You are on page 1of 102

PHASE TRANSFORMATIONS

Nucleation & Growth


TTT and CCT Diagrams
APPLICATIONS
Transformations in Steel
Precipitation
Solidification & crystallization
Glass transition
Recovery, Recrystallization & Grain growth

Phase Transformations in Metals and Alloys


David Porter & Kenneth Esterling
Van Nostrand Reinhold Co. Ltd., New York (1981)

Phase Transformations: an overview


When one phase transforms to another phase it is called phase transformation.
Often the word phase transition is used to describe transformations where there is no
change in composition.
In a phase transformation we could be concerned about phases defined based on:
Structure e.g. cubic to tetragonal phase
Property e.g. ferromagnetic to paramagnetic phase

Phase transformations could be classified based on (pictorial view in next page):


Kinetic: Mass transport Diffusional or Diffusionless
Thermodynamic: Order (of the transformation) 1st order, 2nd order, higher order.
Often subtler aspects are considered under the preview of transformations.
E.g. (i) roughening transition of surfaces, (ii) coherent to semi-coherent transition of
interfaces.

PHASE TRANSFORMATIONS
Based on
Mass
transport

Diffusional
Involves long range mass transport

Diffusionless
E.g. Martensitic

PHASE TRANSFORMATIONS
Based on
order

1nd order
nucleation & growth

2nd (& higher) order


Entire volume transforms

Transformations in Materials
Phase transformations are associated with change in one or more properties.
Hence for microstructure dependent properties we would like to additionally worry about
subtler transformations, which involve defect structure and stress state (apart from
phases).
Therefore the broader subject of interest is Microstructural Transformations.

Phases

Defects

Phases can transform

Defect structures can change

Stress state can be altered

Phase
Transformation

Defect Structure
Transformation

Stress-State
Transformation

Geometrical

Physical

Structural

Property

Phases

Phases Transformations

Residual stress

Microstructure

Microstructural Transformations

Some of the questions we would like to have an answer for


What is a Phase?
What kind of phases exist?
What constitutes a transformation?
How can we cause a phase transformation to occur?
The stimuli: P, T, Magnetic field, Electric field etc.
What kind of phase transformations are there?
Why does a phase transformation occur?
Energy considerations of the system?
Thermodynamic potentials (G, A)

Answers for some these questions may be found in other chapters

Revise concepts of surface and interface energy before starting on these topics

When a volume of material (V) transforms three energies have to be considered :


(i) reduction in G (assume we are working at constant T & P),
(ii) increase in (interface free-energy),
(iii) increase in strain energy.
In a liquid to solid phase transformation the strain energy term can be neglected (as the
liquid can flow and accommodate the volume/shape change involved in the transformationassume we are working at constant T & P).

Bulk Gibbs free energy

New interface created


Energies involved

Interfacial energy

Strain energy
Important in solid to solid transformations

Volume of transformed material

The origin of the strain energy can be understood using the schematics as below. Eshelby
construction is used for this purpose.
In general a solid state phase transformation can involve a change in both volume and
shape. I.e. both dilatational and shear strains may be involved. For simplicity we consider
only change in volume of the material, leading to an increase in the strain energy of the
system (in future considerations).
Bulk Gibbs free energy
Interfacial energy

Energies involved

Strain energy

(a)
(b)

Only volume change

(c)

Considering only
volume change

(d)

Schematic of the Eshelby construction to understand the origin of the stresses due to phase transformation
of a volume (V): (a) region V before transformation, (b) the region V is cut out of the matrix and allowed to
transform (the transformation could involve both shape and volume changes), (c) the transformed volume
(V- shown to be larger in the figure) is inserted into the hole (here only volume change is shown for
simplicity), (c) the system is allowed to equilibrate. The continuity of the system is maintained during the
transformation. The system is strained as a larger volume V is inserted into the hole of volume V.

Let us start understanding phase transformations using the example of the solidification of
a pure metal. (This process is a first order transformation*. First order transformations
involve nucleation and growth**).
There is no change in composition involved as we are considering a pure metal. If we
solidify an alloy this will involve long range diffusion.
Strain energy term can be neglected as the liquid melt can flow to accommodate the
volume change (assume we are working at constant T & P).
The process can start only below the melting point of the liquid (as only below the melting
point the GLiquid < GSolid). I.e. we need to Undercool the system. As we shall note, under
suitable conditions (e.g. container-less solidification in zero gravity conditions), melts can
be undercooled to a large extent without solidification taking place.
Bulk Gibbs free energy
Energies involved

Interfacial energy
Strain energy

1nd order
**
nucleation & growth

Trasformation

Solid-solid transformation

Nucleation
+
of
phase

Growth till
is
exhausted

Click here to know more about order of a phase transformation

Caution: here we are seeing an


increase time experiment and
soon we will be talking of
increasing undercooling
experiments

Liquid
Solid
Growth of Crystal

Two crystal going to join to


form grain boundary

Grain boundary

Crude schematic!

Liquid Solid phase transformation: Solidification

Video snap shots of


solidification of stearic acid

Growth of nucleated crystal

See video here

Solidification complete

For sufficient Undercooling

Liquid Solid phase transformation


On cooling just below Tm solid becomes stable, i.e. GLiquid < GSolid.
But even when we are just below Tm solidification does not start.
E.g. liquid Ni can be undercooled 250 K below Tm.
We will try to understand Why?
The figure below shows G vs T curves for melt and a crystal.
The undercooling is marked as T and the G difference between the liquid and the solid
(which will be released on solidification) is marked as Gv (the subscript indicates that the
quantity G is per unit volume). Hence, Gv is a function of undercooling (T)
Solid stable

Liquid stable

Gv
Solid (GS)

G ve
T

Liquid (GL)
G +ve

T - Undercooling

Tm

As pointed out before solidification is a first order phase transformation involving


nucleation (of crystal from melt) and growth (of crystals such that the entire liquid is
exhausted).
Nucleation is a technical term and we will try to understand that soon.
In solid solid phase transformation, which involve strain energy, heterogeneous
nucleation (defined below) is highly preferred. Even in liquid solid transformations
heterogeneous nucleation plays an very important role.

Solidification

Nucleation
of crystals from melt

Growth
of nucleated crystals till liquid is exhausted

Homogenous
Nucleation
Heterogeneous

Heterogenous nucleation sites


Liquid solid walls of container, inclusions
Solid solid inclusions, grain boundaries,
dislocations, stacking faults

In Homogenous nucleation the probability of nucleation occurring at point in the parent


phase is same throughout the parent phase.
In heterogeneous nucleation there are some preferred sites in the parent phase where
nucleation can occur

Homogenous nucleation

Let us start with a text-book description of nucleation before taking up an alternate


perspective

Let us consider LS transformation taking place by homogenous nucleation. Let the


system be undercooled to a fixed temperature T. Let us consider the formation of a
spherical crystal of radius r from the melt. We can neglect the strain energy contribution.
Let the change in G during the process be G. This is equal to the decrease in bulk free
energy + the increase in surface free energy. This can be computed for a spherical nucleus
as below.

