𝑇
𝐾𝑠 − 𝐾𝑙 = 𝐿𝑓 𝑉, (2.14)
𝜕𝑥 𝑠 𝜕𝑥 𝑙

where 𝐿𝑓 , is the specific heat of fusion and 𝑉 is the velocity of the interface.
The preceding equation is a relation between the thermal gradients on
either side of the interface and the interfacial velocity. A corresponding
relation for the case of mass transfer reads as (simplified for the case of
binary alloys),

(︂ )︂ (︂ )︂
𝜕𝜇 𝜕𝜇
𝑀𝑠 − 𝑀𝑙 = (𝑐𝑙 − 𝑐𝑠 ) 𝑉, (2.15)
𝜕𝑥 𝑠 𝜕𝑥 𝑙

where 𝑐𝑠 and 𝑐𝑙 represent the compositions in the solid and the liquid at
the interface.
18 2 Transport processes in solidification

2.3 One dimensional freezing: Quasi-static


approximation
Consider the situation of a solid in contact with the liquid which is
maintained at a temperature very close to the melting point 𝑇𝑚 . In the
quasi-static approximation, we have that the temperature field in the solid
has enough time to fully relax, upon a release of heat at the solidification
front, thereby leading to a linear temperature profile with the temperature
being 𝑇𝑚 at the solid-liquid interface and 𝑇𝑤 , which is the temperature of
the wall in contact with the surroundings. Also, the diffusional gradients
in the solid will considered small enough such that the temperature profile
in this study will be considered to be constant in the melt. Writing, the
heat flux at the wall, due to newton cooling, with a heat transfer coefficient
of ℎ writes as,

𝑞𝑤𝑎𝑙𝑙 = −ℎ (𝑇𝑤 − 𝑇∞ ) . (2.16)

where 𝑇∞ is the temperature of the surrounding environment. At the


quasi-static state, there is balance between the diffusional heat flux in
the solid and the heat lost due to the heat transfer at the interface. This
balance reads as,

−ℎ (𝑇𝑤 − 𝑇∞ ) = −𝐾𝑠 (𝑇𝑚 − 𝑇𝑤 ) /𝑙, (2.17)

where 𝑙 is the solidified length. Solving, for the wall temperature we get,

(︂ )︂
ℎ𝑙
𝑇𝑚 + 𝑇∞
𝐾𝑠
𝑇𝑤 = . (2.18)
ℎ𝑙
1+
𝐾𝑠

The Stefan boundary condition at the interface reads as,


2.3 One dimensional freezing: Quasi-static approximation 19

𝐾𝑠
(𝑇𝑚 − 𝑇𝑤 ) = 𝐿𝑓 𝑉. (2.19)
𝑙

Inserting the expression for 𝑇𝑤 in the preceding equation we derive,

ℎ (𝑇𝑚 − 𝑇∞ ) 𝑑𝑙
(︂ )︂ = 𝐿𝑓 . (2.20)
ℎ𝑙 𝑑𝑡
1+
𝐾𝑠

𝐿2
Multiplying, both sides of the preceding relation with , where 𝐿 is the
𝐿
length of the slab which is going to solidify, and multiply the numerator
𝐾𝑠
and denóminator of the RHS with the thermal diffusivity 𝛼 = , we
𝜌𝐶𝑝
derive,

(𝑇𝑚 − 𝑇∞ ) 𝑑𝑓𝑠
ℎ𝐿 (︂ )︂ = (𝐿𝑓 𝛼) . (2.21)
ℎ𝐿𝑓𝑠 𝑑 (𝛼𝑡/𝐿2 )
1+
𝐾𝑠

ℎ𝐿
where 𝑓𝑠 = 𝑙/𝐿. Here, we define certain non-dimensional numbers: =
𝐾𝑠
𝛼𝑡 𝐿𝑓
𝐵𝑖(Biot Number) and = 𝐹 𝑜(Fourier number). Also, note has
𝐿2 𝜌𝐶𝑝
units of temperature, such that a third non-dimensional number can be
𝑇𝑚 − 𝑇∞
written down as which scales with the differential temperature
𝐿𝑓
𝜌𝐶𝑝
𝑇𝑚 − 𝑇∞ . This is called the Stefan number. Writing down the equations
in terms of the non-dimensional numbers we derive,

𝑆𝑡 𝑑𝑓𝑠
= . (2.22)
1 𝑑𝐹 𝑜
+ 𝑓𝑠
𝐵𝑖
20 2 Transport processes in solidification

For a given length, the equation can be integrated for 𝑓𝑠 in terms of


the Fourier number and the following result can be derived by using the
boundary condition that at 𝐹 𝑜 = 0, the solidfied fraction is 0. The fraction
solid 𝑓𝑠 as a function of 𝐹 𝑜 writes as,

𝑓𝑠 𝑓2
𝑆𝑡.𝐹 𝑜 = + 𝑠. (2.23)
𝐵𝑖 2
The preceding quadratic equation can be solved for the unknown 𝑓𝑠 by
taking the positive root which writes as,

√︃
1 1
𝑓𝑠 = 2 + 2𝑆𝑡.𝐹 𝑜 − . (2.24)
𝐵𝑖 𝐵𝑖

2.4 Neumann’s solution: Semi-infinite freezing


Neumann’s solution to the moving boundary problem assumes, solidifica-
tion in one direction, with the solidified length increasing as a function of
time. The wall in contact with the first solid that forms is kept at a uniform
temperature of 𝑇𝑤 , while the temperature in the liquid decays from a
temperature of 𝑇𝑚 at the interface to a value 𝑇𝑜 far away. Neumann solves
this semi-infinite problem, by solving the diffusion equations in the bulk,
solid and the liquid and assuming the interface temperature remains at
𝑇𝑚 and applying the Stefan boundary condition at the interface. For this,
𝑥
we introduce a similarity variable 𝜉 = √ , and transform the diffusion
4𝛼𝑡
equations in each phase using this variable. The diffusion equation are of
the following form in the respective domains marked by the solid and the
liquid phases.

𝜕 2 𝑇𝑠 1 𝜕𝑇𝑠
= , 𝑥 < 𝑥𝑓 (2.25)
𝜕𝑥2 𝛼𝑠 𝜕𝑡
𝜕 2 𝑇𝑙 1 𝜕𝑇𝑙
= , 𝑥 > 𝑥𝑓 . (2.26)
𝜕𝑥2 𝛼𝑙 𝜕𝑡
2.4 Neumann’s solution: Semi-infinite freezing 21

where 𝑥𝑓 is the position of the solidification front. Performing the trans-


formation, the diffusion equations are transformed to,

𝜕 2 𝑇𝑠 𝜕𝑇𝑠
2
= −2𝜉 , 𝑥 < 𝑥𝑓 (2.27)
𝜕𝜉𝑠 𝜕𝜉𝑠
𝜕 2 𝑇𝑙 𝜕𝑇𝑙
2 = −2𝜉 , 𝑥 > 𝑥𝑓 , (2.28)
𝜕𝜉𝑙 𝜕𝜉𝑙

𝑥 𝑥
where 𝜉𝑠 = √ and 𝜉𝑙 = √ . Integrating both equations once, and
4𝛼𝑠 𝑡 4𝛼𝑙 𝑡
taking the exponential on both sides of the solution, we derive,

𝜕𝑇𝑠
= 𝐴 exp −𝜉𝑠2 , 𝑥 < 𝑥𝑓
(︀ )︀
(2.29)
𝜕𝜉
𝜕𝑇𝑙
= 𝐵 exp −𝜉𝑙2 , 𝑥 > 𝑥𝑓 .
(︀ )︀
(2.30)
𝜕𝜉

where 𝐴 and 𝐵 are constants of integration. Integrating once more, we


derive the temperature profiles for any time t, on the solid and the liquid
sides as

∫︁ 𝜉 √
𝜋
exp −𝜉 2 𝑑𝜉 = 𝐴
(︀ )︀
𝑇𝑠 (𝑥, 𝑡) − 𝑇𝑠 (0, 𝑡) = 𝐴 𝑒𝑟𝑓 (𝜉) (2.31)
0 2
∫︁ ∞ √
𝜋
exp −𝜉 2 𝑑𝜉 = −𝐵
(︀ )︀
𝑇𝑙 (𝑥, 𝑡) − 𝑇𝑙 (∞, 𝑡) = −𝐵 𝑒𝑟𝑓 𝑐 (𝜉) , (2.32)
𝜉 2

where we have used the definitions for the error function and its complement
to describe the variation of temperature in each phase. We note that
𝑇𝑠 (0, 𝑡) = 𝑇𝑤 and 𝑇𝑙 (∞, 𝑡) = 𝑇𝑜 . Using the information that the front
temperature is at 𝑇𝑚 at times, we have,
22 2 Transport processes in solidification


𝜋
𝑒𝑟𝑓 𝜉𝑓𝑠
(︀ )︀
𝑇𝑤 − 𝑇𝑚 = −𝐴 (2.33)
√2
𝜋
𝑒𝑟𝑓 𝑐 𝜉𝑓𝑙 .
(︀ )︀
𝑇𝑜 − 𝑇𝑚 = 𝐵 (2.34)
2

Dividing both sides of the expression for temperature on the solid side
with 𝑇𝑤 − 𝑇𝑚 and the liquid by 𝑇𝑜 − 𝑇𝑚 , we derive,

𝑇𝑠 (𝑥, 𝑡) − 𝑇𝑚 𝑒𝑟𝑓 (𝜉)


=1− (︁ )︁ (2.35)
𝑇𝑤 − 𝑇𝑚 𝑒𝑟𝑓 𝜉𝑓𝑠
𝑇𝑙 (𝑥, 𝑡) − 𝑇𝑚 𝑒𝑟𝑓 𝑐 (𝜉)
=1− (︁ )︁ . (2.36)
𝑇𝑜 − 𝑇𝑚 𝑒𝑟𝑓 𝑐 𝜉𝑓𝑙

𝑇𝑤 − 𝑇𝑚
Please note that from equations in Eqn.2.34, we can derive √ =
𝜋
−𝐴
2
(︁ )︁ 𝑇𝑜 − 𝑇𝑚 (︁ )︁
𝑒𝑟𝑓 𝜉𝑓𝑠 and √ = 𝑒𝑟𝑓 𝑐 𝜉𝑓𝑙 , where the L.H.S of both relations are
𝜋
𝐵
2
constants, thereby rendering the error functions with 𝜉𝑓𝑠,𝑙 as the arguments
constant in time. This can only occur for constant 𝜉 which implies for
both the solid and the liquid the following relations must hold,

𝑥𝑓
√ = 𝜆𝑠 (2.37)
4𝛼𝑠 𝑡
𝑥
√ 𝑓 = 𝜆𝑙 . (2.38)
4𝛼𝑙 𝑡

where 𝜆𝑠 and 𝜆𝑙 are constants. Therefore the change in the position of the
solidification of either front can be written down as,


𝑥𝑓 (𝑡) = 𝜆𝑖 4𝛼𝑖 𝑡 ∀𝑖 ∈ (𝑠, 𝑙). (2.39)
2.4 Neumann’s solution: Semi-infinite freezing 23

√︂
𝜆𝑠 𝛼𝑙
Also we have = , which can be derived by dividing the relations
𝜆𝑙 𝛼𝑠
for 𝜆𝑖 . To complete the description of the problem we need to specify the
value of 𝜆𝑖 , which are functions of the growth conditions. For this, we will
utilize the Stefan condition, which upon transformation of variables from
𝑥, 𝑡 → 𝜉 we derive,

𝜕𝑇 1 𝜕𝑇 1 𝑑𝑥𝑓
𝐾𝑠 √ − 𝐾𝑙 √ = 𝐿𝑓 . (2.40)
𝜕𝜉𝑠 4𝛼𝑠 𝑡 𝜕𝜉𝑙 4𝛼𝑙 𝑡 𝑑𝑡

Using the relations in Eqn.2.30, and the preceding relations relating 𝜆𝑠


and 𝜆𝑙 , we can derive the following non-linear equation for the constant
𝜆𝑠 ,

(︂ )︂
2 𝛼𝑙
(︀)︀ 𝐾𝑙 (𝑇𝑜 − 𝑇𝑚 ) exp −𝜆𝑠
𝐾𝑠 (𝑇𝑚 − 𝑇𝑤 ) exp −𝜆2𝑠
√︂
𝛼𝑠 𝛼𝑠
√ − √ = 𝐿𝑓 𝜆𝑠 ,
√ 𝛼𝑙 √
(︂ √︂ )︂
𝜋 𝜋 𝑡
𝑒𝑟𝑓 (𝜆𝑠 ) 4𝛼𝑠 𝑡 𝑒𝑟𝑓 𝑐 𝜆𝑠 4𝛼𝑙 𝑡
2 2 𝛼𝑠
(2.41)

where we have differentiated both sides of the equation in 2.39 to derive


𝑑𝑥𝑓 1
. Cancelling √ throughout, we derive a non-linear equation can be
𝑑𝑡 𝑡
solved for 𝜆𝑠 and consequently 𝜆𝑙 , which can in turn be used to derive the
solution of the temperature profiles in every phase.
3 Solute redistribution during
solidification
26 3 Solute redistribution during solidification

During slow solidification of a multi-component alloy, the interfacial com-


positions of the different components are given by the Gibbs-Thomson
condition. Upon change of the fraction solid and the assosciated motion
of the interface, mass balance requires rejection of solute into the melt,
which is transported in the liquid through diffusion and in some cases
accompanied by convection. The change in the liquid composition and the
assosciated change in the solid composition upon change in the fraction
of solid is an important information as it enables the derivation of the
properties after the final solidification. Additionally, a knowledge about
the redistribution of the solute during solidification will allow the design
of thermal heat treatment cycles to homogenize the concentrations in the
solid in-order to derive uniform properties. In the following sections we will
derive expressions for the change in the solid and the liquid compositions
as functions of the fraction solid.