Free energy change on nucleation

Neglected in L S transformations

Reduction in bulk free energy increase in surface energy increase in strain energy

G (Volume).(GV ) (Surface).( )

f (r )

G r 3 .(Gv ) 4r 2 .( )
3

Gv f (T )

Note that GV is negative

r3
r2

r 1

Note that below a value of 1 the lower power of r dominates;


while above 1 the higher power of r dominates.
In the above equation these powers are weighed with other
factors/parameters, but the essential logic remains.

G r 3 .(Gv ) 4r 2 .( )
3

Funda Check A note on minimization versus criticality conditions.

In the above equation, the r3 term is +ve and the r2 term is ve. Such kinds of equations
are often encountered in materials science, where one term is opposing the process and the
other is supporting it. Example of such processes are crack growth (where surface energy
opposes the process and the strain energy stored in the material supports crack growth).
In the current case it is the higher power is supporting the phase transformation. Since the
higher power dominates above 1, the function will go through a maximum as in fig.
below. This implies the G function will go through a maximum. I.e. if the process just
even starts it will lead to an increase in G! (more about this soon).
On the other hand the function with ve contribution from the lower power (to G) will
go through a minimum (fig. below) and such a process will take place down-hill in G and
stop.
1
4

0.6

Goes through a maximum

0.4

2.5

0.2

f(x)

x^n

(x - x^2)
(x^2 - x)

0.8

x^2
x^3

3.5

1.5

-0.2

-0.4
-0.6

0.5

-0.8
0
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

-1

0.2

0.4

0.6

0.8

Goes through a minimum

1.2

1.4

1.6

G r 3 .(Gv ) 4r 2 .( )
3

Note that G is a function of T, r &

As we have noted previously G vs r plot will go through a maximum (implying that as a


small crystal forms G will increase and hence it will tend to dissolve).
The maximum of G vs r plot is obtained by by setting dG/dr = 0. The maximum value of
G corresponds to a value of r called the critical radius (denoted by superscript *).
If by some accident (technically a statistical random fluctuation) a crystal (of preferred
crystal structure) size > r* (called supercritical nuclei) forms then it can grow down-hill in
G. Crystals smaller than r* (called embryos) will tend to shrink to reduce G. The critical
value of G at r* is called G*.
Reduction in G (below the liquid state) is obtained only after r0 is obtained (which can be
obtained by setting G = 0).
dG
dG
0
dr

r*

r2*

Trivial solution

2
Gv

2
Gv

16
G
3 Gv2
3

Note that we are


at a constant T

G 0

3
r0
Gv

As Gv is ve, r*is +ve

r1* 0

dr

G 0

r*
Embryos

r0
Supercritical nuclei

What is the effect of undercooling (T) on r* and G*?


We have noted that GV is a fucntion of undercooling (T). At larger undercoolings GV
increases and hence r* and G* decrease. This is evident from the equations for r* and G*
as below (derived before).
At Tm GV is zero and r* is infinity!

That the melting point is not the same as the freezing point!!
This energy (G) barrier to nucleation is called the nucleation barrier.

Decreasing G*

Tm

Gv f (T )
The bulk free energy reduction is a function of undercooling

r*

2
Gv

16 3
G
3 Gv2
*

Decreasing r*
Using the Turnbull approximation (linearizing the G-T curve
close to Tm), we can get the value of G interms of the enthalpy
of solidification.
Tm2
16 3

G
3
T 2 H 2

Turnbulls approximation

G H f

Solid (GS)

Tm T
T
H f
Tm
Tm
H f heat of fusion

16 3 Tm
*
G
H f T
3

Liquid (GL)
Tm

Jump

Quantum

How are atoms assembled to form a nucleus of r* Statistical Random Fluctuation

Tm
T

r*

Chances of nucleation increases

To cause nucleation (or even to form an embryo) atoms of the liquid (which are randomly moving
about) have to come together in a order, which resembles the crystalline order, at a given instant of
time.
Typically, this crystalline order is very different from the order (local order), which exists in the liquid.
This coming together is a random process, which is statistical in nature i.e. the liquid is exploring
locally many different possible configurations and randomly (by chance), in some location in the
liquid, this order may resemble the preferred crystalline order.
Since this process is random (& statistical) in nature, the probability that a larger sized crystalline
order is assembled is lower than that to assemble a smaller sized crystal.
Hence, at smaller undercoolings (where the value of r* is large) the chance of the formation of a
supercritical nucleus is smaller and so is the probability of solidification (as at least one nucleus is
needed which can grow to cause solidification). At larger undercoolings, where r* value is relatively
smaller, the chance of solidification is higher.

What is meant by the Nucleation Barrier an alternate perspective

Funda Check

Here we try to understand: What exactly is meant by the nucleation barrier?.


It is sometime difficult to fathom out as to the surface energy can make freezing of a small
embryo energetically infeasible (as we have already noted that unless the crystallite size is >
r0 the energy of the system is higher). Agreed that for the surface the energy lowering is not as
much as that for the bulk*, but even the surface (with some unsaturated bonds) is expected to
have a lower energy than the liquid state (where the crystal is energetically favoured). I.e. the
specific concern being: can state-1 in figure below be above the zero level (now considered for
the liquid state)? Is the surface so bad that it even negates the effect of the bulk lowering?
We will approach this mystery from a different angle by first asking the question: what is
meant by melting point? & what is meant by undercooling?.

* refer to surface energy and surface tension slides.

Melting point, undercooling, freezing point (in the realm of homogenous nucleation)
The plot below shows melting point of Au nanoparticles, plotted as a function of the particle radius. It is to
be noted that the melting point of nanoparticles decreases below the bulk melting point (a 5nm particle
melts more than 100C below Tmbulk). This is due to surface effects (surface is expected to have a lower
melting point than bulk!?*) actually, the current understanding is that the whole nanoparticle melts
simultaneously (not surface layer by layer).
Let us continue to use the example of Au. Suppose we are below Tmbulk (1337K=1064C, i.e. system is undercooled
w.r.t the bulk melting point) at T1 (=1300K T = 37K) and suppose a small crystal of r2 = 5nm forms in the liquid.
Now the melting point of this crystal is ~1200K this crystal will melt-away. Now we have to assemble a
crystal of size of about 15nm (= r1) for it not to melt. This needless to say is much less probable (and it is
better to undercool even further so that the value of r* decreases) . Thus the mystery of nucleation barrier
vanishes and we can think of melting point freezing point (for a given size of particle)!
Tm is in heating for the bulk material and in cooling if we take into account the size dependence of melting
point everything sort-of falls into place .

T1

r1

Other materials like Pb, Cu, Bi,


Si show similar trend lines
* Surface atoms are loosely bound as compared to the bulk atoms.

Atomic perspective of nucleation: Nucleation Rate


The process of nucleation (of a crystal from a liquid melt, below Tmbulk) we have described so
far is a dynamic one. Various atomic configurations are being explored in the liquid state some
of which resemble the stable crystalline order. Some of these crystallites are of a critical size
r*T for a given undercooling (T). These crystallites can grow to transform the melt to a
solid by becoming supercritical. Crystallites smaller than r* (embryos) tend to dissolve.
As the whole process is dynamic, we need to describe the process in terms of rate the
nucleation rate [dN/dt number of nucleation events/time].
Also, true nucleation is the rate at which crystallites become supercritical. To find the
nucleation rate we have to find the number of critical sized crystallites (N*) and multiply it by
the frequency/rate at which they become supercritical.
If the total number of particles (which can act like potential nucleation sites in homogenous
nucleation for now) is Nt , then the number of critical sized particles given by an Arrhenius type
function with a activation barrier of G*.