3.1 Solidification at equilibrium


As a first approximation, we assume both the solid and the liquid have
enough time to diffuse any inhomogeneity in composition, such that we
always have a uniform composition in both the phases. With this assump-
tion mass balance, requires that the fraction solid and the compositions in
the solid and the liquid satisfy the relation,
𝑐𝑠 𝑓𝑠 + 𝑐𝑙 𝑓𝑙 = 𝑐𝑜 , (3.1)

where 𝑐𝑠 and 𝑐𝑙 are the compositions in the solid and the liquid, while
𝑐𝑜 is the initial composition in the melt. 𝑓𝑠 and 𝑓𝑙 denote the fractions
of the solid and the liquid phases, with 𝑓𝑠 + 𝑓𝑙 = 1. Assuming that the
partition coefficient of the phase diagram is 𝑘 = 𝑐𝑠 /𝑐𝑙 , which is assumed to
be relatively constant for the range of solidification conditions, we derive
the composition in the melt as,

𝑐𝑙 1
= (3.2)
𝑐𝑜 1 + 𝑓𝑠 (𝑘 − 1)
𝑐𝑠 𝑘
= . (3.3)
𝑐𝑜 1 + 𝑓𝑠 (𝑘 − 1)
3.2 Non-equilibrium solidification (Scheil): No diffusion in the melt 27

3.2 Non-equilibrium solidification (Scheil): No


diffusion in the melt
The more realistic case of alloy solidification is where the diffusion in
the solid is much smaller compared to that in the liquid. As a first
approximation in this case, we will assume that there is no diffusion in
the solid while the diffusivity in the liquid is fast enough such that in
the time-scale in which the interace moves, the inhomogeneities in the
composition in the melt are completely relaxed. In this situation, one
cannot write the mass-balance using global solid and liquid compositions as
in the equilibrium case because of the non-uniformity in the compositions
in the solid. Thus, only instantaneous mass balance relations using the
local compositions of the solid and liquid at the interface and the global
liquid composition(fast diffusion), for small infinitesimal motions of the
interface can be written down. Such a mass balance writes as,

(𝑐𝑙 − 𝑐𝑠 ) 𝑑𝑓𝑠 = 𝑑𝑐𝑙 (1 − 𝑓𝑠 ) . (3.4)

The above equation can be interpretted as the rejected solute due to the
advance of the front by 𝑑𝑓𝑠 , must equal the change of composition in
the remaining liquid to satisfy mass balance. Using the definition of the
partition coefficient we have the differential equation,

𝑑𝑓𝑠 𝑑𝑐𝑙
= . (3.5)
1 − 𝑓𝑠 𝑐𝑙 (1 − 𝑘)

Integrating both sides, we have,

(︂ )︂ 1
𝑐𝑙 (𝑘 − 1)
ln (1 − 𝑓𝑠 ) = ln (3.6)
𝑐𝑜
𝑐𝑙 𝑘−1
= (1 − 𝑓𝑠 ) (3.7)
𝑐𝑜
𝑐𝑠 𝑘−1
= 𝑘 (1 − 𝑓𝑠 ) (3.8)
𝑐𝑜
28 3 Solute redistribution during solidification

3.3 Solidification with small diffusion in the


melt (Brody-Flemmings model)
In the scheil equation the balance of mass at the interface is without the
influence of back-diffusion of the solute in the already solidified solid. In
order, to include this, one must subtract the necessary amount of solute
from the rejected solute on account of diffusion. In the Brody-Flemmings
model, the amount of solute back-diffusion is accounted for by calculating
the area under the curve of the solute-field integrated until the diffusion
length. The mass- balance relation with the approximate incorporation
of back-diffusion across a cross-sectional area of magnitude unity, writes
as,
1
(𝑐𝑙 − 𝑐𝑠 ) 𝐿𝑑𝑓𝑠 − 𝛿𝑠 𝑑𝑐𝑠 = 𝐿 (1 − 𝑓𝑠 ) 𝑑𝑐𝐿 , (3.9)
2

where 𝐿 is the length to be solidified, and 𝛿𝑠 is the diffusion length.


2𝐷𝑠
Substituting the diffusion length as 𝛿𝑠 = , where 𝐷𝑠 is the diffusivity
𝑉
in the solid and 𝑉 is the velocity of the solidification front also written as
𝑑𝑥
, we derive,
𝑑𝑡

(︂ )︂
𝑑𝑐𝑠
(𝑐𝑙 − 𝑐𝑠 ) 𝐿𝑑𝑓𝑠 − 𝐷𝑠 𝑑𝑡 = 𝐿 (1 − 𝑓𝑠 ) 𝑑𝑐𝐿 . (3.10)
𝑑𝑥

The variation of 𝑐𝑠 can be expressed as a function of the change in the


solid fraction 𝑓𝑠 and written as,

(︂ )︂
𝐷𝑠 𝑑𝑐𝑠
(𝑐𝑙 − 𝑐𝑠 ) 𝑑𝑓𝑠 − 2 𝑑𝑡 = (1 − 𝑓𝑠 ) 𝑑𝑐𝐿 . (3.11)
𝐿 𝑑𝑓𝑠

Assuming that we are in a regime where the growth of the solid is diffusion

limited, that we can borrow from Neumann’s growth law that, 𝑓𝑠 = 𝜆 4𝛼𝑡,
𝑑𝑓𝑠 𝛼𝜆2
which can be used to determine = 2 2 . Substituting, re-arranging
𝑑𝑡 𝐿 𝑓𝑠
3.4 Error in the Brody-Flemmings model 29

and forming the differential equation between the change in the liquid
composition 𝑑𝑐𝑙 in response to change in the fraction of solid 𝑓𝑠 as,
𝑑𝑐𝑙 𝑑𝑓
= (︂ )︂ 𝑠 (3.12)
𝑐𝑙 (1 − 𝑘) 𝐷𝑠 𝑓𝑠 𝑘
+ (1 − 𝑓𝑠 )
2𝛼𝜆2
𝑑𝑐𝑙 (1 − 𝑘) 𝑑𝑓𝑠
= (︂ )︂ (3.13)
𝑐𝑙 𝐷𝑠 𝑓𝑠 𝑘
+ (1 − 𝑓𝑠 )
2𝛼𝜆2

Integrating, the preceding differential equation with limits of integration


𝑐0 → 𝑐𝑙 for the LHS and correspondingly 0 → 𝑓𝑠 for the solid, we have,

(𝑘 − 1)
(︂ )︂
(︂ (︂ )︂)︂ 1 − 𝐷𝑠 𝑘
𝐷𝑠 𝑘 2𝛼𝜆2 .
𝑐𝑙 = 𝑐0 1 − 𝑓𝑠 1 − (3.14)
2𝛼𝜆2

The preceding equation predicts the correct limit as for 𝐷𝑠 → 0, we


recover the Scheil prediction. However for 𝐷𝑠 tending to larger values, the
prediction does not converge to the equilibrium value, which happens for
𝐷𝑠
the special case of → 1. This is an obvious anomaly in the model
2𝛼𝜆2
which arises because of the following assumptions in the model.

3.4 Error in the Brody-Flemmings model


For systems where the diffusion lengths are going to be small as compared to
the system sizes the Brody-Flemmings gives a good estimate. However, as
the diffusion length increases the error due to Brody-Flemmings increases.
This was noticed by Flemmings himself, in that he confirmed that the
system of equations written down in the Brody-Flemmings model seized
to me mass conserving on certain occassions. The error relates to the over
estimation of the back diffusion. A modified analysis is present from Clyne
and Kurz. Essentially, the overestimation occurs as part of the diffusion
30 3 Solute redistribution during solidification

profile which is integrated in the Brody-Flemmings model, lies outside


the domain for cases where the diffusion length is large compared to the
system size. The amount of error, performed with this overestimation
can be estimated by plotting the difference in the composition profiles
which are obtained before and after the advancement of the interface by
an infinitesimal amount. As an approximation, this is assumed to decay
in an exponential manner starting from the interface and going backwards
into the solid; 𝑑𝑐𝑠 exp (−2𝑦/𝛿𝑠 ). Integrating this we get the following
expression,

∫︁ ∞
𝑑𝑐𝑠 𝛿𝑠
𝐴𝑇 = 𝑑𝑐𝑠 exp (−2𝑦/𝛿𝑠 ) 𝑑𝑥 = . (3.15)
0 2

However, for systems where 𝛿𝑠 ≈ 𝐿, part of this difference lies outside the
domain. Hence, the error performed is,

∫︁ ∞ (︂ )︂
𝑑𝑐𝑠 𝛿𝑠 𝑦
𝐴𝐸 = 𝑑𝑐𝑠 exp (−2𝑦/𝛿𝑠 ) 𝑑𝑦 = exp −2 . (3.16)
𝑦𝑖 2 𝛿𝑠

where 𝑦𝑖 is the position of the interface.


∑︀
A parameter constructed as the ratio of the areas and integrated over
time would help in encapsulating the total error perfomed. The parameter
can be explicitly written as,

∫︁ 𝑡𝑓 (︂ )︂
∑︁ 1 𝐴𝐸
= 𝑑𝑡. (3.17)
𝑡𝑓 0 𝐴𝑇

In the above expressions, the ratio of the areas change as a function of


time, due to the change in the position of the interface.
∑︀ For cases, where
the velocity is constant, we can derive the parameter as,

∑︁ 𝐷𝑠 (︀
1 − exp −4𝛼𝜆2 /𝐷𝑠 ,
(︀ )︀)︀
= (3.18)
4𝛼𝜆2
3.4 Error in the Brody-Flemmings model 31

while for a parabolic decay of the velocity profile, we can write down the
parameter as,

∑︁
= exp −2𝛼𝜆2 /𝐷𝑠 .
(︀ )︀
(3.19)

Plotting of the two functions shows that the curves converge for significant
𝐷𝑠
values 𝛼 to a value of 1.0 and are considerably different for values of
4𝛼𝜆2
greater than 0.1. Thus the estimates from Brody-Flemmings give results
𝐷𝑠
which are reasonable, only for values of which are fairly small, where
4𝛼𝜆2
the deviation from Scheil is slight. To improve the applicability of the
𝐷𝑠
model for values of which are substantial Clyne and Kurz, propose
4𝛼𝜆2
the construction of the parameter which approaches the correct values of
𝐷𝑠
the parameter , for both limits (Scheil and equilibrium). To achieve
4𝛼𝜆2
𝐷𝑠
this, we note that for = 0.5, we retrieve the the equilibrium limit and
4𝛼𝜆2
𝐷𝑠
for = 0, we retrieve the Scheil limit. So the aim is to construct such
4𝛼𝜆2
𝐷𝑠
a parameter 𝜉 which assumes the value 0 for tending to zero and
∑︀ 4𝛼𝜆2
equal to 0.5 for tending to 1. This function can be created exemplarily
as,

𝐷𝑠 (︀ )︀)︀ 1
1 − exp −4𝛼𝜆2 /𝐷𝑠 − exp −2𝛼𝜆2 /𝐷𝑠 .
(︀ (︀ )︀
𝜉= 2
(3.20)
4𝛼𝜆 2

𝐷𝑠
Using 𝜉 in place in Brody-Flemmings equation gives the right limits.
4𝛼𝜆2
The modified Brody-Flemmings equation writes as,

(𝑘 − 1)
𝑐𝑙 = 𝑐0 (1 − 𝑓𝑠 (1 − 2𝜉𝑘)) − 2𝜉𝑘) .
(1 (3.21)
32 3 Solute redistribution during solidification

3.5 Zone refining

3.5.1 Purification as result of cyclic unidirectional


solidification

In all the segregation models we have see that as a result of solidification,


the solid becomes pure in terms of the solute content for 𝑘 < 1, with
respect to the alloy composition one started with. Hence, such a technique
of directionally solidifying a bar leads to segregation of impurities from
one section to another leading to a purer solid. This is a common route for
producing purer materials. To estimate the fraction of the solidified metal
which is purified with respect to the starting composition of the alloy, we
write down the composition of the solid as a function of the fraction solid
from the Scheil equation which writes as,

𝑘−1
𝑐𝑠 = 𝑘𝑐𝑜 (1 − 𝑓𝑠 ) , (3.22)

which is a function which increases with 𝑓𝑠 for 𝑘 < 1. Therefore, there


exists a position in the bar for which the composition will be above
𝑐𝑜 , i.e the initial alloy composition. The 𝑓𝑠* corresponding to this alloy
composition, can be derived by substituting 𝑐𝑠 = 𝑐𝑜 and solving for 𝑓𝑠
which derives as,

(︂ )︂ 1
1 𝑘−1
𝑓𝑠* = 1 − . (3.23)
𝑘

Clearly, if this process is repeated the composition of the purer solid


will decrease further at the expense of a lesser amount of pure solid, i.e
composition below the starting alloy composition.
To get an estimate of the average composition in the solid after one such
solidification cycle, we perform a spatial average of the compositional field
in the solid which writes as,
3.5 Zone refining 33

𝑓𝑠*
⟨𝑐1𝑠 ⟩
∫︁
𝑘 𝑘−1
= * (1 − 𝑓𝑠 ) 𝑑𝑓𝑠 (3.24)
𝑐𝑜 𝑓𝑠 0
𝑘
1−𝑘 (1 − 𝑘)
= . (3.25)
1
1−𝑘 1 − 𝑘

Repeating, the cycle on the homogenized sample we arrive, we can derive,

𝑓𝑠*
⟨𝑐2𝑠 ⟩
∫︁
𝑘 𝑘−1
1
= * (1 − 𝑓𝑠 ) 𝑑𝑓𝑠 (3.26)
⟨𝑐𝑠 ⟩ 𝑓𝑠 0
𝑘
1−𝑘 (1 − 𝑘)
= . (3.27)
1
1 − 𝑘1 − 𝑘

At the end of 𝑛 cycles the composition in the solid writes as,

⎞𝑛
𝑘

⎜ 1 − 𝑘 (1 − 𝑘) ⎟
⟨𝑐𝑛𝑠 ⟩ = 𝑐𝑜 ⎜ ⎟ . (3.28)
⎜ ⎟
⎝ 1 ⎠
1−𝑘 1 − 𝑘

The idea for zone refining extends from the above result. However, the
repeated cycles are not performed by removing material and homogenizing
after each step, rather it is performed by repeatedly performing passes
where a small area of the solid is melted which is also called the melt-zone.
The melt-zone moves through the entire solid, with each pass impurities
get segregated in the melt and finally get segregated towards one direction
of the bar. In order to analyze this process an important question one
attempts to address is to find the change in the composition of the solid
34 3 Solute redistribution during solidification

along the bar during a single pass. The transport phenomena involved in
the change of concentration in the melt zone are the following,

∙ Part of the impurities from the liquid in the melt-zone get refrozen
back in the solid
∙ Impurities from the solid enter the melt zone from the new solid that
melts.
∙ A balance of the above processes leads to an increase in the compo-
sition of the liquid constituting the melt-zone of length (𝑧).

In the form of equations, the above relations write as,

(︂ )︂
𝑑𝑥 𝑑𝑥 𝑑𝑐𝑙
𝑐𝑜 − 𝑘𝑐𝑙 =𝑧 . (3.29)
𝑑𝑡 𝑑𝑡 𝑑𝑡

Re-arranging to form a differential equation between the 𝑐𝑙 and 𝑥, we


have,

(︂ )︂
𝑐𝑙
𝑑 (︂ )︂
𝑑𝑥 𝑐𝑜 𝑐𝑙
=𝑧 / 1−𝑘 . (3.30)
𝑑𝑡 𝑑𝑡 𝑐𝑜

Integrating, we derive the equation for the composition along the pass-
direction as,

[︂ (︂ )︂]︂
𝑐𝑙 1 𝑘𝑥
= 1 − (1 − 𝑘) exp − . (3.31)
𝑐𝑜 𝑘 𝑧

The equation is valid until 𝑥 + 𝑧 = 𝐿, that is the region where the


farthest point in the bar matches the end-point of the melt-zone. Here,
the conditions resemble the unidirectional solidification that we have dealt
with in all the segregation models. Assuming, the Scheil segregation model,
3.6 Multi-pass zone refining 35

we derive that the distribution of the impurities in the final melt-zone


beyond 𝑥 = 𝐿 − 𝑧,

𝑘−1
𝑐𝑠 = 𝑐𝐿−𝑧 𝑘 (1 − 𝑓𝑠 ) , (3.32)

𝑥 − (𝐿 − 𝑧)
where the fraction of solid here is represented as, 𝑓𝑠 = , which
𝑧
𝐿−𝑥
can be concisely be represented as 1−𝑓𝑠 = . Putting things together,
𝑧
the composition in this final solidified melt zone can be expressed as,

[︂ (︂ )︂]︂ (︂ )︂𝑘−1
𝑘 (𝐿 − 𝑧) 𝐿−𝑥
𝑐𝑠 = 𝑐𝑜 1 − (1 − 𝑘) exp − . (3.33)
𝑧 𝑧

3.6 Multi-pass zone refining


One can perform multiple passes for the process just described. However,
there is an ’ultimate impurity distribution’ that can be achieved as with
repeated beyond a point, impurities start back segregating into the frozen
solid thus not allowing further purification. An estimate of this composition
can be achieved by setting the average melt-zone composition equal to the
composition at the position 𝑥. This average melt composition writes as,

∫︁ 𝑥+𝑧
1
⟨𝑐∞
𝑙 (𝑥)⟩ = 𝑐∞
𝑠 (𝑥) 𝑑𝑥, (3.34)
𝑧 𝑥

and the ultimate impurity distribution would be described by the solution


to the equation,

∫︁ 𝑥+𝑧
𝑘
⟨𝑐∞
𝑠 (𝑥)⟩ = 𝑐∞
𝑠 (𝑥) 𝑑𝑥. (3.35)
𝑧 𝑥
36 3 Solute redistribution during solidification

A trial function which satisfies the above equation would be of the form
𝑐∞
𝑠 (𝑥) = 𝛼 exp (𝜂𝑥), where one needs to determing the constants 𝛼 and 𝑘.
Substituting, the form of the solution back into the preceding equation
allows us to derive a relation between 𝜂 and k which writes as,

𝜂𝑧
𝑘= . (3.36)
exp (𝜂𝑧) − 1

The preceding equation can be numerically solved for 𝜂 given 𝑧 and 𝑘.


We use conservation of mass to determine the unknown 𝛼. The mass
conservation equation writes as,

∫︁ 𝐿
1
𝑐𝑜 = 𝛼 exp (𝜂𝑥) 𝑑𝑥. (3.37)
𝐿 0

Evaluating the integral and solving for 𝛼 gives,

(︂ )︂
𝜂𝐿
𝛼 = 𝑐𝑜 . (3.38)
exp (𝜂𝐿) − 1

With this the ultimate impurity distribution writes as,

𝑐∞
𝑠 (𝑥) 𝜂𝐿
= exp (𝜂𝑥) . (3.39)
𝑐𝑜 exp (𝜂𝐿) − 1

The most pure material appears at the very start of the pass i.e. 𝑥 = 0.
After the first pass 𝑐𝑠 (0) /𝑐𝑜 = 𝑘, after second pass it would be 𝑐𝑠 (0) /𝑐𝑜 =
𝑘 2 and therefore after 𝑛 passes would be 𝑘 𝑛 . To determine the number
of passes 𝑛* in order to attain the ultimate impurity distribution one can
solve for the equation,
3.6 Multi-pass zone refining 37

𝜂𝐿
𝑘 𝑛* = (3.40)
exp (𝜂𝐿) − 1
(︂ )︂
𝜂𝐿
ln
exp (𝜂𝐿) − 1
𝑛* = . (3.41)
ln 𝑘
4 Mullins-Sekerka Instability in a
nut-shell
40 4 Mullins-Sekerka Instability in a nut-shell

It is well known that a planar solidification front can break up into cells
and dendrites beyond certain critical velocity of the interface, as a result
of the amplification of fluctuations at the interface arising out of thermal
or composition distrubances.
In this chapter, we get a flavor of doing a stability analysis of the growth of
a planar interface towards the formation of cells and dendrites. The basis
of the analysis is the derivation of the time rate of change a perturbation to
the interface, which is described as a fourier mode. The interface remains
stable if the growth rates of all fourier modes are negative and unstable if
any given mode amplifies. The analysis will allow us to derive a stability
criterion for the various solidification parameters.
The analysis presented here is a simplification of the one present in the
original paper by Mullins and Sekerka(1964), wherein we work out the
stability problem in the limit where the gradients in the temperature are
too small and hence the stability of the interface is purely determined
through the feedback of the concentration field to given perturbations.
Therefore, the only governing equation for our case is,

(︂ )︂ (︂ )︂
2 𝑉 𝜕𝑐
∇ 𝑐+ = 0, (4.1)
𝐷 𝜕𝑧

where 𝑥 is the direction parallel to the growth front, and 𝑧 is the growth
direction. 𝑉 is the steady-state velocity of the interface and 𝐷 is the
diffusivity of the liquid. We are going to assume a one sided model,
wherein, the diffusivity in the solid is zero.

4.1 Solution for steady state planar interface


growth
For a planar interface growth, the governing equation simplifies to,

𝜕2𝑐 𝑉 𝜕𝑐
+ = 0. (4.2)
𝜕𝑧 2 𝐷 𝜕𝑧
4.2 Boundary conditions at the interface 41

Integrating once and using the information, that we do indeed know the
gradient of the concentration at the interface 𝐺𝑐 , we get,

(︂ )︂
𝜕𝑐 𝑉𝑧
ln =− + ln 𝐺𝑐 , (4.3)
𝜕𝑧 𝐷

the variation of the gradient of the composition can be derived as,

(︂ )︂
𝜕𝑐 −𝑉 𝑧
= 𝐺𝑐 exp . (4.4)
𝜕𝑧 𝐷

Integrating again, and using the information that the concentration at the
interface is 𝑐𝑜 , we can write, the variation of the composition as,

[︂ (︂ )︂]︂
𝐺𝑐 𝐷 −𝑉 𝑧
𝑐 = 𝑐𝑜 + 1 − exp . (4.5)
𝑉 𝐷

4.2 Boundary conditions at the interface

4.2.1 Gibbs-Thomson condition

When an interface is curved, thermodynamic equilibrium at the interface


requires that the concentration at the interface be perturbed from its
equilibrium value at a planar interface. For, the present situation let us
assume a linear phase diagram where the liquidus slope can be written
down as 𝑚. Then the temperature at the interface can be written down
as,

𝑇𝜑 = 𝑇 = 𝑇𝑁 − 𝑚𝑐𝜑 − Γ𝜅. (4.6)


42 4 Mullins-Sekerka Instability in a nut-shell

where 𝑇 is the isothermal temperature of solidification. For a perturbed


interface, where the perturbation is of the form 𝑧 = 𝛿 sin 𝜔𝑥,

𝑇𝜑 = 𝑇 = 𝑇𝑁 − 𝑚𝑐𝜑 − 𝛿Γ𝜔 2 sin 𝜔𝑥. (4.7)

Therefore 𝑐𝜑 can be written down as,

𝑇𝑁 − 𝑇 − 𝛿Γ𝜔 2 sin 𝜔𝑥
𝑐𝜑 = (4.8)
𝑚

4.2.2 Stefan condition (Moving boundary problem)

For a moving interface, we have mass balance at the interface to be satisfied


which gives us,

𝜕𝑐
(𝑐𝜑 − 𝑘𝑐𝜑 ) 𝑣 = −𝐷 𝑜𝑟 (4.9)
𝜕𝑧
𝐷 𝜕𝑐
𝑣 (𝑥) = . (4.10)
𝑐𝜑 (𝑘 − 1) 𝜕𝑧

4.3 Perturbation of the boundary-layer solution


The scheme of perturbation analysis is usually performed around a solu-
tion, i.e we perturb the given solution through for example a sinusoidal
perturbation, and derive from the set of governing equations and boundary
conditions, as to what is the rate of growth of the perturbation.
In the stability analysis performed by Mullins, we perturb the interface
through a sinusoidal wave of the form,

𝑧 = 𝜑 (𝑥, 𝑡) = 𝛿 (𝑡) sin (𝜔𝑥) , (4.11)


4.3 Perturbation of the boundary-layer solution 43

where 𝛿 is the height of the perturbation, and 𝑧 is measured with respect


to the steady state planar interface as the reference. 𝜔 is the frequency.
The assumption in this analysis, is for such small perturbations to the
interface will result in corresponding linear response to the composition
field such that the composition at the interface can be written down as

𝑐𝜑 = 𝑐𝑜 + 𝑏𝑧 = 𝑐𝑜 + 𝑏𝛿 sin (𝜔𝑥) , (4.12)

where 𝑏 is a parameter, which needs to be identified such that the boundary


conditions are satisfied.
We propose that the solution to ’𝑐’ can be written as a linear super-position
of the steady state solution of the planar front and a sinusoidal wave of
amplitude 𝐶1 which decays in the growth direction, such that beyond
several times the wavelength of perturbation in the liquid, we retrieve the
same solution as for a planar front. Therefore, we write the total solution
as,

𝑐 = 𝑐𝑝𝑙𝑎𝑛𝑎𝑟 + 𝐶1 sin(𝜔𝑥)𝑒𝑥𝑝(−𝑘𝜔 𝑧), (4.13)

where 𝑘𝜔 is the decay rate of the frequency 𝜔. Since, for small amplitude
perturbations, the planar solution 𝑐𝑝𝑙𝑎𝑛𝑎𝑟 confirms to the governing equa-
tion, the peturbed part of the solution, must also satisfy the governing
equation, such that the linear sum of the two parts is indeed a solution to
the concentration field at a perturbed interface. To have this we introduce
the perturbed part of the solution into the governing equation, which
gives,

(︂ )︂
𝑉
𝐶1 𝑘𝜔2 2
− 𝜔 − 𝑘𝜔 = 0, (4.14)
𝐷

which is quadratic equation in 𝑘𝜔 and can be solved for 𝑘𝜔 . The solution


to the above equation writes,
44 4 Mullins-Sekerka Instability in a nut-shell

√︃
𝑉 𝑉 2
𝑘𝜔 = + + 𝜔2 . (4.15)
2𝐷 2𝐷

In order to retrieve back a form 𝑐𝜑 as elaborated in Eqn. 4.12 for a


perturbed interface, we write the the expression for the composition as
given in Eqn.4.13, at 𝑧 = 𝜑 = 𝛿 sin 𝜔𝑥, where 𝛿 is small and determine,
the value of 𝐶1 .

𝑐 = 𝑐𝑝𝑙𝑎𝑛𝑎𝑟 + 𝐶1 sin(𝜔𝑥) exp(−𝑘𝜔 𝛿 sin 𝜔𝑥) (4.16)


(︂ (︂ )︂)︂
𝐺𝑐 𝐷 𝑉𝑧
𝑐 ≈ 𝑐𝑜 + 1− 1− + 𝐶1 sin (𝜔𝑥) (1 − 𝑘𝜔 𝛿 sin 𝜔𝑥) . (4.17)
𝑉 𝐷

At first order in 𝛿 , we have

𝑐 ≈ 𝑐𝑜 + 𝐺𝑐 𝛿 sin 𝜔𝑥 + 𝐶1 sin 𝜔𝑥. (4.18)

Comparison with equation 𝑐𝜑 , we have that 𝐶1 = 𝛿 (𝑏 − 𝐺𝑐 ). Therefore,


the concentration ahead of the perturbed interface can be written down
as,

𝑐 = 𝑐𝑝𝑙𝑎𝑛𝑎𝑟 + 𝛿 (𝑏 − 𝐺𝑐 ) sin(𝜔𝑥) exp(−𝑘𝜔 𝛿 sin 𝜔𝑥). (4.19)

4.4 Pertubation Analysis

To determine the form of the solution 𝑐 we need to determine the parameter


𝑏. This can de done by utilizing the Gibbs-Thomson condition as given in
Eqn. 4.8 and the form for 𝑐𝜑 as given in Eqn. 4.12.
4.4 Pertubation Analysis 45

(𝑇𝑁 − 𝑇 ) + 𝛿Γ𝜔 2 sin 𝜔𝑥


𝑐𝜑 = 𝑐𝑜 + 𝑏𝛿 sin 𝜔𝑥 = . (4.20)
𝑚

Equating the fourier terms we have,

𝑇𝑁 − 𝑇
= 𝑐𝑜 (4.21)
𝑚
Γ𝜔 2
𝑏=− . (4.22)
𝑚

Next we utilise the Stefan boundary condition in Eqn.4.10 and noting that
𝑣 (𝑥) = 𝑉 + 𝛿˙ sin 𝜔𝑥, where 𝑉 is the velocity of the steady state planar
˙ For this, we require
front, to determine the rate of change of amplitude 𝛿.
the gradients in the compositions, computed in the growth direction at
the perturbed interface. Here again, we limit ourselves to terms which are
first order in 𝛿. The expression writes as,

(︂ )︂ (︂ )︂
𝜕𝑐 𝑉 𝛿 sin 𝜔𝑥
= 𝐺𝑐 1 − − 𝑘𝜔 𝛿 (𝑏 − 𝐺𝑐 ) sin 𝜔𝑥 (4.23)
𝜕𝑧 𝜑 𝐷
(︂ [︂ ]︂)︂
𝑉
= 𝐺𝑐 − 𝑘𝜔 𝛿 sin 𝜔𝑥 𝑏 − 𝐺𝑐 1 − . (4.24)
𝑘𝜔 𝐷

Inserting the preceding expression in the Stefan condition Eqn.4.10 and


using,

(︂ )︂
1 1 𝑏
= 1 − 𝛿 sin 𝜔𝑥 , (4.25)
𝑐𝜑 𝑐𝑜 𝑐𝑜

𝐷𝐺𝑐
𝑉 = and (4.26)
𝑐𝑜 (𝑘 − 1)
𝛿˙
[︂ ]︂ (︂ )︂
𝐷𝑏 𝐺𝑐 𝐺𝑐 𝐷𝑘𝜔 𝑉
=− + 𝑘𝜔 + 1− , (4.27)
𝛿 𝑐𝑜 (𝑘 − 1) 𝑐𝑜 𝑐𝑜 (𝑘 − 1) 𝑘𝑤 𝐷
46 4 Mullins-Sekerka Instability in a nut-shell

where we have compared the fourier components, and limited ourselves only
𝐷𝐺𝑐 𝐺𝑐 𝑉 (𝑘 − 1)
to first order in 𝛿. Using 𝑉 = and therefore, = ,
𝑐𝑜 (𝑘 − 1) 𝑐𝑜 𝐷
we can write,

𝛿˙
[︂ ]︂ (︂ )︂
𝐷𝑏 𝑉 𝑉
= 𝑘𝜔 − (1 − 𝑘) + 𝑉 𝑘𝜔 − (4.28)
𝛿 𝑐𝑜 (1 − 𝑘) 𝐷 𝐷

Using the equation for 𝑘𝜔 as stated in Eqn. 4.15, it is easy to see that
𝑘𝜔 > 𝑉 /𝐷 > 𝑉 /𝐷(1 − 𝑘). Therefore, factorizing the term we derive,

𝛿˙
[︂ ]︂ (︂ )︂
𝑉 𝐷𝑏
= 𝑘𝜔 − (1 − 𝑘) + 𝑉 𝑔 (𝜔) , (4.29)
𝛿 𝐷 𝑐𝑜 (1 − 𝑘)

𝑉
𝑘𝜔 −
where 𝑔 (𝜔) = 𝐷 . Expanding, 𝑘𝜔 using Eqn. 4.15 we can
𝑉
𝑘𝜔 − (1 − 𝑘)
𝐷
analyze the behavior of 𝑔 (𝜔).