N * Nt e

G *

kT

The number of potential atoms, which can jump to make the critical nucleus supercritical are
the atoms which are adjacent to the liquid let this number be s*.
If the lattice vibration frequency is and the activation barrier for an atom facing the nucleus
(i.e. atom belonging to s*) to jump into the nucleus (to make in supercritical) is Hd , the
frequency with which nuclei become supercritical due atomic jumps into the nucleus is given
H
by:

*
' s e kT
d

Rate of nucleation =

dN
dt

No. of critical sized


particles

N * Nt e

Frequency with which they


become supercritical

G *

kT

No. of particles/volume in L

' s* e

lattice vibration frequency (~1013 /s)

s* atoms of the liquid facing the nucleus

Critical sized nucleus (r*)


Jump taking particle to supercriticality
nucleated (enthalpy of activation = Hd)
Outline of critical sized nucleus

H d

kT

The nucleation rate (I = dN/dt) can be written as a product of the two terms as in the equation
below.
How does the plot of this function look with temperature?
At Tm , G* is I = 0 (as expected if there is no undercooling there is no nucleation).
At T = 0K again I = 0
This implies that the function should reach a maximum between T = Tm and T = 0.
A schematic plot of I(T) (or I(T)) is given in the figure below.
An important point to note is that the nucleation rate is not a monotonic function of
undercooling.

Tm

I Nt s* e

T (K)

Increasing T

T = Tm G* = I = 0

T=0I=0

G * H d

kT

G* I

Note: G* is a function of T

T I

Heterogenous nucleation
We have already talked about the nucleation barrier and the difficulty in the nucleation
process. This is all the more so for fully solid state phase transformations, where the strain
energy term is also involved (which opposes the transformation).
The nucleation process is often made easier by the presence of defects in the system.
In the solidification of a liquid this could be the mold walls.
For solid state transformation suitable nucleation sites are: non-equilibrium defects such
as excess vacancies, dislocations, grain boundaries, stacking faults, inclusions and
surfaces.
One way to visualize the ease of heterogeneous nucleation
heterogeneous nucleation at a defect will lead to destruction/modification of the defect
(make it less defective). This will lead to some free energy Gd being released thus
reducing the activation barrier (equation below).

G hetro,defect (V) Gv Gs A (Gd )


Increasing Gd (i.e. decreasing G*)
Homogenous sites
Vacancies
Dislocations
Stacking Faults
Grain boundaries (triple junction), Interphase boundaries
Free Surface

Heterogenous nucleation
Consider the nucleation of from on a planar surface of inclusion .
The nucleus will have the shape of a lens (as in the figure below).
Surface tension force balance equation can be written as in equation (1) below. The contact angle
can be calculated from this equation (as in equation (3)).
Keeping in view the interface areas created and lost we can write the G equation as below (2).

Interfacial Energies

Created

Alens

Created

Acircle

is the contact angle


Cos


(3)

Surface tension force balance

Acircle

Lost

Cos

(1)

G (Vlens )Gv (Alens ) ( Acircle ) ( Acircle )


Vlens = h2(3r-h)/3

Alens = 2rh

h = (1-Cos)r

(2)

rcircle = r Sin

Using the procedure as before (for the case of the homogenous nucleation) we can find r* for
heterogeneous nucleation. Using the surface tension balance equation we can write the
formulae for r*and G* using a single interfacial energy (and contact angle ).
*
*
Further we can write down Ghetero
in terms of Ghomo
and contact angle .

dG
0
dr
G

*
hetero

*
hetero

2
Gv

*
Ghetero

2 3Cos Cos 3
2
3 Gv

1 *

Ghomo
2 3Cos Cos 3
4

*
Ghetero
1

2 3Cos Cos 3
*
Ghomo
4

Just a function of
the contact angle

Decreasing tendency to wet the substrate

Increasing
contact angle

= 0 f() = 0
*
Ghetero
1

2 3Cos Cos3 f ( )
*
Ghomo
4

= 90 f() =

Complete wetting

Partial wetting

= 180 f() = 1
No wetting
*
*
The plot of Ghetero
/ Ghomo
is shown in the next page.

Plot of G*hetero/G*homo is shown below. This brings out the benefit of heterogeneous nucleation vs homogenous nucleation.
If the phase nucleus (lens shaped) completely wets the substrate/inclusion (-phase) (i.e. = 0)
then G*hetero = 0 there is no barrier to nucleation.

On the other extreme if -phase does not we the substrate (i.e. = 180)
then G*hetero = G*homo there is no benefit of the substrate.
In reality the wetting angle is somewhere between 0-180
Hence, we have to chose a heterogeneous nucleating agent with a minimum value.

G*hetero (180o) = G*homo


no benefit

G*hetero / G*homo

0.75

G*hetero (0o) = 0
no barrier to nucleation

G*hetero (90o) = G*homo/2

0.5


Cos

0.25
Complete wetting

No wetting

Partial wetting

0
0

30

60

90
120
(degrees)

150

180

Choice of heterogeneous nucleating agent


Heterogeneous nucleation has many practical applications.
During the solidification of a melt if only a few nuclei form and
these nuclei grow, we will have a coarse grained material (which
will have a lower strength as compared to a fine grained
material- due to Hall-Petch effect).
Hence, nucleating agents are added to the melt (e.g. Ti for Al
alloys, Zr for Mg alloys) for grain refinement.

Cos

How to get a small value of ? (so that easy heterogeneous nucleation).


Choosing a nucleating agent with a low value of (low energy interface)
(Actually the value of ( ) will determine the effectiveness of the heterogeneous
nucleating agent high or low )
How to get a low value of ?
We can get a low value of if:
(i) crystal structure of and are similar and
(ii) lattice parameters are as close as possible
Examples of such choices:
In seeding rain-bearing clouds AgI or NaCl are used for nucleation of ice crystals
Ni (FCC, a = 3.52 ) is used a heterogeneous nucleating agent in the production of
artificial diamonds (FCC, a = 3.57 ) from graphite.

Why does heterogeneous nucleation dominate? (arent there more number of homogenous nucleation sites?)
To understand the above questions, let us write the nucleation rate for both cases as a preexponential term and an exponential term. The pre-exponential term is a function of the
number of nucleation sites.
However, the term that dominates is the exponential term and due to a lower G* the
heterogeneous nucleation rate is typically higher.

0
I homo I homo
e

*
Ghomo

kT

0
I hetero I hetero
e

*
Ghetero

kT

= f(number of nucleation sites)


~ 1026

= f(number of nucleation sites)


~ 1042

BUT
the exponential term dominates
Ihetero > Ihomo

Heterogeneous nucleation in AlMgZn alloy

Growth
Diffusional transformations involve nucleation and growth. Nucleation involves the
formation of a different phase from a parent phase (e.g. crystal from melt). Growth
involves attachment of atoms belonging to the matrix to the new phase (e.g. atoms
belonging to the liquid phase attach to the crystal phase).
Nucleation we have noted is uphill in G process, while growth is downhill in G.
Growth can proceed till all the prescribed product phase forms (by consuming the parent
phase).