2𝑘
𝑔 (𝜔) = 1 − √︃ (︂ )︂2 . (4.30)
2𝐷𝜔
1+ + 2𝑘 − 1
𝑉

For perturbations, whose wavelength is much smaller as compared to the


diffusion length 2𝐷/𝑉 , the following approximation holds,
4.5 Directional solidification 47

2𝑘
𝑔 (𝜔) = 1 − (4.31)
2𝜔 2 𝐷2
2𝑘 +
𝑉2
1
𝑔 (𝜔) = 1 − (4.32)
𝜔2 𝐷2
1+ 2
𝑉 𝑘
𝜔2 𝐷2
𝑔 (𝜔) = (4.33)
𝑉 2𝑘

Substituting in the expression for the rate of change of amplitude,

𝛿˙ 𝐷Γ𝜔 2 𝐷2 𝜔2
[︂ ]︂ (︂ )︂
𝑉
= 𝑘𝜔 − (1 − 𝑘) − + (4.34)
𝛿 𝐷 𝑚𝑐𝑜 (1 − 𝑘) 𝑉𝑘

Therefore, the condition for stability 𝛿˙ < 0 derives as,

Γ𝑉 𝑘
> 1. (4.35)
𝐷𝑚𝑐𝑜 (1 − 𝑘)

𝐷 𝑉
One can use the relation = , to derive a equivalent
𝑚𝑐𝑜 (𝑘 − 1) −𝑚𝐺𝑐
relation for stability as,

Γ𝑉 2 𝑘
>1 (4.36)
(−𝑚)𝐷2 𝐺𝑐

4.5 Directional solidification


In the case of directional solidification, a growth velocity is imposed at
the interface, through either the pulling of a sample with respect to a
fixed temperature gradient or the movement of the interface at a given
48 4 Mullins-Sekerka Instability in a nut-shell

velocity, keeping the sample fixed. As the interface velocity matches, that
of the imposed temperature gradient which is positive going from the
solid to the liquid, the mean relative position of the interface with respect
to the furnace does not change. This setup is normally referred to as a
Bridgman furnace. This sort of setup, is normally used to grow crystals in
a given direction, (directional solidification). In the following section, we
will perform a perturbation analysis as before.
It is important to note, that the only major change in the analysis will be
the modification of the Gibbs-Thomson equation which can be derived for
the case of the directional solidification, where the temperature field at any
given point with an ordinate 𝑧 ′ due to a furnace moving at a given velocity
𝑉 can be written as 𝑇 = 𝑇𝑏 + 𝐺 (𝑧 ′ − 𝑉 𝑡), 𝐺 being the magnitude of the
temperature gradient, 𝑉 is the velocity, 𝑡 the time and 𝑇𝑏 is a reference
base temperature. At the interface, 𝑧 ′ = 𝑧 + 𝑍, where 𝑍 is the co-ordinate
of the flat interface, and 𝑧 = 𝛿 sin 𝜔𝑥 is the small perturbation, the growth
of which we want to analyze. The modified Gibbs-Thomson equaton writes
as,

(𝑇𝑏 + 𝐺 ((𝑧 + 𝑍) − 𝑉 𝑡)) = 𝑇𝑁 − 𝑚𝑐𝜑 − Γ𝜅. (4.37)

Re-arranging the expression for 𝑐𝜑 and using the relation, 𝑐𝜑 = 𝑐𝑜 + 𝑏𝜑,


we can work out a modified expression for 𝑏 as,

(︀ )︀
𝐺 + Γ𝜔 2
𝑏=− . (4.38)
𝑚

Replacing this relation in the equation for the rate of change of amplitude
in Eqn. 4.29 we derive,

𝛿˙ 𝐷Γ𝜔 2
[︂ ]︂ (︂ )︂
𝑉 𝐷𝐺
= 𝑘𝜔 − (1 − 𝑘) − − + 𝑉 𝑔 (𝜔) .
𝛿 𝐷 𝑚𝑐𝑜 (1 − 𝑘) 𝑚𝑐𝑜 (1 − 𝑘)
(4.39)
4.6 Discussion 49

We can see in the preceding expression, that there is now an additional


term which is frequency independent. One can plot, the rate of change
of perturbation as a function of 𝜔, to determine the critical wavelength
which will amplify.
For perturbations, whose wavelength 𝜆 is much smaller than the diffusion
length 2𝐷/𝑉 . In such cases, we can show that 𝑔 (𝜔) ≈ 1. Additionally,
for perturbations which are large such that the capillarity terms do not
dominate, a criterion of stability can be derived from the preceding equation
as,

𝑣𝑐𝑜 (1 − 𝑘) 1
< . (4.40)
−𝐷𝐺 𝑚

4.6 Discussion
∙ Considering Eqn.4.29 it is clear that in the absense of capillarity,
the first term would vanish and perturbations of all wavelengths
would amplify, i.e a stability criterion exists only in the presence of
capillarity.
∙ In the classical paper by Mullins-Sekerka, the authors work out the
stability criterion for the case of coupled thermal and solutal diffusion.
In order to pull an interface with a given speed V, the interface is
held between a hot and a cold furnace, i.e. a positive temperature
gradient exists, going from the solid and the liquid. Following, the
analysis one can write down a stability criterion as,

(︂ )︂
𝜅𝑠 − 𝜅𝑙 𝐿𝑉
−𝐺 + 𝑚𝐺𝑐 + 𝐺− < |𝑓 (𝜔) |. (4.41)
𝜅𝑠 + 𝜅𝐿 𝜅𝑠 + 𝜅𝐿

𝐷2 𝑚𝐺𝑐
(︂ )︂
2
where 𝑓 (𝜔) = −𝑚𝐺𝑐 + 𝜔 − Γ . This term is gener-
𝑉 2𝑘
ally small, however one must verify the magnitude to ascertain its
influence on the stability criterion.
50 4 Mullins-Sekerka Instability in a nut-shell

One can directly see that there are other stabilizing factors in the
presence of positive thermal gradients (𝐺). Additionally, the release
−𝐿𝑉
of latent heat at the interface stabilizes the interface. This
𝜅𝑠 + 𝜅𝐿
can be physically, seen as an effect of rise of local temperature
and thereby hindering the further growth of the perturbation. The
last term depicts a non-trivial effect depending on the difference
of conductivities 𝜅𝑠 − 𝜅𝐿 . When 𝜅𝑠 > 𝜅𝐿 , the perturbation is
destabilized, and vice-versa, when 𝜅𝑠 < 𝜅𝐿 . This effect can be
qualitatively understood as follows: 𝜅𝐺 determines the amount of
heat flux. −𝜅𝑠 𝐺 determines the amount of heat flux taken away
from the interface into the solid and −𝜅𝑙 𝐺 is the amount of heat flux
from the liquid to the interface. If the former current is stronger,
increase of gradients due to a perturbation in the solid would cause
cooling of the interface through greater heat dissipation in the solid
than would be the increase in heat current from the liquid to the
solid. This would help in making the interface more unstable. The
converse is true, for the case when 𝜅𝑠 < 𝜅𝐿 . One must note, that
the former is normally true for real systems.
∙ One can verify that in the absence of thermal gradients, the above
expression gives back the stability criterion previously derived.
5 Physics of Dendritic growth
52 5 Physics of Dendritic growth

Dendritic morphologies are ubiquitous in material science and are good


examples of fractal like structures. Examples of dendritic structures are
simple snow-flakes that are abundant in nature, while also occurring in
industrial metallic and organic alloys. A key characteristic of dendritic
growth is the phenomena of self-similarity between the patterns observed
at various scales, which are characteristics common to fractals. The
morphology can be characterized by growth of primary arms along well
defined crystallographic directions with each primary arm also giving rise
to self-similar secondary arms and tertiary arms.

In this chapter, we are going to survey the development of theories for


describing dendritic growth. Physically, one would like to predict the
morphological features describing the dendrite, which are the dendrite
tip radius, inter-dendritic and secondary arm spacings and how they
change as a function of the processing conditions, alloy composition and
properties. The requirement for understanding the scale and variation of
these parameters lies in the fact that the macroscopic parameters such as
tensile strength, and resistance to fracture are often determined by the
scale of the microstructure.

Dendritic microstructures result as a de-stabilization of a planar or spherical


growth front. A critical component giving rise to dendritic growth, where
the primary growth directions predominantly occur along well defined
growth directions, is surface energy anisotropy. As we shall see in the
course of our analysis a physical theoretical prediction of a unique tip
radius and velocity for a given set of growth conditions can only be framed
in the presence of surface energy anisotropy.

5.1 Free-dendritic growth: The paraboloid


approximation and Ivantsov solution

The first attempt towards formulating a solution for predicting the shape
of the dendrite was from Ivantsov. The approach was to assume the
shape of the dendrite near to the tip of the dendrite. as a parabola in
2D and parabaloid in 3D. The parabolic tip was assumed to move in a
self-similar manner at steady state, with a velocity given by 𝑣 * and radius
5.1 Free-dendritic growth: The paraboloid approximation and Ivantsov solution
53

of curvature at the tip given by 𝑅* . In this first analysis, the effect of the
Gibbs-Thomson effect is assumed to be absent so that the entire growth
shape of the solid is assumed to be isothermal. This assumption, as we will
see later is important to relax in order to formulate a self-consistent theory.
Also, in the present analysis we will consider the dendritic growth of a
pure material. However, the results will be easily carried over to represent
the c ase of solutal dendrites.
Consider a co-ordinate system where the growth directions is given by 𝑧
and thereby in the cylindirical co-ordinate system the shape of the dendrite
can be written as,

𝑅𝑡𝑖𝑝 𝑟2
𝑧= − , (5.1)
2 2𝑅𝑡𝑖𝑝

where 𝑟 is the distance from the central axis of revolution of the paraboloid.
Non-dimensionalizing the length by using 𝑅𝑡𝑖𝑝 as the scale, we derive,

1 (︀
1 − 𝑅2 = 𝑃 (𝑅) ,
)︀
𝑍= (5.2)
2

where 𝑍 = 𝑧/𝑅𝑡𝑖𝑝 and 𝑅 = 𝑟/𝑅𝑡𝑖𝑝 . The temperature in the solid is isother-


mal and fixed at 𝑇 * . Scaling the temperature using this temperature at the
interface, formulating a variable representing the variation of temperature
𝑇𝑙 − 𝑇∞ 𝐶𝑝𝑙 Δ𝑇
in the liquid 𝜃𝑙 as * definining the Stefan number 𝑆𝑡 = ,
𝑇 − 𝑇∞ 𝐿𝑓
*
𝑣 𝑅
where Δ𝑇 = 𝑇 * − 𝑇 ∞ and the Peclet number Pe = , 𝛼𝑙 being the
2𝛼𝑙
thermal diffusivity in the liquid phase. With this scaling the governing
equations for the diffusion in the solid written in the moving co-ordinate
system fixed at the center of the circle sharing a tangent with the tip of
the dendrite and translating with a velocity 𝑣 * derives as,

𝜕𝜃𝑙
−2𝑃 𝑒 = ∇2 𝜃 𝑙 𝑍 > 𝑃 (𝑅) . (5.3)
𝜕𝑍
54 5 Physics of Dendritic growth

The surface of the dendrite is isothermal hence,

𝜃𝑙 = 1 𝑍 = 𝑃 (𝑅) . (5.4)

The Stefan condition in this dimensionaless co-ordinate system writes as,

2𝑃 𝑒
𝑛𝑧 = −∇𝜃𝑙 .n, 𝑍 = 𝑃 (𝑅) (5.5)
𝑆𝑡

where 𝑛𝑧 is the normal vector in the growth direction and n is the normal
vector to the interface of the dendrite.
A convenient transformation adopted by Ivantsov here was to utilize a
parabolic co-ordinate system, formed out of a intersection of a continuos
set of two orthogonal parabolas. This allows for easier writing down of the
Stefan boundary conditions which depend on the interface normal. The
transformation writes as,

𝑅 = 𝜉𝜂 (5.6)
1 (︀ 2
𝜉 − 𝜂2 ,
)︀
𝑍= (5.7)
2

where 𝜉 and 𝜂 are our new transformed co-ordinates each of which can
be expressed in terms of 𝑍 and 𝑅 and drawn as parabolas. This can be
seen as follows: eliminate 𝜂 in favour of 𝜉 using the above transformations,
This writes as,

𝑅2
(︂ )︂
1 2
𝑍= 𝜉 − 2 or (5.8)
2 𝜉
𝜉 4 − 2𝑍𝜉 2 − 𝑅2 = 0. (5.9)

Solving for 𝜉 2 and taking the real root we derive,


5.1 Free-dendritic growth: The paraboloid approximation and Ivantsov solution
55

√︀
𝜉2 = 𝑍 + 𝑅2 + 𝑍 2 . (5.10)

Similarly, eliminating 𝜉 in favour of 𝜂 one can derive the other relation


as,

√︀
𝜂 2 = −𝑍 + 𝑅2 + 𝜉 2 , (5.11)

which represent the two sets of different parabolas for each set of constant
𝜉, 𝜂. Please note that for 𝜉 = 1, one derives the parabola representing the
1 (︀ )︀
surface of the dendrite, 𝑍 = 1 − 𝑅2 .
2
After this, the evolution equations and the boundary conditions are trans-
formed using the preceding set of transformations. We will not go into the
solution discourse at present, and will directly present the set of solutions
that one derives,

(︀ )︀
𝐸1 𝑃 𝑒𝜉 2
𝜃𝑙 (𝜉) = , 𝜉 ≥ 1, (5.12)
𝐸1 (𝑃 𝑒)

where 𝐸1 is the exponential integral,


𝑒−𝑠
∫︁
𝐸1 (𝑢) = 𝑑𝑠; (5.13)
𝑢 𝑠

the differential of the above integral with respect to the variable 𝜉 can be
calculated using the preceding definition as,

𝜕𝐸1 (𝑢) 𝑒−𝑢 𝑑𝑢


=− . (5.14)
𝜕𝜉 𝑢 𝑑𝜉
56 5 Physics of Dendritic growth

Using the Stefan condition computed at the tip in the transformed co-
ordinate system which now reads as,

2𝑃 𝑒 𝜕𝜃𝑙
=− . (5.15)
𝑆𝑡 𝜕𝜉

Evaluating the boundary condition at 𝜉 = 1, which represents the surface


of the dendrite at the dendrite tip, we derive the relation between the
Stefan number and the Peclet number, which is the basic continuity
equation arising out of the diffusion equations with the approximation of
the parabolic shape of the dendrite. The relation derives as,

𝑆𝑡 = 𝑃 𝑒 exp (𝑃 𝑒) 𝐸1 (𝑃 𝑒) . (5.16)

Similarly, in 2D the solution for the temperature field in the liquid writes
as,

√ (︁ √ )︁
𝜋𝑃 𝑒
𝜃𝑙 (𝜉) = exp (𝑃 𝑒) 𝑒𝑟𝑓 𝑐 𝜉 𝑃 𝑒 , 𝜉 ≥ 1, (5.17)
𝑆𝑡
∫︀ 𝑥 (︀ )︀
where 𝑒𝑟𝑓 𝑐 (𝑥) is the error function described as 0
exp −𝑢2 𝑑𝑢 and the
relation to the Stefan number written as,

√ (︁√ )︁
𝑆𝑡 = 𝜋𝑃 𝑒 exp (𝑃 𝑒) 𝑒𝑟𝑓 𝑐 𝑃𝑒 . (5.18)

√ 𝑃 𝑒 exp (𝑃 𝑒) 𝐸1 (𝑃
We will represent the functions 𝑒) as 𝐼𝑣3𝐷 (𝑃 𝑒) and
(︁ √ )︁
𝜋𝑃 𝑒
the two dimensional variant exp (𝑃 𝑒) 𝑒𝑟𝑓 𝑐 𝜉 𝑃 𝑒 as 𝐼𝑣2𝐷 (𝑃 𝑒).
𝑆𝑡
Hypothetically consider that the inverse of the integral functions can be
computed, such that we can express the Peclet number as a function of
the Stefan number as follows,
5.2 Subsequent theories 57

𝑃 𝑒 = 𝐼𝑣 −1 (𝑆𝑡) . (5.19)

Clearly, this solution derives a relation between the the Peclet number
which is the product of the tip velocitc 𝑣 * and the tip radius 𝑅𝑡𝑖𝑝 and the
scaled undercooling in the system, which however, does not allow one to
determine a unique combination of 𝑣 * and 𝑅𝑡𝑖𝑝 for a given set of processing
conditions, as has been reported in experiments.

5.2 Subsequent theories

To resolve the issue that only a continuum family of solution results out
of the Ivantsov calculation, a number of subsequent theories were placed.
Temkin, was the first to point out that a critical parameter which is the
surface tension, that relates to the surface tension, has been left out of the
calculation scheme. Temkin performed a calculation where he used the a
modified expression for the Stefan number by correcting for the curvature
undercooling. The relation upon rendering this correction reads as,

2Γ𝑠𝑙 𝑐𝑝
𝑆𝑡 − = 𝑃 𝑒 exp (𝑃 𝑒) 𝐸1 (𝑃 𝑒) . (5.20)
𝑅𝑡𝑖𝑝 𝐿𝑓

This now admits a solution for a undercooling corresponding to 𝑆𝑡 =


2Γ𝑠𝑙 𝑐𝑝
, such that the right hand side is zero, which is derived at zero
𝑅𝑡𝑖𝑝 𝐿𝑓
velocity of the interface. As R increases, now there are two effects, decrease
of velocity as an increase of radius since the product is contrained to be
2Γ𝑠𝑙 𝑐𝑝
some function of 𝑆𝑡 − , and the decrease of the effective Stefan
𝑅𝑡𝑖𝑝 𝐿𝑓
number with increase in R. This brings about a maximum, beyond which
2Γ𝑠𝑙 𝑐𝑝
the contribution of is significantly lower and thereby there is
𝑅𝑡𝑖𝑝 𝐿𝑓
decrease in velocity with increasing 𝑅.
58 5 Physics of Dendritic growth

Temkin put forth a conjecture that the point of extremum must be the
operating point of the tip. This was however, held to be inconsistent with
experimental observations.
One of the errors in the computations was of course, that the Gibbs-
Thomson effect was accounted for, only at the tip of the dendrite. This was
corrected with accurate analytical calculations using Green’s functions for
the solutions to the heat equation by Nash and Glicksman. Unfortunately,
the solutions were only slightly different from that of Temkin.
A first computation of a dynamical evolution of a dendritic surface was per-
formed by Oldfield. Starting with an Ivantsov shape for the parabola, the
temperature field was solved at each step of evolution by using the Gibbs-
Thomson condition at the interface and convergence with the Ivantsov
solution far away from the tip of the dendrite. The interface was advanced
using the velocity derived out of the Stefan boundary condition. Oldfield
reported that the interface with surface energy, was unstable and was
able to derive only “split tips”. Thus, he formulated a stability criterion,
resulting from a balance between the surface energy and the perturbation
in the heat equation. This is on the same lines as computation of the
critical wavelengths in the Mullins-Sekerka analysis.
Langer and Muller-Krumbhaar formalized the observation by Oldfield,
by performing a stability analysis and proving that the needle crystal
is unstable for all finite wavelengths of isotropic surface energy. This
analysis goes by the name of marginal stability. Essentially, it is statement
motivated from the critical wavelength arising in Mullins-Sekerka analysis
of a planar front, and thus,

√︀
𝑅𝑡𝑖𝑝 = 𝜆𝑚𝑖𝑛 = 2𝜋 𝑑𝑜 𝐷𝑙 , (5.21)

where 𝑑𝑜 is the capillary length and 𝐷𝑙 is the diffusion length. Reformu-


2
𝑅𝑡𝑖𝑝 𝑣* 2
lating, the marginal stability criterion reads as, = (1/2𝜋) , with
𝑑𝑜 𝛼𝑙
1
𝜎* = .
2𝜋
The fundametal problem however, was that still there was no unique
𝑣 * , 𝑅𝑡𝑖𝑝 , that resulted out of the analysis. The next steps were in a
5.3 Simplified models 59

direction towards formulating a solution for the diffusion problem, with


surface energy anisotropy. Since, the problem needs to be solved in a
large domain the models needed to be efficient. The first modes were
approximate, with the full solution limited only to a few layers near the
interface, while far away the Ivantsov solution is used as the boundary
condition. The results of such models, retrieved the basic results of Oldfield,
Nash and Glicksman, that with surface energy, the interface is unstable.
With the incorporation of anisotropy however, the authors retrieved a
special observation, that instead of continuous family of solutions only a
few discrete pairs of solutions were allowed, and among them only one of
them, which is the fastest growing solution is stable.
Eventually, Kessler and Levine put-forth a numerical stability analysis
of the dendritic interface. The general solutions of this stability analysis
give rise to shapes which are cusps, (non-zero slopes at the tip). Only one
distinct solution with a smooth dendrite tip exists, which is derived by
the microscopic solvability condition. The microscopic solvability criterion
2
also yields a condition of the form 𝑅𝑡𝑖𝑝 𝑣 * = 𝐶 (𝜖), where the constant is
now a function of 𝜖, the strength of anisotropy.

5.3 Simplified models


A number of simplified models have been proposed to analyze the problem
of determinining the velocity and tip radius given the undercooling. We
will survey one of the major theories, whose variants are also used in
macroscopic simulations of dendritic growth.

5.3.1 Lipton-Glicksman-Kurz (LGK) theory

The LGK theory is one of the mostly used approximate theories of dendritic
growth. Here, the authors use the parabaloid to represent the dendritic
shape. The total undercooling of the interface is written with respect
to the melting point of the starting alloy composition and thereafter the
radius and the velocity of the tip are derived as a function of undercooling.
The undercooling at the tip contains all the contributions, namely, the
effect of the Gibbs-Thomson condition, the change in melting point due
60 5 Physics of Dendritic growth

to the variation of the composition and the imposed thermal field. The
operating state of the tip of the parabola is assumed to be derived from
the marginal stability criterion, wherein the radius of the tip is equated
to the minimum wavelength of perturbation that is critical in the Mullin-
Sekerka theory of planar front growth. It is essential to point out that the
authors choose a factor of this wavelength in order to match the observed
experimental results. The solutions obtained from the intersection of the
Ivantsov solution and the stability criterion derives, the radius and velocity
of the tip of the parabola, which are found to be in good agreement with
the experimental observations.

In the following, we work through the framework of LGK, and its simplifi-
cations.

We start by writing the total undercooling in the system as,

Δ𝑇 = Δ𝑇𝑡 + Δ𝑇𝑐 + Δ𝑇𝜅 . (5.22)

Given, that the temperature at the interface is 𝑇 * and the temperature


𝐿𝑓
at 𝑇∞ , we can write Δ𝑇𝑡 = 𝑇 * − 𝑇∞ = 𝐼𝑣3𝐷 (𝑃 𝑒𝑇 ), where 𝑃 𝑒𝑇 is the
𝐶𝑝
𝑣𝑅
thermal Peclet number given by . The constitutional undercooling
2𝛼𝑙
with respect to the starting alloy composition can be written by using
the properties of the phase diagram namely the liquidus slope and the
partition coefficients. Given that the composition at the interface is
𝐶𝑙* , the far-field composition is 𝐶0 , the partition coefficient is k and the
negative of the liquidus slope is 𝑚, the constitutional undercooling is
given by 𝑚 (𝐶𝑙* − 𝐶0 ), where 𝐶𝑙* can be eliminated using the Ivantsov
𝐶 * − 𝐶0
solution for the composition *𝑙 = 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 ) and written as
𝐶𝑙 (1 − 𝑘)
𝑚𝐶0 (1 − 𝑘) 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 )
, where 𝑃 𝑒𝐶 is the solutal Peclet number given
1 − (1 − 𝑘) 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 )
𝑣𝑅 2Γ𝑠𝑙
by . The curvature undercooling is given by where Γ𝑠𝑙 is the
2𝐷𝑙 𝑅
Gibbs-Thomson coefficient. Thus, the total undercooling writes as,
5.3 Simplified models 61

𝐿𝑓 𝑚𝐶0 (1 − 𝑘) 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 ) 2Γ𝑠𝑙


Δ𝑇 = 𝐼𝑣3𝐷 (𝑃 𝑒𝑇 ) + + . (5.23)
𝐶𝑝 1 − (1 − 𝑘) 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 ) 𝑅𝑡𝑖𝑝

The stability criterion which equates the radius of the tip to the minimum
critical wavelength in the Mullins-Sekerka analysis of a solidification front
in presence of a thermal gradient writes as,

√︃
Γ𝑠𝑙
𝑅 ≈ 2𝜋 . (5.24)
(−𝑚𝑙 𝐺𝐶 − 𝐺)

To proceed further, we require the solutal and the thermal derivatives


𝐺𝑐 and 𝐺 respectively. These derivatives can be computed using the
composition and the temperature profiles derived as solution to the Ivantsov
solutions. These can be written as,

(︀ )︀
𝑇𝑙 − 𝑇∞ 𝐸1 𝑃 𝑒𝑇 𝜉 2
𝜃𝑙 = * = (5.25)
𝑇 − 𝑇∞ 𝐸1 (𝑃 𝑒𝑇 )
(︀ 2
)︀
𝐶𝑙 − 𝐶0 𝐸1 𝑃 𝑒𝐶 𝜉
Ω𝑙 = * = . (5.26)
𝐶𝑙 − 𝐶0 𝐸1 (𝑃 𝑒𝐶 )

𝜕𝜃𝑙 𝜕𝐶𝑙
We need the derivatives , where 𝑧 is the growth direction. We
𝜕𝑧 𝜕𝑧
use the definition of the parabola that is given at 𝜉 = 1 and at the tip
1
𝑅 = 𝜉𝜂 = 0 on account of 𝜂 = 0, also 𝑍 = 𝜉 2 . Therefore 𝑑𝑍 = 𝑑𝑧/𝑅𝑡𝑖𝑝 =
2
𝜉𝑑𝜉 which at 𝜉 = 1 (representing the envelope of the dendrite) reads,
𝜕𝜃𝑙 𝜕𝐶𝑙
𝑑𝑧 = 𝑅𝑡𝑖𝑝 𝑑𝜉. Therefore, the derivatives and can be written as
(︂ )︂ (︂ 𝜕𝑧 )︂ 𝜕𝑧
𝜕𝜃𝑙 1 𝜕𝜃𝑙 𝜕𝐶𝑙 1 𝜕𝐶𝑙
= and = .
𝜕𝑧 𝑅𝑡𝑖𝑝 𝜕𝜉 𝜉=1 𝜕𝑧 𝑅𝑡𝑖𝑝 𝜕𝜉 𝜉=1
Using the definition of exponential integrals, we can now write the deriva-
𝜕𝜃𝑙 exp (−𝑃 𝑒𝑇 ) 1
tives explicitly as = − (𝑇 * − 𝑇∞ ) , which upon using
𝜕𝑧 𝐸1 (𝑃 𝑒𝑇 ) 𝑅𝑡𝑖𝑝
62 5 Physics of Dendritic growth

𝑃 𝑒𝑇 𝐿𝑓
the Ivantsov solution reads as − . Similarly, we can derive the
𝑅𝑡𝑖𝑝 𝐶𝑝
*
−𝑃 𝑒𝐶 𝐶𝑙 (1 − 𝑘)
solutal gradient as − . Using the Ivantsov solution and
𝑅𝑡𝑖𝑝
*
eliminating 𝐶𝑙 , we have the 𝑅𝑡𝑖𝑝 determined in terms of the Peclet number
from the stability criterion as,

(︂ )︂−1
2 𝑃 𝑒𝑇 𝐿𝑓 𝑃 𝑒𝐶 𝑚𝐶0 (1 − 𝑘)
𝑅𝑡𝑖𝑝 = 4𝜋 Γ𝑠𝑙 + . (5.27)
𝐶𝑝 1 − (1 − 𝑘) 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 )

Solving Eqns. 5.27 and 5.23 iteratively one can determine the product
𝑣𝑅𝑡𝑖𝑝 and 𝑅𝑡𝑖𝑝 and thereby derive both 𝑣 and 𝑅𝑡𝑖𝑝 .