Transformation

= Nucleation of phase +

Growth of phase
till is exhausted*

Growth
At transformation temperature the probability of jump of atom from (across the
interface) is same as the reverse jump
Growth proceeds below the transformation temperature, wherein the activation barrier for the
reverse jump is higher than that for the forward jump.
Hd

Hd vatom Gv

phase

phase

Transformation rate
As expected transformation rate (Tr) is a function of nucleation rate (I) and growth rate
(U).
In a transformation, if X is the fraction of -phase formed, then dX/dt is the
transformation rate.
The derivation of Tr as a function of I & U is carried using some assumptions (e.g.
Johnson-Mehl and Avarami models).

We have already seen the curve for the nucleation rate (I) as a function of the
undercooling.
The growth rate (U) curve as a function of undercooling looks similar. The key difference
being that the maximum of U-T* curve is typically above the I-T curve*.
This fact that T(Umax) > T(Imax) give us an important handle on the scale of the
transformed phases forming. We will see examples of the utility of this information later.

Transformation rate f(Nucleation rate, Growth rate) Tr

Increasing T

Tm

dX
dt

f (I , U )

Maximum of growth rate usually at higher


temperature than maximum of nucleation rate

Tr

T (K)

I, U, Tr

[rate sec1]

* The U-T curve is an alternate way of stating the U-T curve

Fraction of the product () phase forming with time the sigmoidal growth curve
Many processes in nature (etc.), e.g. growth of bacteria in a culture (number of bacteria
with time), marks obtained versus study time(!), etc. tend to follow a universal curve the
sigmoidal growth curve.
In the context of phase transformation, the fraction of the product phase (X) forming with
time follows a sigmoidal curve (function and curve as below).

1.0

X 1 e

0.5

Saturation phase decreasing growth rate


with time

I U3 t 4

Linear growth regime ~constant high


growth rate

Incubation period slow growth (but with


increasing growth rate with time)

From Rate to time: the origin of Time Temperature Transformation (TTT) diagrams
A type of phase diagram

The transformation rate curve (Tr-T plot) has hidden in it the I-T and U-T curves.
An alternate way of plotting the Transformation rate (Tr) curve is to plot Transformation
time (Tt) [i.e. go from frequency domain to time domain]. Such a plot is called the TimeTemperature-Transformation diagram (TTT diagram).
High rates correspond to short times and vice-versa. Zero rate implies time (no transformation).
This Tt-T plot looks like the C alphabet and is often called the C-curve. The minimum
time part is called the nose of the curve.

Tm

Rate f (T , t )

Tm
Small driving
force for nucleation

Tr

Time for transformation

Replot

T (K)

T (K)

Nose of the C-curve

Tr (rate sec1)

Tt
Sluggish growth

Tt (time sec)

Understanding the TTT diagram


Though we are labeling the transformation temperature Tm , it represents other transformations, in
addition to melting.
Clearly the Tt function is not monotonic in undercooling. At Tm it takes infinite time for
transformation.
Till T3 the time for transformation decreases (with undercooling) [i.e. T3 < T2 < T1] due to
small driving force for nucleation.
After T3 (the minimum) the time for transformation increases [i.e. T3 < T4 < T5] due to sluggish
growth.
This is a phase diagram where the blue region is the Liquid (parent) phase field and purplish region is
the transformed product (crystalline solid).
The diagram is called the TTT diagram because it plots the
time required for transformation if we hold the sample at
fixed temperature (say T1) or fixed undercooling (T1). The
time taken at T1 is t1.
To plot these diagrams we have to isothermally hold at
various undercoolings and note the transformation time.
I.e. instantaneous quench followed by isothermal hold.

Hence, these diagrams are also called Isothermal


Transformation Diagrams.
Similar curves can be drawn for (solid state)
transformation.

Clearly the picture of TTT diagram presented before is incomplete transformations may
start at a particular time, but will take time to be completed (i.e. between the L-phase field
and solid phase field there must be a two phase region L+S!).
This implies that we need two C curves one for start of transformation and one for
completion. A practical problem in this regard is related to the issue of how to define start
and finish (is start the first nucleus which forms? Does finish correspond to 100%?) . Since practically it is
difficult to find % and 100%, we use practical measures of start and finish, which can
be measured experimentally. Typically this is done using optical metallography and a
reliable resolution of the technique is about 1% for start and 99% for finish.
Another obvious point: as x-axis is time any transformation paths have to be drawn such
that it is from left to right (i.e. in increasing time).
TTT diagram phase transformation

How do we define the fractions transformed?

Increasing % transformation

99% = finish

T (K)

1% = start

Fraction
transformed

volume fraction of at t
final volume of


f volume fractionof

t (sec)

How can we compute Tt(T) (transformation time for each T)


The C curve depends on various factors as listed in diagram below.
Some common assumptions used in the derivation are: (i) constant number of nuclei, (ii)
constant nucleation rate, (iii) constant growth rate.

Nucleation rate
Growth rate
f(t,T) determined by

Density and distribution of nucleation sites


Overlap of diffusion fields from adjacent transformed volumes
Impingement of transformed volumes

Constant number of nuclei (these form at the beginning of the transformation)


One assumption to simplify the derivation is to assume that the number of nucleation sites
remain constant and these form at the beginning of the transformation.
This situation may be approximately valid for example if a nucleating agent (inoculant) is
added to a melt (the number of inoculant particles remain constant).
In this case the transformation rate is a function of the number of nucleation sites (fixed)
and the growth rate (U).
Growth rate is expected to decrease with time.
In Avrami model the growth rate is assumed to be constant (till impingement).

f F (number of nucleation sites, growth rate)

growth rate withtime

Derivation of f(T,t): Avrami Model


Parent phase has a fixed number of nucleation sites Nn per unit volume (and these sites are
exhausted in a very short period of time
Growth rate (U = dr/dt) constant and isotropic (as spherical particles) till particles impinge
on one another
At time t the particle that nucleated at t = 0 will have a radius r = Ut
Between time t = t and t = t + dt the radius increases by dr = Udt
The corresponding volume increase dV = 4r2 dr
Without impingement, the transformed volume fraction (f) (the extended transformed
volume fraction) of particles that nucleated between t = t and t = t + dt is:

f Nn 4 r 2 dr Nn 4 Ut Udt Nn 4U 3t 2 dt
2

This fraction (f) has to be corrected for impingement. The corrected transformed volume
fraction (X) is lower than f by a factor (1X) as contribution to transformed volume
fraction comes from untransformed regions only:

dX
1 X

dX
N n 4U 3t 2 dt

1 X

t t

dX
3 2

N
4

U
t dt
0 1 X t 0 n

X 1 e

4 Nn U3 t 3

Based on the assumptions note that the growth rate is not part of the equation it is only the
number of nuclei.

Where do we see constant growth rate?


In cellular transformations constant growth rate is observed.
Termination of transformation does not occur by a gradual reduction in the growth rate but
by the impingement of the adjacent cells growing with a constant velocity.
E.g.: Pearlitic transformation, Cellular precipitation, Massive transformation,
recrystallization.

Cellular Transformations Constant growth rate


All of the parent phase is consumed by the product phase

Pearlitic transformation
Cellular Precipitation

Massive Transformation
Recrystallization

Constant nucleation rate


Another common assumption is that the nucleation rate (I) is constant.
In this case the transformation rate is a function of both the nucleation rate (fixed) and the
growth rate (U).
Growth rate decreases with time.
If we further assume that the growth rate is constant (till impingement), then we get the
Johnson-Mehl model.

f F (nucleation rate, growth rate)

growth rate withtime

Derivation of f(T,t): Johnson-Mehl Model


Parent phase completely transforms to product phase ( )
Homogenous Nucleation rate of in untransformed volume is constant (I)
Growth rate (U = dr/dt) constant and isotropic (as spherical particles) till particles impinge
on one another
At time t the particle that nucleated at t = 0 will have a radius r = Ut
The particle which nucleated at t = will have a radius r = U(t )
Number of nuclei formed between time t = and t = + d Id
Without impingement, the transformed volume fraction (f) (called the extended
transformed volume fraction) of particles that nucleated between t = and t = + d is:

4
4
3
f r 3 Id U (t ) Id
3
3
This fraction (f) has to be corrected for impingement. The corrected transformed volume
fraction (X) is lower than f by a factor (1X) as contribution to transformed volume
fraction comes from untransformed regions only:

dX
1 X

dX
4 3
4
3
r Id U (t ) Id

1 X 3
3

dX
4
3

U
(
t

)
Id X 1 e

0 1 X 0 3

I U3 t 4

Note that X is both a function of I and


U. I & U are assumed constant

1.0

For a isothermal transformation

I U3

is a constant during isothermal transformation


3
0.5

APPLICATIONS
of the concepts of nucleation & growth
TTT/CCT diagrams

Phase Transformations in Steel


Precipitation
Solidification, Crystallization and Glass Transition
As hyperlinks

Recovery recrystallization & grain growth

Phase Transformations in Steel


Now we have the necessary wherewithal to understand phase transformations in steel
Phase diagram (Fe-Fe3C) and Concept of TTT diagrams
We shall specifically consider TTT and CCT diagrams for eutectoid, hypo- and hypereutectoid steels.
Further we will consider the use of these diagrams to design heat treatments to get a
specific microstructure (each microstructure will give us a different set of properties).

We have already seen the Fe-Fe3C phase diagram (please have a second look!)
Fe-Cementite diagram
Peritectic
L+

Eutectic
L + Fe3C

1493C

L+

0.1 %C
Eutectoid
+ Fe3C

1147C

2.06
+ Fe3C

723C

+ Fe3C

0.025 %C

%C

Fe
0.16

0.8

4.3

Fe3C
6.7

For every composition of steel we should draw a different TTT diagram.


To the left of the start C curve is the Austenite () phase field.
To the right of finish C curve is the ( + Fe3C) phase field.
TTT diagram for
Eutectoid steel (0.8%C)

Above Eutectoid
temperature there is no
transformation

800
Eutectoid temperature

723C

700
Pearlite

600
+ Fe3C

Nose of C curve

500

Nose of the C curve is in


~sec and just below TE
transformation times may be
~day.

The starting phase has to


.
The ( + Fe3C) phase
field has more labels
included.
There are horizontal
lines labeled Ms and Mf.

400

Important points to be
noted:
The x-axis is log scale.

Eutectoid steel (0.8%C)

Pearlite + Bainite
Bainite

Austenite

300
200
100

Ms
Mf

Martensite
0.1

10

102

t (s)
103

104

105

As pointed out before one of the important utilities of the TTT diagrams comes from the
overlay of microconstituents (microstructures) on the diagram.
Depending on the T, the ( + Fe3C) phase field is labeled with microconstituents like
Pearlite, Bainite.
We had seen that TTT diagrams are drawn by instantaneous quench to a temperature
followed by isothermal hold.
Suppose we quench below (~225C, below the temperature marked Ms), then Austenite
transforms via a diffusionless transformation (involving shear) to a (hard) phase known as
Martensite. Below a temperature marked Mf this transformation to Martensite is complete.
Once is exhausted it cannot transform to ( + Fe3C).
Hence, we have a new phase field for Martensite. The fraction of Martensite formed is not
a function of the time of hold, but the temperature to which we quench (between Ms and
Mf).
800
Eutectoid steel (0.8%C)
Eutectoid temperature

723C

700
Pearlite

600
+ Fe3C

500

400

Pearlite + Bainite
Bainite

Austenite

300
200
100

Ms
Mf
Martensite

0.1

10

102

t (s)
103

104

105

Strictly speaking cooling curves (including finite quenching rates) should not be overlaid
on TTT diagrams (remember that TTT diagrams are drawn for isothermal holds!).
Isothermal hold at: (i) T1 gives us Pearlite, (ii) T2 gives Pearlite+Bainite, (iii) T3 gives
Bainite. Note that Pearlite and Bainite are both +Fe3C (but their morphologies are
different).
To produce Martensite we should quench at a rate such as to avoid the nose of the start C
curve. Called the critical cooling rate.
If we quench between Ms and Mf we
will get a mixture of Martensite and
(called retained Austenite).

800
723C Eutectoid temperature
700
600
500

Austenite

200
100
T

Not an isothermal
transformation

Coarse
Pearlite

T1

Fine
Pearlite + Bainite

T2

400
300

Eutectoid steel (0.8%C)

Bainite

T3

Ms

Austenite

Mf
Martensite

0.1

10

103
102
t (s)

104

105

Funda Check

For the transformations to both Pearlite and Bainite, why do we have only
one C curve?

In principle two curves exist for Pearlitic and Bainitic transformations


they are usually not resolved in plain C steel (In alloy steels they can be distinct).
Eutectoid steel (0.8%C)

TTT Diagram: hypoeutectoid steel

C1

Atlas of Isothermal Transformation and Cooling Transformation


Diagrams, ASM International, Metals Park, OH, 1977.

In hypo- (and hyper-) eutectoid steels (say composition C1) there is one more branch to the
C curve-NP (marked in red).
The part of the curve lying between T1 and TE (marked in figs. below) is clear, because in this range of
temperatures we expect only pro-eutectoid to form and the final microstructure will
consist of and .(E.g. if we cool to Tx and hold- left figure).
The part of the curve below TE is a bit of a mystery (since we are instantaneously cooling
steel
to below TE, we should get a mix of + Fe3C what is the meaning Hypo-Eutectoid
of a pro-eutectoid
phase in a TTT diagram? (remember pro- implies pre-).(Considered next)

Funda Check

Why do we get pro-eutectoid phase below TE?

Suppose we quench instantaneously an hypo-eutectoid composition (C1) to Tx we should expect the


formation of +Fe3C (and not pro-eutectoid first).
The reason we see the formation of pro-eutectoid first is that the undercooling w.r.t to Acm is more
than the undercooling w.r.t to A1. Hence, there is a higher propensity for the formation of pro-eutectoid
.