5.3.2 Simplifications

Given the numerical complexity of solving the coupled system of equations


in the preceding section, it sometimes becomes numerically demanding
when one wants to perform this computation for a large number of grains
in an evolving domain. Some simplifications could be made, which are
applicable for most practical situations. Firstly, normal castings work
in the limit 𝑃 𝑒𝐶 ≪ 1, i.e the radius of the tip is much smaller than the
diffusion length. In this limit, given that the thermal diffusivity is about
103 times the solutal diffusivity, the thermal Peclet number tends to zero in
this limit. Following this, it is clear that the thermal undercooling arising
because of the variation of the temperature field is much smaller than
the variation in undercooling arising out of the variation in composition.
Additionally, the magnitude of the curvature undercooling is normally of
very small magnitude(≈0.01K) for dendrites at the mesoscopic scale. Thus,
the constitutional undercooling is the dominant contributor to the total
undercooling. With these approximations we can write an approximate
version of the 𝑅𝑡𝑖𝑝 derived from the marginal stability criterion as,

(︂ )︂
2 2 𝐷𝑙
𝑅𝑡𝑖𝑝 𝑣 = 8𝜋 Γ𝑠𝑙 (5.28)
𝑚𝐶0 (1 − 𝑘)
Δ𝑇 = 𝑚𝐶0 (1 − 𝑘) 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 ) , (5.29)
5.4 Primary and secondary dendrite arm spacings 63

where we have also used the approximation 𝐼𝑣3𝐷 (𝑃 𝑒𝐶 ) ≪ 1 when 𝑃 𝑒𝐶 ≪


1.
The above system of equations is still difficult to solve given that it involves
the exponential integral. A way out, is to derive an approximate value for
𝐼𝑣3𝐷 (𝑃 𝑒𝐶 ) in the limit of small peclet numbers. One of the commonly
used ones is,

0.8
𝐼𝑣3𝐷 (𝑃 𝑒𝐶 ) = 1.5 (𝑃 𝑒𝐶 ) . (5.30)

Using the expressions for the previous two relations of 𝑅𝑡𝑖𝑝 and Δ𝑇 , we
can derive closed form relations for 𝑣 and 𝑅, such that large system of
grains can also be solved for, in an efficient manner.

5.4 Primary and secondary dendrite arm


spacings
Until now we have seen how to determine one of the length scales relevant
to dendritic growth which is the radius at the tip of the dendrite. In
the following section, we determine some more relevant microstructural
measures relevant to dendrites which are the primary dendrite arm spacing
and the secondary dendrite arm spacing. While the primary arm spacing is
a reflection of the history of the solidification of the sample the secondary
arm spacing as we will see is a function of the solidification time, the
important mechanism of change in spacing with time being coarsening.
In the following we first derive expressions for the primary arm spacing in
the case of constrained solidification, wherein, the solidification occurs in a
temperature gradient which is positive in the growth direction and initially
moves with respect to the sample at a prescribed velocity. At steady state,
the interface moves with the same velocity as that of isotherms, such that
in the frame of reference fixed to the isotherms, there is no change in the
position of the mean position of the interface.
In this setting, the variables are indeed the radius of the tip and the
inter-dendritic spacing, which indicates the extent of flux coupling between
64 5 Physics of Dendritic growth

the adjacent dendrites. A simple estimate of the primary dendrite arm


spacing can be achieved by first approximating the shape of the envelope
formed by the singular primary arms with its secondary arms. Clearly, this
shape cannot be a parabola, since this would imply that the envelope is
diverging with distance from the tip, which cannot be the case, because of
the overlap of the diffusion fields of neigboring dendrites. An approximate
shape would be an ellipsoid with a major axes 𝑎 and minor axes 𝑏. We
assume a hexagonal configuration of these envelopes, with the envelopes
intersecting touching at their lowest point. The distance between the
neighboring tips of the envelopes is the primary dendrite arm spacing 𝜆1 .
Relating the geometric features with the properties of the ellipsoid we
get 𝜆1 = 2𝑏. Similarly, the radius of curvature a the tip of the envelope
matches the dendrite tip radius of the primary arm, which is given by the
𝑏2 𝜆2
properties of the ellipsoid as 𝑅𝑡𝑖𝑝 = = 1.
𝑎 4𝑎

The length of the major axes 𝑎 can be fixed using the information of the
temperature gradient. Usually, the difference of temperature between the
tips of the dendrite, which is the first liquid that solidifies, and the base
which is the last liquid that solidifies, is nothing but the freezing range
corresponding to the alloy composition, Δ𝑇0 . In some instances, the final
reaction is interrupted through a secondary reaction which is the formation
of an eutectic in which case, this difference of temperatures between the
tip of the dendrite and the base would be modified. Therefore, we will
assume, that we have some knowledge of the difference of temperatures

between the tip of the dendrite and the base and write it as Δ𝑇0 . Thus,

Δ𝑇0
the length of the major axes can now be determined as, 𝑎 = , where
𝐺
𝐺 is the magnitude of the imposed temperature gradient.
(︂ )︂
2 2 𝐷𝑙
Now using, 𝑅𝑡𝑖𝑝 𝑣 = 8𝜋 Γ𝑠𝑙 as derived from the marginal
𝑚𝐶0 (1 − 𝑘)
stability criterion, we can write the radius of the tip as,

(︂ (︂ )︂)︂1/2
2 𝐷𝑙
𝑅𝑡𝑖𝑝 = 8𝜋 Γ𝑠𝑙 (5.31)
𝑣𝑚𝐶0 (1 − 𝑘)
5.4 Primary and secondary dendrite arm spacings 65

𝜆21 𝐺
Also 𝑅𝑡𝑖𝑝 can be expressed as a 𝑅𝑡𝑖𝑝 = ′ , using the properties of the
4Δ𝑇0
ellipsoid. Simplifying, we derive an expression for the primary dendrite
arm spacing as,

′ 2
⎛ (︁ )︁ ⎞
2
128𝜋 Γ 𝐷
𝑠𝑙 𝑙 Δ𝑇0 ⎟ −1/4 −1/2
𝜆1 = ⎝ ⎠𝑣 𝐺 , (5.32)

Δ𝑇0

where we have used Δ𝑇0 = 𝑚𝐶0 (1 − 𝑘). A point to note is that the
exponents of 𝑣 and 𝐺 turn out to be different, which implies, that the rate
of change of 𝜆1 cannot be related to the change in the solidification rate
which is proportional to the cooling rate 𝑇˙ = 𝐺𝑣.

5.4.1 Secondary dendrite arm spacing

As was mentioned previously the length scale of the microstructure rep-


resented by the distance between the secondary arms changes over time
through a mechanism called coarsening. During coarsening, the concen-
tration in the far-field liquid is saturated and the diffusion is governed
by small gradients created due to the difference in compositions arising
out of the Gibbs-Thomson effects which cause geometrical entities to have
equilibrium interfacial compositions as functions of their radii of curvature.
Thus, in the case where the compositions are represented with respect to
the solute, the deviation of the composition in the liquid in front of a solid
of radius of curvature (𝑅) may be written as,

1
𝐶𝑅 − 𝐶𝑒𝑞 = Γ𝑠𝑙 , (5.33)
(−𝑚𝑅)

where 𝑚 is the modulus of the liquidus slope. Thus the difference in


equilibrium compositions between two secondary arms (assumed to have
cylindrical geometries), of radii 𝑅 and 𝑟 such that 𝑅 ≫ 𝑟, writes as,
66 5 Physics of Dendritic growth

(︂ )︂ (︂ )︂
Γ𝑠𝑙 1 1
𝐶𝑅 − 𝐶𝑟 = − . (5.34)
−𝑚 𝑅 𝑟

Thus, the mass transport of solute from the liquid near the solid of radius
of larger curvature towards the liquid near the smaller arm, writes as,

(︂ )︂ (︂ )︂
Γ𝑠𝑙 𝐷𝑙 1 1
𝑗= − . (5.35)
𝑚𝜆2 𝑅 𝑟

where 𝑑 effective distance over which the gradient exists. Using the Stefan
condition at the interface of the smaller radii, one can write down the rate
of change of radii as,

(︂ )︂ (︂ )︂
𝑑𝑟 Γ𝑠𝑙 𝐷𝑙 1 1
= − . (5.36)
𝑑𝑡 𝑚Δ𝐶𝑑 𝑅 𝑟

𝑙 𝑠
where Δ𝐶 = 𝐶𝑒𝑞 − 𝐶𝑒𝑞 , which is positive when the slope of the liquidus
is negative. Since, in the present derivation 𝑚 has been written as the
modulus of the liquidus slopes, therefore, the term in the first brackets
on the right hand side of the equation is positive. Given that 𝑅 ≫ 𝑟, the
smaller arm is going to shrink. Similarly, the larger arm is growing grow.
However, in the present illustration, since 𝑅 ≫ 𝑟, the percentage change in
the radii of the larger arm is small compared to the change in the smaller
arm. Therefore, we assume 𝑅 stays constant in the present derivation.
We also assume for simplicity that the Δ𝐶 remains constant in time, i.e.,
the difference in composition between the phases does not change with
temperature or with radius.
Also, note that we can write 𝑑 = 𝜆2 − 𝑅 − 𝑟. Inserting, this into the
relation, integrating, and making use of the approximation 𝑟/𝑅 ≪ 1, we
have

(︂ )︂
𝜆2 − 𝑅 2 1 3 Γ𝑠𝑙 𝐷𝑙
𝑟 − 𝑟 =− 𝑡. (5.37)
2 3 𝑚Δ𝐶
5.5 Rayleigh instability 67

Clearly, given the magnitudes 𝜆2 , 𝑅, 𝑟, the coefficient of exponent 𝑡1/3 is


going to be the largest, which is the characteristic of a coarsening mecha-
nism. The rate of change of the smaller radius of curvature reflects the
rate at which the secondary arm spacing changes, since the disappearance
of smaller arms results in the increase in the size of the secondary arm
spacing.

5.5 Rayleigh instability

A key mechanism operating in the breakdown of dendritic arms in 3D into


fragments is through an instability wherein a cylindrical object is unstable
to perturbations beyond a particular wavelength. This process is also
fragmentation. Fragmentation as a process is important to understand,
as it is often used to tailor the microstructural scale. Through, imposed
temperature cyling fragments are created which then act as nuclei for more
dendrites to form thus reducing the grain-size in the microstructure.

In the following, we briefly survey the origin of this instability which was
first noticed in the case of flow of droplets though a nozzle. Beyond a
particular flow velocity, the stream of droplets breaks down into smaller
droplets.

The origin of Rayleigh instability lies in the fact that beyond a particular
size, the surface area for the same unit volume of matter is smaller and
hence there exists a thermodynamic driving force for the change of form
from a cylindrical object to a sphere.

Consider a cylinder of initial radius 𝑅0 , and let us perturb the cylinder


2𝜋𝑧
with a time dependent cosine wave of the form 𝜖 (𝑡) cos , where 𝜆 is
𝜆
the wavelength of the perturbation and 𝑧 is the co-ordinate along the
cylindrical axis. The radius as a function of space and time with this
imposed perturbation can be written as,

2𝜋𝑧
𝑅 = 𝑅𝑐𝑦𝑙 (𝑡) + 𝜖 (𝑡) cos . (5.38)
𝜆
68 5 Physics of Dendritic growth

To derive the variation of the radius of cylinder as a function of time, we


can equate the volume at a given time 𝑡 to te initial volume, for a one
wavelength of the parturbed of cylinder. This writes as,

∫︁ 𝜆
𝜋𝑅02 𝜆 = 𝜋𝑅2 𝑑𝑧 (5.39)
0
∫︁ (︂𝜆 )︂
2 2𝜋𝑧 2 2 2𝜋𝑧
= 𝜋 𝑅𝑐𝑦𝑙 + 2𝑅𝑐𝑦𝑙 𝜖 (𝑡) cos + 𝜖 (𝑡) cos (5.40)
0 𝜆 𝜆
2
2 𝜋𝜖 (𝑡) 𝜆
= 𝜋𝑅𝑐𝑦𝑙 𝜆+ (5.41)
2

Thus, the radius as a function of time writes as,

)︂1/2
𝜖2 (𝑡)
(︂
𝑅𝑐𝑦𝑙 (𝑡) = 𝑅0 1 − . (5.42)
2𝑅02

In the limit of smaller perturbations, we can use binomial approximation


and write it as,

𝜖2 (𝑡)
𝑅𝑐𝑦𝑙 (𝑡) = 𝑅0 − . (5.43)
4𝑅0

Using the preceding relation, we can now write down the surface area as a
function of time as,

∫︁ 𝜆
𝐸𝑠 (𝑡) = 2𝜋𝑅𝑑𝑠, (5.44)
0

√︃ )︂2

(︂
𝑑𝑟
where 𝑑𝑠 is the arc length written as 𝑑𝑟2
= 𝑑𝑧 1 + + 𝑑𝑧 2 . Using
𝑑𝑧
(︂ )︂
𝑑𝑟 2𝜋 2𝜋𝑧
𝑅𝑡 one can determine the derivative =− sin .
𝑑𝑧 𝜆 𝜆
5.5 Rayleigh instability 69

Inserting the preceding relation into the expression for surface energy we
have,

⎛⎯ (︂ )︂ ⎞
2𝜋𝑧

4𝜋 2 𝜖2 sin2

∫︁ 𝜆 (︂
𝜖2
)︂ ⎜⎸ ⎟
cos 2𝜋𝑧 ⎜ ⎷ 𝜆 ⎟
𝐸𝑠 (𝑡) = 2𝜋 𝑅0 − +𝜖 ⎜ 1+

2

0 4𝑅0 𝜆 ⎝ 𝜆 ⎟

(5.45)

Using the binomial approximation and limiting the analysis only to 𝜖2 , we


have the expression for the surface area as,

𝜖2 𝜆 𝜋 3 𝑅0 𝜖2
𝐸𝑠 (𝑡) = 2𝜋𝑅0 𝜆 − 2𝜋 +4 (5.46)
4𝑅0 2𝜆
𝜖2 (︀ 2
𝜆 − 4𝜋 2 𝑅02 .
)︀
= 2𝜋𝑅0 𝜆 − (5.47)
2𝑅0 𝜆

Clearly, for any perturbation with 𝜆 > 2𝜋𝑅0 we are going to have reduction
in surface energy with increase in the amplitude of the perturbation, making
the cylinder unstable. This is called the Rayleigh instability.
To describe the actuall mechanism through which the perturbation grows
𝜖 (𝑡), one requires more information about the coupling between the motion
of the interface and mass transport.
6 Multi-phase solidification
72 6 Multi-phase solidification

In materials science there exists a number of invariant reactions, namely


points on the phase diagram corresponding where the degrees of freedom
with respect to the thermodynamical variables is zero. Among them, there
are certain, reactions which involve transformation of liquid to one or
more solids. These reactions are three in number, i) eutectic which is the
transformation of one liquid to two or more solids 𝐿 → 𝛼 + 𝛽 + . . ., ii)
peritectic, which is the transformation of a liquid in combination with a
solid to give one or more solids and thirdly iii) monotectic, which is the
transformation of a liquid to another liquid and a solid.