Undercooling wrt Acm


(formation of pro-eutectoid )

undercooling wrt A1 line


(formation of + Fe3C)

C1

Similar to the hypo-eutectoid case, hyper-eutectoid compositions (e.g. C2 in fig. below) have a
+Fe3C branch.
For a temperature between T2 and TE (say Tm (not melting point- just a label)) we land up with +Fe3C.
For a temperature below TE (but above the nose of the C curve) (say Tn), first we have the
formation of pro-eutectoid Fe3C followed by the formation of eutectoid +Fe3C.

T2
TE

Hyper-Eutectoid steel

C2

Continuous Cooling Transformation (CCT) Curves


The TTT diagrams are also called Isothermal Transformation Diagrams, because the
transformation times are representative of isothermal hold treatment (following a instantaneous quench).
In practical situations we follow heat treatments (T-t procedures/cycles) in which (typically)
there are steps involving cooling of the sample. The cooling rate may or may not be
constant. The rate of cooling may be slow (as in a furnace which has been switch off) or
rapid (like quenching in water).
Hence, in terms of practical utility TTT curves have a limitation and we need to draw
separate diagrams called Continuous Cooling Transformation diagrams (CCT), wherein
transformation times (also: products & microstructure) are noted using constant rate cooling
treatments. A diagram drawn for a given cooling rate (dT/dt) is typically used for a range of
cooling rates (thus avoiding the need for a separate diagram for every cooling rate).
However, often TTT diagrams are also used for constant cooling rate experiments keeping
in view the assumptions & approximations involved.
The CCT diagram for eutectoid steel is considered next. Blue curve is the CCT curve and
TTT curve is overlaid for comparison.
Important difference between the CCT & TTT transformations is that in the CCT case
Bainite cannot form.

Continuous Cooling Transformation (CCT) Curves

T1

dT
T
dt

> T2

800

Eutectoid temperature

723

Eutectoid steel (0.8%C)

Start

600

Pearlite

Finish

500

Original TTT lines

400

Important points to be
noted:
As before the x-axis is
log scale.
Bainite cannot form by
continuous cooling.
Constant rate cooling
curves look like curves
due to log scale in xaxis. The higher cooling
rate curve has a higher
(negative) slope.
As time is one of the
axes, no treatment curve
can be drawn where time
decreases or remains
constant.

Constant Cooling rate

300
200

Cooling curves
Constant rate

Ms
Mf

100

Martensite
0.1

T2

T1
10

102

103
104
t (s)

105

Funda Check The CCT curves are to the right of the corresponding TTT curves. Why?
As the cooled sample has spent more time at higher temperature, before it intersects the
TTT curve (virtually superimposed) and the transformation time is longer at higher T
(above the nose) CCT curves should be to the right of TTT curves.

800

Eutectoid temperature

723

600

Eutectoid steel (0.8%C)

Pearlite

500

Original TTT lines

400
300
200

Cooling curves
Constant rate

Ms
Mf

100

0.1

T2

T1

Martensite
10

102

103
104
t (s)

105

Common heat treatments involving cooling

Common cooling heat treatment labels (with increasing cooling rate) are:
Full anneal < Normalizing < Oil quench < Water quench
The microstructures produced for these treatments are:
Full Anneal (furnace cooling) Coarse Pearlite
Normalizing (Air cooling) Fine Pearlite
Oil Quench Matensite (M) + Pearlite (P)
Water Quench Matensite
To produce full martensite we have to avoid the nose of the TTT diagram (i.e. the
quenching rate should be fast enough).
Within water or oil quench further parameters determine the actual quench rate (e.g. was the
sample shaken?).

Different cooling treatments

It is important to note that for a single composition, different cooling treatments give
different microstructures these give rise to a varied set of properties.
After even water quench to produce Martensite, further heat treatment (tempering) can be
given to optimize properties like strength and ductility.
800
Eutectoid steel (0.8%C)

723

600

zin

al

nne

ll a

Fu

ali
rm

No

ench

400

Water qu

500

ch
uen

lq
Oi

300
200

Coarse P

100
M
Pearlite
PP == Pearlite
M == Martensite
Martensite
M

0.1

M + P
10

103
102
t (s)

Fine P
104

105

What are the typical cooling rates of various processes?

Process

Cooling rate (K/s)

Furnace cooling (Annealing)

105 103

Air Cooling

1 10

Oil Quenching*

~100

Water Quenching*

~500

Splat Quenching

105

Melt-Spinning

106 108

Evaporation, sputtering

109 (expected)

* Depends on conditions discussed later

Pearlite

+ Fe3C

Lamellae of Pearlite in ~0.8% carbon steel

Nucleation and growth


Heterogeneous nucleation at grain boundaries
Interlamellar spacing is a function of the temperature of transformation
Lower temperature finer spacing higher hardness

Mechanism of Pearlitic transformation: arising of the lamellar microstructure


1 Let us consider the heterogeneous nucleation of one of the phases of the pearlitic
microconstituent (say Fe3C), at a grain boundary of Austenite (). Further let this
precipitate be bound by a coherent interface on one side and a incoherent interface on the
other side. The incoherent interface will be glissile (mobile) and will grow into the
corresponding grain (2).
The orientation relation (OR) between and Fe3C is refered to as the KurdyumovSachs OR (as in fig. below).
2,3 The region surrounding this Fe3C precipitate will be depleted in Carbon and the
conditions will be right for the nucleation of adjacent to it.
4 The process is repeated to give rise to a pearlitic colony. Branching of an advancing
plate may also be observed.

(100)C || (1 11)
Orientation Relation:
Kurdyumov-Sachs

(010)C || (110)

(001)C || (112)

Branching mechanism

Bainite

+ Fe3C**

Micrograph courtesy: Prof. Sandeep Sangal

Bainite formed at high temperature (~ 350C) has a feathery appearance and is called
Feathery Bainite.
Bainite formed at lower temperature (~ 275C) has a needle-like appearance and is called
acicular Bainite.
The process of formation of bainite involves nucleation and growth
Acicular, accompanied by surface distortions
** Lower temperature carbide could be carbide (hexagonal structure, 8.4% C)
Bainite plates have irrational habit planes
Ferrite in Bainite plates possess different orientation relationship relative to the parent
Austenite than does the Ferrite in Pearlite

More images of Bainite

Micrograph courtesy: Prof. Sandeep Sangal, Swati Sharma

AFM image

0.8% C steel, the sample was quenched


in a salt bath having 400C temperature
and then it was held for 2 hours.
Micrograph courtesy: Prof. Sandeep Sangal, Swati Sharma

Characteristic of Martensitic transformations


Shape of the Martensite formed Lenticular (or thin parallel plates)
Associated with shape change (shear)
But: Invariant plane strain (observed experimentally) Interface plane between Martensite and
Parent remains undistorted and unrotated
This condition requires:
1) Bain distortion Expansion or contraction of the lattice along certain crystallographic
directions leading to homogenous pure dilation
2) Secondary Shear Distortion Slip or twinning
3) Rigid Body rotation

Surface deformations caused by the Martensitic plate

Martensite
Change in Crystal Structure

( FCC )
0.8 %C

Quench

' ( BCT )
0.8 %C

Martensitic transformation can be understood by first considering an alternate unit cell for the
Austenite phase as shown in the figure below.
If there is no carbon in the Austenite (as in the schematic below), then the Martensitic
transformation can be understood as a ~20% contraction along the c-axis and a ~12% expansion of
the a-axis accompanied by no volume change and the resultant structure has a BCC lattice (the
usual BCC-Fe) c/a ratio of 1.0.