All of these reactions, exhibit a rich variety of multi-phase co-existince


and pattern forming systems and thus causing interest among both physi-
cists and material scientists. While, eutectic growth has been fairly well
understood with a fair deal of analytical, experimental and mathematical
modeling effort, the corresponding knowledge in the case of peritectics and
monotectics is still in progess.

In the following, we review a classical coupled growth analysis performed by


Jackson and Hunt. The assumption, here is that the solute profiles in front
of the two growing phases is derived using a planar front approximation,
while the contributionn to the curvature undercooling is additionally
incorporated when accounting for the undercooling at the interface.
6.1 Mechanism of eutectic growth 73

6.1 Mechanism of eutectic growth

𝑔(𝑐)

(𝑇 = 𝑇𝐸 )

𝑐
(𝐶𝐸 )

𝑔(𝑐)

(𝑇 < 𝑇𝐸 )

𝑐
(𝐶𝛽𝑙 ) (𝐶𝛼𝑙 )

The mechanism of eutectic growth can be understood on the basis of the


free energy curves depicted in figures 6.1 and 6.1. At the eutectic point,
the solid phases are in equilibrium with the liquid, with the composition of
the liquid being 𝐶𝐸 . For an undercooled temperature, the solid phases are
at a lower temperature as compared to the liquid phase. The equilibrium
74 6 Multi-phase solidification

compositions of liquid in equilibrium with the phase 𝛼 is given by 𝐶𝛼𝑙 ,


which is given by the common tangent construction between the two phases.
Similarly, the composition of the liquid 𝐶𝛽𝑙 is in equilibrium with the solid
of 𝛽. Ahead of a lamellar front as shown in the following figure,

(︁ )︁
𝐶𝛽𝑙 − 𝐶𝛼𝑙
the compositions thus have gradient proportional to , where 𝜆
𝜆
is the spacing between the 𝛼 and the 𝛽. This diffusional length scale which
is much smaller compared to the diffusion length in the liquid, dominates
the mass transfer. One can see that the flux due to this gradient, would
encourage the flux of atoms, such that the growth is co-operative.

6.2 Jackson-Hunt analysis

To start with, our growth direction is going to be 𝑧, and the transverse


direction is 𝑥. For this geometry the relevant approximate concentration
profile for a binary alloy, is the following,


∑︁
𝑐= 𝑋𝑛 𝑒𝑖𝑘𝑛 𝑥−𝑞𝑛 𝑧 + 𝑐∞ , . (6.1)
𝑛=−∞
6.2 Jackson-Hunt analysis 75

In Eq. (6.1), 𝑘𝑛 = 2𝜋𝑛/𝜆 are wave numbers and 𝑞𝑛 can be determined


from the solutions of the stationary diffusion equation

𝑣𝜕𝑧 𝑐 + 𝐷∇2 𝑐 = 0,

which yields √︂
𝑣 (︁ 𝑣 )︁2
𝑞𝑛 = + 𝑘𝑛2 + .
2𝐷 2𝐷

Thus, we define

Δ𝑐 = 𝑐𝑙 − 𝑐𝜈 with 𝜈 = 𝛼, 𝛽.

with 𝜈 = 𝛼, 𝛽 being the two solid phases.


In this approximation, the Stefan condition at a 𝜈-𝑙 interface, which
expresses mass conservation upon solidification, reads
𝑣𝑛 𝜈
𝜕𝑛 𝑐𝑗 = − Δ𝑐𝑗 , (6.2)
𝐷
where 𝜕𝑛 𝑐𝑗 denotes the partial derivative of 𝑐𝑗 in the direction normal
to the interface, 𝑣𝑛 is the normal velocity of the interface (positive for a
growing solid), and 𝐷 is the chemical diffusion coefficient, for simplicity
assumed to be equal for all the components. For, the present discussion,
the normal direction is going to be z.
We consider a general periodic lamellar array with 𝑀 repeating units
consisting of phases (𝜈0 , 𝜈1 , 𝜈2 , . . . , 𝜈𝑀 −1 ) where each 𝜈𝑖 represents the
name of one solid phase (𝛼, 𝛽) in the sequence, with a repeat distance
(lamellar spacing) 𝜆. The width of the 𝑗-th single solid phase region is
(𝑥𝑗 − 𝑥𝑗−1 ) 𝜆, with 𝑥0 = 0 and 𝑥𝑀 = 1, and the sum of all the widths
corresponding to any given phase is its volume fraction 𝜂𝜈 . The eutectic
front is assumed to grow in the 𝑧 direction with a constant velocity 𝑣.
For all the modes 𝑛 ̸= 0, we thus have 𝑞𝑛 ≃ |𝑘𝑛 | for small Peclet number
Pe = 𝜆/ℓ ≪ 1 with ℓ = 2𝐷/𝑣𝑛 , which will always be the case for slow
growth. The mode 𝑛 = 0 describes the concentration boundary layer which
is present at off-eutectic concentrations, and which has a characteristic
length scale of ℓ.
76 6 Multi-phase solidification

To determine the coefficients 𝑋𝑛 in the above Fourier series, we assume the


eutectic front to be at the 𝑧 = 0 position. Using the Stefan condition in Eq.
(6.2) and taking the derivative of 𝑐𝑋 with respect to the 𝑧-coordinate

∑︁
𝜕𝑧 𝑐𝑋 |𝑧=0 = −𝑞𝑛 𝑋𝑛 𝑒𝑖𝑘𝑛 𝑥 ,
𝑛=−∞

integration across one lamella period 𝜆 of arbitrary partitioning of phases


gives
𝑀 −1 ∫︁
2 ∑︁ 𝑥𝑗+1 𝜆 −𝑖𝑘𝑚 𝑥 𝜈𝑗
𝑞𝑛 𝑋𝑛 𝛿𝑛𝑚 𝜆 = 𝑒 Δ𝑐 𝑑𝑥, (6.3)
ℓ 𝑗=0 𝑥𝑗 𝜆

so that the coefficients 𝑋𝑛 , 𝑛 ∈ IN in the series ansatz, Eq. (6.1) follow


𝑀 −1
4 ∑︁
𝑋𝑛 = Δ𝑐𝜈𝑗 𝑒−𝑖𝑘𝑛 𝜆(𝑥𝑗+1 +𝑥𝑗 )/2 sin(𝑘𝑛 𝜆(𝑥𝑗+1 − 𝑥𝑗 )/2)
ℓ𝑞𝑛 𝜆𝑘𝑛 𝑗=0
. (6.4)
Applying symmetry arguments for the sinus and cosinus functions, we can
formulate real combinations of these coefficients if we additionally take
the negative summation indices into account. We obtain
𝑀 −1
8 ∑︁ 𝜈
𝑋𝑛 +𝑋−𝑛 = Δ𝑐𝑋𝑗 cos(𝑘𝑛 𝜆(𝑥𝑗+1 +𝑥𝑗 )/2) sin(𝑘𝑛 𝜆(𝑥𝑗+1 −𝑥𝑗 )/2),
ℓ𝑞𝑛 𝜆𝑘𝑛 𝑗=0

𝑀 −1
8 ∑︁ 𝜈
𝑖(𝑋𝑛 −𝑋−𝑛 ) = Δ𝑐𝑋𝑗 sin(𝑘𝑛 𝜆(𝑥𝑗+1 +𝑥𝑗 )/2) sin(𝑘𝑛 𝜆(𝑥𝑗+1 −𝑥𝑗 )/2).
ℓ𝑞𝑛 𝜆𝑘𝑛 𝑗=0

Herewith, Eq. (6.1) reads:


𝑀 −1 ∑︁

∑︁ 8
𝑐𝑋 = 𝑐∞
𝑋 + 𝑋0 + cos(𝑘𝑛 𝜆(𝑥𝑗+1 + 𝑥𝑗 )/2)×
𝑗=0 𝑛=1
ℓ𝑞𝑛 𝜆𝑘𝑛

sin(𝑘𝑛 𝜆(𝑥𝑗+1 − 𝑥𝑗 )/2) cos(𝑘𝑛 𝑥)+


𝑀 −1 ∑︁

∑︁ 8
sin(𝑘𝑛 𝜆(𝑥𝑗+1 + 𝑥𝑗 )/2) sin(𝑘𝑛 𝜆(𝑥𝑗+1 − 𝑥𝑗 )/2) sin(𝑘𝑛 𝑥).
𝑗=0 𝑛=1
ℓ𝑞 𝑛 𝜆𝑘𝑛
6.2 Jackson-Hunt analysis 77

The general expression for the mean concentration ⟨𝑐𝑋 ⟩𝑚 ahead of the
𝑚-th phase of the phase sequence can be calculated to yield
∫︁ 𝑥𝑚+1 𝜆
1
⟨𝑐𝑋 ⟩𝑚 = 𝑐𝑋 𝑑𝑥
(𝑥𝑚+1 − 𝑥𝑚 )𝜆 𝑥𝑚 𝜆
∞ 𝑀 −1 {︁
1 ∑︁ ∑︁ 16 𝜈
= 𝑐∞
𝑋 + 𝑋0 + Δ𝑐 𝑗
𝑥𝑚+1 − 𝑥𝑚 𝑛=1 𝑗=0 𝜆 𝑘𝑛2 ℓ𝑞𝑛 𝑋
2

}︁
sin[𝜋𝑛(𝑥𝑚+1 − 𝑥𝑚 )] × sin[𝜋𝑛(𝑥𝑗+1 − 𝑥𝑗 )] cos[𝜋𝑛(𝑥𝑚+1 + 𝑥𝑚 − 𝑥𝑗+1 − 𝑥𝑗 )] .
(6.5)

For a repetitive appearance of a phase 𝜈 in the phase sequence, the mean


concentration of component 𝑋 ahead of this phase follows by taking the
weighted average of all the lamellae of phase 𝜈,
∑︀𝑀 −1 {︂
𝑚=0 ⟨𝑐𝑋 ⟩𝑚 (𝑥𝑚+1 − 𝑥𝑚 )𝛿𝜈𝑚 𝜈 1 for 𝜈 = 𝜈𝑚
⟨𝑐𝑋 ⟩𝜈 = with 𝛿𝜈𝑚 𝜈 =
∑︀𝑀 −1 0 for 𝜈 ̸= 𝜈𝑚 .
𝑚=0 (𝑥𝑚+1 − 𝑥𝑚 )𝛿𝜈𝑚 𝜈

6.2.1 Average front temperature

The average front temperature is now found by taking the average of the
Gibbs-Thomson equation along the front, separately for each phase (𝛼, 𝛽
and 𝛾):

Δ𝑇𝜈 = 𝑇𝐸 − 𝑇𝜈 = −𝑚𝜈𝐵 (⟨𝑐𝐵 ⟩𝜈 − 𝑐𝐸 𝜈 𝐸


𝐵 ) − 𝑚𝐶 (⟨𝑐𝐶 ⟩𝜈 − 𝑐𝐶 ) + Γ𝜈 ⟨𝜅⟩𝜈 ,
(6.6)

for 𝜈 = 𝛼, 𝛽, 𝛾. Here, ⟨𝜅⟩𝜈 is the average curvature of the solid-liquid


interface which can be evaluated by exact geometric relations to be
∑︀𝑀 −1
𝑚=0 ⟨𝜅⟩𝑚 (𝑥𝑚+1 − 𝑥𝑚 )𝛿𝜈𝑚 𝜈
⟨𝜅⟩𝜈 = ∑︀𝑀 −1
𝑚=0 (𝑥𝑚+1 − 𝑥𝑚 )𝛿𝜈𝑚 𝜈

and
sin 𝜃𝜈𝑚 𝜈𝑚+1 + sin 𝜃𝜈𝑚 𝜈𝑚−1
⟨𝜅⟩𝑚 = .
(𝑥𝑚+1 − 𝑥𝑚 )𝜆
78 6 Multi-phase solidification

Here, 𝜃𝜈𝑚 𝜈𝑚−1 are the contact angles that are obtained by applying Young’s
law at the trijunction points. More precisely, 𝜃𝜈𝑚 𝜈𝑚+1 is the angle, at the
triple point (identified by the intersection of the two solid-liquid interfaces
and the solid-solid one), between the tangent to the 𝜈𝑚 −𝑙 interface and the
horizontal (the 𝑥 direction). For a triple point with the phases 𝜈𝑚 , 𝜈𝑚+1
and liquid, the two contact angles 𝜃𝜈𝑚 𝜈𝑚+1 , 𝜃𝜈𝑚+1 𝜈𝑚 satisfy the following
relations, obtained from Young’s law,
𝜎
˜𝜈𝑚+1 𝑙 𝜎
˜ 𝜈𝑚 𝑙 𝜎˜𝜈𝑚 𝜈𝑚+1
= = .
cos(𝜃𝜈𝑚 𝜈𝑚+1 ) cos(𝜃𝜈𝑚+1 𝜈𝑚 ) sin(𝜃𝜈𝑚 𝜈𝑚+1 + 𝜃𝜈𝑚+1 𝜈𝑚 )

Note that, in general, 𝜃𝜈𝑚 𝜈𝑚+1 ̸= 𝜃𝜈𝑚+1 𝜈𝑚 .

A short digression is in order to motivate the closure of our system of


equations. Although we have not given the explicit expressions, the
coefficients 𝐴0 and 𝐵0 can be simply calculated by using Eq.(6.3) with
𝑛 = 0. However, to carry out this calculation, the width of each lamella has
to be given. If these widths are chosen consistent with the lever rule, that is,
the cumulated lamellar width of phase 𝜈 corresponds to the nominal volume
fraction of phase 𝜈 for the given sample concentration 𝑐∞ ∞ ∞
𝐴 , 𝑐𝐵 , and 𝑐𝐶 , the

use of Eq.(6.3) yields 𝑋0 = 𝑐𝐸𝑋 − 𝑐𝑋 (𝑋 = 𝐴, 𝐵, 𝐶). However, this result is
incorrect: the concentrations of the solids are not equal to the equilibrium
concentrations at the eutectic temperature because solidification takes
place at a temperature below 𝑇E . Therefore, the true volume fractions
depend on the solidification conditions. Their determination would require
a self-consistent calculation which is exceedingly difficult. Therefore, we
will take the same path as Jackson and Hunt in their original paper we
will assume that the volume fractions of the three phases are fixed by the
lever rule at the eutectic temperature, but we will treat the amplitudes of
the two boundary layers, 𝐴0 and 𝐵0 , as unknowns. As in, one can expect
that the difference to the true solution is of order Pe and therefore small
for slow solidification.

With this assumption, the equations developed above can now be used in
two ways. For isothermal solidification, the temperatures of all interfaces
must be equal to the externally set temperature, and the three equations
Δ𝑇𝜈 = Δ𝑇 for 𝜈 = 𝛼, 𝛽, 𝛾, can be used to determine the three unknowns
𝐴0 , 𝐵0 and the velocity 𝑣 of the solid-liquid front. All of these quantities
will be a function of the lamellar spacing 𝜆. In directional solidification,
6.2 Jackson-Hunt analysis 79

the growth velocity in steady state is fixed and equal to the speed with
which the sample is pulled from a hot to a cold region. The third unknown
is now the total front undercooling. In the classic Jackson-Hunt theory for
binary eutectics, the system of equations is closed by the hypothesis that
the average undercoolings of the two phases are equal. This is only an
approximation which is quite accurate for eutectics with comparable volume
fractions of the two solids, but becomes increasingly inaccurate when the
volume fractions are asymmetric. We will use the same approximation for
the ternary case here, and set Δ𝑇𝛼 = Δ𝑇𝛽 = Δ𝑇𝛾 = Δ𝑇 . This then leads
to expressions for Δ𝑇 as a function of the growth speed 𝑣 and the lamellar
spacing 𝜆.
Setting 𝑥0 = 0, 𝑥1 = 𝜂𝛼 , 𝑥2 = 1, and applying Eq. (6.5) gives

1 ∑︁ {︁ 16 (︁
𝛽
)︁ }︁
⟨𝑐𝑋 ⟩𝛼 = 𝑐∞
𝑋 + 𝑋0 + 2 2
Δ𝑐𝛼 2
𝑋 − Δ𝑐𝑋 sin (𝜋𝑛𝜂𝛼(6.7)
)
𝜂𝛼 𝑛=1 𝜆 𝑘𝑛 ℓ𝑞𝑛

∼ 2𝜆
= 𝑐∞
𝑋 + 𝑋0 + 𝒫(𝜂𝛼 )Δ𝑐𝑋 and (6.8)
𝜂𝛼 ℓ
2𝜆
⟨𝑐𝑋 ⟩𝛽 = 𝑐∞
𝑋 + 𝑋0 − 𝒫(1 − 𝜂𝛼 )Δ𝑐𝑋 (6.9)
(1 − 𝜂𝛼 )ℓ
𝛽
with 𝑘𝑛 = 2𝜋𝑛/𝜆, 𝑞𝑛 ≈ 𝑘𝑛 , 𝜆/ℓ ≪ 1, Δ𝑐𝑋 = Δ𝑐𝛼
𝑋 − Δ𝑐𝑋 , and the dimen-
sionless function

∑︁ 1
𝒫(𝜂) = 3
sin2 (𝜋𝑛𝜂) (6.10)
𝑛=1
(𝜋𝑛)

which has the properties 𝒫(𝜂) = 𝒫(1 − 𝜂) = 𝒫(𝜂 − 1).


Furthermore, Eq. (6.6) together with ℓ = 2𝐷/𝑣 leads to
𝜆𝑣
Δ𝑇𝛼 = −𝑚𝛼
𝐵 𝐵0 − 𝒫(𝜂𝛼 )𝑚𝛼𝐵 Δ𝑐𝐵 + Γ𝛼 ⟨𝜅⟩𝛼 ,
𝜂𝛼 𝐷
𝜆𝑣
Δ𝑇𝛽 = −𝑚𝛽𝐴 𝐴0 − 𝒫(𝜂𝛽 )𝑚𝛽𝐴 Δ𝑐𝐴 + Γ𝛽 ⟨𝜅⟩𝛽 ,
𝜂𝛽 𝐷

where ⟨𝜅⟩𝛼 = 2 sin 𝜃𝛼𝛽 /(𝜂𝛼 𝜆) and ⟨𝜅⟩𝛽 = 2 sin 𝜃𝛽𝛼 /(𝜂𝛽 𝜆). In addition, for
a binary alloy 𝐵0 = −𝐴0 . The unknown 𝐴0 and the global front undercool-
ing are determined using the assumption of equal interface undercoolings,
80 6 Multi-phase solidification

Δ𝑇𝛼 = Δ𝑇𝛽 . The result is identical to the one of the Jackson-Hunt


analysis.

The undercooling can be expressed as a function which can be written


as,

(︂ )︂
Δ𝑇𝑚𝑖𝑛 𝜆𝑚𝑖𝑛 𝜆
Δ𝑇 = + , (6.11)
2 𝜆 𝜆𝑚𝑖𝑛

where Δ𝑇𝑚𝑖𝑛 and 𝜆𝑚𝑖𝑛 are the minimum undercooling and the minimum
undercooling spacing respectively. An illustrative graph for the variation
of the undercooling as a function of spacing for a fixed velocity is drawn
in figure 6.2.1.

Figure 6.1: Undercooling vs spacing plotted for a fixed velocity of the


growth front.

Eutectics exhibit a wide-range of co-existing patterns ranging from simple


lamellae, to rods, and other complex co-existing patterns. Some examples
of co-existing patterns can be seen in the following figure 6.2.1.

There exist simulation methods for modeling multi-phase eutectic solidifi-


cation, which are shown exemplarily in the following images as,
6.2 Jackson-Hunt analysis 81

(a) (b) (c)

Figure 6.2: Experimental images of two and three-phase growth during


directional solidification.

(a) (b)

Figure 6.3: Simulated microstructures of two and three-phase coupled


growth.
82 6 Multi-phase solidification

6.3 Long-wavelength perturbation theory

Consider a large array of lamellae consisting of phases 𝛼, 𝛽, with non-


uniform spacings in a directional solidification setup. The variation in the
spacings is on a scale which is larger than that of the lamellar spacing, and
thus will be referred to as long wavelength perturbations. In the following,
we will elaborate on the evolution of the spacings as a function of time,
as in the analysis by Langer. Critical to this analysis, is the assumption
that the growth of lamellae is normal to the local front envelope, which
is Cahn’s hypothesis. Let the displacement of the triple-points be given
by the co-ordinates (𝜉, 𝑦), where 𝜉 is the displacement in the growth
direction, with respect to the unperturbed front and 𝑦 is the displacement
in the transverse direction, see figure 6.4, With this notation, the local

Figure 6.4: Schematic of a long-wavelength pertubation of the growth


front.

strain, in the growth envelope can be written approximately by relating


the deviation in the local lamellar spacing𝜆 with respect to the initial
unperturbed setting 𝜆0 , with the variation of the displacement of the triple
points in the transverse directions as,
6.3 Long-wavelength perturbation theory 83

𝜆 − 𝜆0 𝜕𝑦
= or (6.12)
𝜆0 𝜕𝑥
(︂ )︂
𝜕𝑦
𝜆 = 𝜆0 1 + . (6.13)
𝜕𝑥

Secondly, with Cahn’s hypothesis that the velocity is perpendicular to the


local growth envelope gives us,

𝜕𝑦 𝜕𝜉
= −𝑣 , (6.14)
𝜕𝑡 𝜕𝑥

where we have used the small angle approximation for writing sin 𝜃 = tan 𝜃.
Additionally, the position of the growth front is related to the undercooling
as,

Δ𝑇 = (𝑇𝐸 − 𝑇0 ) − 𝐺𝜉, (6.15)

where 𝑇0 is the temperature of the original unperturbed front. Differen-


tiating, Eqn.6.14 once with respect to 𝑥 and Eqn.6.15 twice w.r.t x, we
have,

𝜕 𝜕𝑦 𝜕2𝜉 𝑣 𝜕 2 Δ𝑇
= −𝑣 2 = (6.16)
𝜕𝑡 𝜕𝑥 𝜕𝑥 𝐺 𝜕𝑥2

Using Eqn.6.13, we can then derive and evolution equation for the spacing
which writes as,

(︃(︂ )︃
𝜕 2 Δ𝑇
(︂ )︂ (︂ )︂ (︂ )︂ (︂ )︂ )︂
𝜕𝜆 𝜆0 𝑣 𝜆0 𝑣 𝜕 𝜕Δ𝑇 𝜆0 𝑣 𝜕 𝜕Δ𝑇 𝜕𝜆
= = = .
𝜕𝑡 𝐺 𝜕𝑥2 𝐺 𝜕𝑥 𝜕𝑥 𝐺 𝜕𝑥 𝜕𝜆 𝜆0 𝜕𝑥
(6.17)
84 6 Multi-phase solidification

The preceding equation for the evolution


(︂ of)︂ the spacing, is a diffusion
𝜕Δ𝑇
equation with the diffusivity given by . Clearly, there are stable
𝜕𝜆 𝜆0
solutions for perturbations
(︂ )︂ to the right of the minumum undercooling
𝜕Δ𝑇
spacing where > 0. Thus, perturbations in spacing in this
𝜕𝜆 𝜆0
region decay monotonically in time, through diffusion equation. To the
left of the minimum undercooling spacing, the perturbations are unstable
giving rise to lamella termination and thereby a resultant increase of
spacing towards the minimum undercooling spacing. This is also referred
to as the Eckhaus Instability.

6.4 Instabilities, for large spacings

Apart from the Eckhaus instability that leads to lamellar termination


for spacings larger than the minimum undercooling spacing, we have
oscillatory instabilities.

In addition to oscillatory instabilities there is also an instability which


leads to the tilting of the lamellae. As one can see the lamellar distance
perpendicular to the tilt is reduced with respect to untilted system, by a
factor which is proportional to the tilt angle. The magnitude of the tilt
increases with increase in lamellar distance. One can characterize each of
these instabilities with their inherent symmetries.

In the first of the images figs. 6.5(b), is represented a 1 − 𝜆 − 𝑂 instability,


wherein, the symmetry passing through both growing phases is preserved
as shown in fig. 6.7(a), thereby preserving the original periodicity in the
lamellae pair. Relaxing one of the symmetries planes though one of the
phases, leads to the 2 − 𝜆 − 𝑂 oscillation (see fig. 6.5(c)), which has now a
periodicity which is double the original setting (fig.6.7(b)). When all the
symmetry planes are broken we derive the tilted state (fig. 6.6).

In 3D, there exist other instabilities such as the zig-zag shown in fig. 6.8.
As in the tilt instability, here too the lamellar distance is modified due to
the formation of the zig-zag structure.
6.4 Instabilities, for large spacings 85

(a) (b) (c)

Figure 6.5: Experimental and simulation images of the different oscillatory


modes seen in binary eutectics.

Figure 6.6: The tilt instability in a binary eutectic alloy.


86 6 Multi-phase solidification

(a) (b)

Figure 6.7: Illustration of the different symmetry elements among the


oscillatory modes.

Figure 6.8: Example of a zig-zag instability in 3D bulk eutectics


6.5 Peritectic growth 87

6.5 Peritectic growth

𝑔(𝑐)

(𝑇 = 𝑇𝑝 )

𝑐
(𝐶𝑝 )

𝑔(𝑐)

(𝑇 < 𝑇𝑝 )

(𝐶𝛼𝑙 ) 𝑐
(𝐶𝛽𝑙 )

The mechanism of growth of the peritectic phase below the peritectic


temperature is different from that of eutectic growth. Here, below the
peritectic temperature the compositions of the liquid in equilibrium with
the peritectic phase 𝛼 is higher as compared to the liquid ahead of the
pro-peritectic phase 𝛽 (see fig. 6.5). Thereby, when placed in contact,
diffusion occurs from the melt ahead of the 𝛼 − 𝑙 interface towards the 𝛽 − 𝑙
88 6 Multi-phase solidification

Figure 6.9: Growth mechanism of the peritectic phase. The solute flux
from the liquid ahead of the peritectic phase, towards the liquid ahead of
the pro-peritectic phase allows it to grow, while the pro-peritectic phase
melts just ahead of the peritectic front.

interface, leading to the growth of the peritectic phase and the consequent
melting of the pro-peritectic phase. The reverse is true for temperatures
above the transformation temperature 𝑇𝑝 , where the pro-peritectic phase 𝛽
would grow at the expense of the 𝛼− phase. The mechanism is illustrated
in the following images, fig.6.9
However, there are regimes of compositions, which are off-peritectic, where
the growth of the both phases can occur in a coupled manner. The
mechanism of such growth patterns is still under research.
The common morphologies that are seen in peritectic systems are displayed
in the following images, fig.6.10,
6.5 Peritectic growth 89

(a) (b)

Figure 6.10: The engulfing microstructure is normally seen in peritectics,


wherein, the peritectic phase engulfs the pro-peritectic phase during phase
transition.

You might also like