In Pure Fe after
the Matensitic transformation
c=a

FCC Austenite alternate choice of Cell

~20% contraction of c-axis


~12% expansion of a-axis

FCC BCC

Martensite

In the presence of Carbon in the octahedral voids of CCP (FCC) -Fe (as in the schematic below)
the contraction along the c-axis is impeded by the carbon atoms. (Note that only a fraction of the
octahedral voids are filled with carbon as the percentage of C in Fe is small).
However the a1 and a2 axis can expand freely. This leads to a product with c/a ratio (c/a) >1
1-1.1.
In this case there is an overall increase in volume of ~4.3% (depends on the carbon content) the
Bain distortion*.
C along the c-axis
obstructs the contraction

Austenite to Martensite ~4.3 % volume increase

Tetragonal
Martensite

* Homogenous dilation of the lattice (expansion/contraction along crystallographic axis) leading to the formation of a new lattice is called
Bain distortion. This involves minimum atomic movements.

Martensite in 0.6%C steel

But shear will distort the lattice!


Twinning

Slip

Average shape
remains undistorted

Summary of characteristics of Martensitic transformation


The martensitic transformation occurs without composition change

The transformation occurs by shear without need for diffusion


The atomic movements required are only a fraction of the interatomic spacing
The shear changes the shape of the transforming region
results in considerable amount of shear energy
plate-like shape of Martensite
The amount of martensite formed is a function of the temperature to which the sample is
quenched and not of time
Hardness of martensite is a function of the carbon content
but high hardness steel is very brittle as martensite is brittle
Steel is reheated to increase its ductility
this process is called TEMPERING

Hardness (Rc)

60

40

Harness of Martensite as a
function of Carbon content

20

% Carbon

0.2

0.4

0.6
Properties of 0.8% C steel
Hardness (Rc)

Tensile strength (MN / m2)

Coarse pearlite

16

710

Fine pearlite

30

990

Bainite

45

1470

Martensite

65

Martensite tempered at 250 oC

55

1990

Constituent

ROLE OF ALLOYING ELEMENTS


Interstitial
Segregation / phase separation

Solid solution

Substitutional

Element Added
Compound (new crystal structure)

Plain Carbon Steel

Alloying elements

+ Simplicity of heat treatment and lower cost


Low hardenability
Loss of hardness on tempering
Low corrosion and oxidation resistance
Low strength at high temperatures

hardenability
Provide a fine distribution of alloy carbides during tempering
resistance to softening on tempering
corrosion and oxidation resistance
strength at high temperatures
Strengthen steels that cannot be quenched
Make easier to obtain the properties throughout a larger section
Elastic limit (no increase in toughness)
Alter temperature at which the transformation occurs
Alter solubility of C in or Iron
Alter the rate of various reactions

Sample elements and their role


P Dissolves in ferrite, larger quantities form iron phosphide brittle (cold-shortness)
S Forms iron sulphide, locates at grain boundaries of ferrite and pearlite poor ductility
at forging temperatures (hot-shortness)
Si (0.2-0.4%) increases elastic modulus and UTS
Cu 0.8 % soluble in ferrite, can be used for precipitation hardening
Pb Insoluble in steel
Cr Corrosion resistance, Ferrite stabilizer, hardness/strength, > 11% forms passive
films, carbide former
Ni Austenite stabilizer, strength ductility and toughness,
Mo Dissolves in & , forms carbide, high temperature strength,
temper embrittlement, strength, hardenability

Brinell Hardness

Mn +0.1% C
180

Mn
Cr + 0.1%C
140

Cr

100

60

10

Alloying Element (%)


Alloying elements increase hardenability but the major contribution to hardness comes from
Carbon

Temperature

Mn, Ni are Austenite stabilizers

0.35% Mn

6.5% Mn
0.4

Temperature

15% Cr

0.4

12% Cr

0.8

1.2

C (%)

1.6

Outline of the phase field

0% Cr

5% Cr

0.8

1.2

C (%)

1.6

Cr is Ferrite stabilizer
Shrinking phase field with Cr

TTT diagram for Ni-Cr-Mo low alloy steel


800
Pearlite

Austenite
600

500
400
300
200
100

Bainite
Ms
Mf
Martensite
~1 min

TTT diagram of low alloy steel (0.42%


C, 0.78% Mn, 1.79% Ni, 0.80% Cr,
0.33% Mo)
U.S.S Carilloy Steels, United States Steel Corporation,
Pittsburgh, 1948)

Effect of carbon content and heat treatment on properties of steel


1000
0.4% C - Slow cooled

900

0.8% C - Slow cooled


800
700
600
500

Tensile Test

400
300
200
100
0
0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Engineering Strain (e)

450
Slowly cooled- 0.6%C
Quenched- 0.8% C
Slowly cooled- 0.8% C
Slowly cooled- 1.0% C

400

Hardness

Vikers Hardness

Engineering Stress (s) [MPa]

0.8% C - quenched

350
300
250
200
150
0.5

0.6

0.7

C%

0.8

0.9

1.1

Precipitation

Precipitation Hardening
The presence of dislocation weakens the crystal easy plastic deformation.
Putting hindrance to dislocation motion increases the strength of the crystal.
Fine precipitates dispersed in the matrix provide such an impediment.
Strength of Al 100 MPa
Strength of Duralumin with proper heat treatment (Al + 4% Cu + other alloying elements)
500 MPa

Al-Cu phase diagram: the sloping solvus line and the design of heat treatments

T (C)

600

400

200

Al

Sloping Solvus line


high T high solubility
low T low solubility of Cu in Al

15

30
% Cu

45

60

+
Slow equilibrium cooling gives rise to coarse
precipitates which is not good in impeding
dislocation motion.*

4 % Cu
( FCC )
( FCC ) CuAl 2 (Tetragonal )

slowcool

4
%
Cu

0
.
5
%
Cu

52
%
Cu

550o C
RT

RT

*Also refer section on Double Ended Frank-Read Source in the chapter on plasticity: max

= Gb/L

To obtain a fine distribution of precipitates the cycle A B C is used


Note: Treatments A, B, C are for the same
composition

B
A

4 % Cu
A

Heat (to 550oC) solid solution


supersaturated solution

Quench (to RT)


Increased vacancy concentration

Age (reheat to 200oC) fine precipitates

Hardness

100oC

Schematic curves
Real experimental curves
are in later slides

180oC
20oC

Log(t)
Higher temperature

less time of aging to obtain peak hardness

Lower temperature
increased peak hardness
optimization between time and hardness required
Note: Schematic curves shown- real curves considered later

Hardness

180oC

Peak-aged

Dispersion of
fine precipitates
(closely spaced)

Coarsening
of precipitates
with increased
interparticle spacing
Overaged

Underaged

Log(t)
Region of solid solution
strengthening
(no precipitation hardening)

Region of precipitation
hardening
(but little/some solid solution
strengthening)

Peak-aged

Hardness

180oC

CRSS Increase

Section of GP zone parallel to (200) plane

Log(t)
Particle
shearing

Particle
By-pass

r
r
Particle radius (r)
1
2

r f (t )

Hardness

A complex set of events are happening in parallel/sequentially during the aging process
These are shown schematically in the figure below

Log(t)
Increasing size of precipitates with increasing interparticle (inter-precipitate) spacing
Interface goes from coherent to semi-coherent to incoherent
Precipitate goes from GP zone

GP Zones
Cu rich zones fully coherent with the matrix low interfacial energy
(Equilibrium phase has a complex tetragonal crystal structure which has incoherent
interfaces)
Zones minimize their strain energy by choosing disc-shape to the elastically soft <100>
directions in the FCC matrix
The driving force (Gv Gs) is less but the barrier to nucleation is much less (G*)
2 atomic layers thick, 10nm in diameter with a spacing of ~10nm
The zones seem to be homogenously nucleated (excess vacancies seem to play an
important role in their nucleation)

Selected area diffraction (SAD) pattern, showing


streaks arising from the zones.

5nm
Bright field TEM micrograph of an Al-4% Cu alloy
(solutionized and aged) GP zones.

5nm

Atomic image of Cu layers in Al matrix

Due to large surface to volume ratio the fine precipitates have a tendency to coarsen
small precipitates dissolve and large precipitates grow
Coarsening
in number of precipitate
in interparticle (inter-precipitate) spacing
reduced hindrance to dislocation motion (max = Gb/L)

''

Distorted FCC
10 nmthick ,100 nmdiameter
UC composition Al6Cu2 = Al3Cu

(001) '' || (001)

[100] '' || [100]

'
Becomes incoherent
as ppt. grows

(001) ' || (001)

[100] ' || [100]

Tetragonal
UC composition Al4Cu2 = Al2Cu

BCT, I4/mcm (140),


a = 6.06, c = 4.87, tI12
UC composition Al8Cu4 = Al2Cu
Phase Transformations in Metals and Alloys, D.A. Porter and K.E. Easterling, Chapman & Hall, London, 1992.

Successive lowering if free


energy of the system

Schematic diagram showing the lowering of the Gibbs free energy of the system on sequential transformation:
GP zones
Phase Transformations in Metals and Alloys, D.A. Porter and K.E. Easterling,Chapman & Hall, London, 1992.

The activation barrier for


precipitation of equilibrium ()
phase is large

But, the free energy benefit in each step is small compared to the
overall single step process

Single step
(equilibrium) process
Schematic plot

In this diagram additionally information has


been superposed onto the phase diagram (which
strictly do not belong there- hence this diagram
should be interpreted with care)
The diagram shows that on aging at various
temperatures in the + region of the phase
diagram various precipitates are obtained first

At higher temperatures the stable phase is


produced directly
At slightly lower temperatures is produced
first

At even lower temperatures is produced first

The normal artificial aging is usually done in


this temperature range to give rise to GP zones
first
Precipitation processes in solids, K.C. Russell, H.I. Aaronson (Eds.), The Metallurgical Society of AMIE, 1978, p.87

Precipitation Sequence in some precipitation hardening systems


(Morphology and compound stoichiometry are given in brackets)

Base
Al

Cu

Fe

Ni

Alloy

Precipitation Sequence

Al-Ag

GPZ (Spheres) ' (plates) (Ag2Al)

Al-Cu

GPZ (Discs) '' (Discs) ' (Plates) (CuAl2)

Al-Cu-Mg

GPZ (Rods) S' (Laths) S (Laths, CuMgAl2)

Al-Zn-Mg

GPZ (Spheres) ' (Plates) (Plates/Rods, Zn2Mg)

Cu-Be

GPZ (Discs) ' (CuBe)

Cu-Co

GPZ (Spheres) (Plates, Co)

Fe-C

-carbide (Discs) Fe3C (Plates)

Fe-N

'' (Discs) Fe4N (Plates)

Ni-Cr-Ti-Al

' (Cubes/Spheres)

Details in practical aging curves

at start

Points to be noted:
In low T aging (130C) The aging curves have more detail than the single peak as discussed
schematically before.
In low T aging (130C) the full sequence of precipitation is observed (GPZ '' ').
At high T aging (190C) '' directly forms (i.e. the full precipitation sequence is not observed).
Peak hardness increases with increasing Cu%.
For the same Cu%, the peak hardness is lower for the 190C aging treatment as compared to the
130C aging treatment.
Peak hardness is achieved when the microstructure consists of a ' or combination of (' + '').

[1] J.M. Silcock, T.J. Heal and H.K. Hardy, J. Inst. Metal. 82 (1953-54) 239.

Particle/precipitate Coarsening

There will be a range of particle sizes due to time of nucleation and rate of
growth
As the curvature increases the solute concentration (XB) in the matrix adjacent to
the particle increases
Concentration gradients are setup in the matrix solute diffuses from near the
small particles towards the large particles
small particles shrink and large particles grow
with increasing time * Total number of particles decrease
* Mean radius (ravg) increases with time

Gibbs-Thomson effect

Gibbs-Thomson effect

3
avg

r0 ravg at t = 0
D Diffusivity
Xe XB (r = )

r kt
3
0

k D Xe

ravg

D is a exponential function of temperature


coarsening increases rapidly with T

r0

t
Volume diffusion rate

Rate controlling
factor

dravg
dt

k small ppts coarsen more


2
rapidly
ravg

3
Linear ravg
versus t relation may break down due to diffusion short-circuits

or if the process is interface controlled

Interface diffusion rate

Rateof coarsening depends on D X e (diffusion controlled)


Precipitation hardening systems employed for high-temperature applications must
avoid coarsening by having low: , Xe or D
Low
Nimonic alloys (Ni-Cr + Al + Ti)
Strength obtained by fine dispersion of [ordered FCC Ni3(TiAl)] precipitate in FCC Ni
rich matrix
Matrix (Ni SS)/ matrix is fully coherent [low interfacial energy = 30 mJ/m2]
Misfit = f(composition) varies between 0% and 0.2%
Creep rupture life increases when the misfit is 0% rather than 0.2%

Nimonic 90: Ni 54%, Cr 18-21%, Co 15-21%, Ti 2-3%, Al 1-2%

Low Xe
ThO2 dispersion in W (or Ni) (Fine oxide dispersion in a metal matrix)
Oxides are insoluble in metals
Stability of these microstructures at high temperatures due to low value of Xe
The term DXe has a low value

Low D
ThO2 dispersion in W (or Ni) (Fine oxide dispersion in a metal matrix)
Cementite dispersions in tempered steel coarsen due to high D of interstitial C
If a substitutional alloying element is added which segregates to the carbide rate of
coarsening due to low D for the substitutional element

You might also like