Вы находитесь на странице: 1из 49

PHASE TRANSFORMATIONS

Phase Transformations in Metals and Alloys


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

 Nucleation & Growth


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

1
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.
 In transformations involving Diffusion T & t are „involved‟. These transformations can
be „quenched‟. First order transformations involve nucleation and growth, while in
second order transformations the entire volume transforms.
 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 Diffusionless
Involves long range mass transport E.g. Martensitic

PHASE TRANSFORMATIONS
Based on
order
1nd order 2nd (& higher) order
nucleation & growth Entire volume transforms

Classification of phase transformations

2
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.
Recrystallization of a cold worked metal is an example of a microstructural transformation.

Phases Defects Residual stress

Phases can transform Defect structures can change Stress state can be altered

Phase Defect Structure Stress-State


Transformation Transformation Transformation

Geometrical Physical
Structural Property

Phases Microstructure

Phases Transformations 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…)
 Is melting point the same as the freezing point?
Further: Does there exist a freezing point?

3
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 (S) stable Liquid (L) stable


G →

Gv
Solid (GS)

GL→S → ve
Assume for now that
we are at a fixed T (=
T1) below the Tm
T Liquid (GL)
T - Undercooling GL→S → +ve Note that Tm is the melting point
of the bulk solid
T → T1 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 + Growth


of crystals from melt of nucleated crystals till liquid is exhausted

Homogenous
Nucleation Heterogenous nucleation sites
Heterogeneous  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

4
Homogenous nucleation

 Let us consider LS transformation taking place by homogenous nucleation. Let the
system be undercooled to a fixed temperature (T held constant). 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 interface 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 interface energy  increase in strain energy

ΔG  (Volume).(GV )  (Interface).( )

Gv  f (T ) 4 
ΔG   r 3 .(Gv )  4r 2 .( )  
3 
 Note that GV is negative

f (r ) r3  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
r2 „factors/parameters‟, but the essential logic remains.

r 1

4 
 
ΔG   r 3 .(Gv )  4r 2 .( ) Note that G is a function of T, r & 
3 

 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 dG/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). dG
0
dr
r1*  0 Trivial solution
dG
dr
0
r2*  
2 G *
Gv
G  0
2 As Gv is ve, r*is +ve
r*  
Gv

16  3 Note that we are r* r0


G *  
G →

at a constant T
3 Gv2
Embryos Supercritical nuclei
3
r0  
G  0 Gv
r →

5
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‟.

Tm Gv  f (T )
Decreasing G*

The bulk free energy reduction is a function of undercooling

2
r*  
Gv
r → 16  3
G *  
3 Gv2

Decreasing r*
G →

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.
16 Tm2
G   3
3 T 2 H 2

Turnbull‟s approximation

Tm  T T
G  H f  H f
Tm Tm

Solid (GS) ΔH f  heat of fusion


G 2
16 3  Tm 
G*    
3  H f T
 
G →

T
Liquid (GL)
Tm
T →

6
Is the melting point same as the freezing point? Take the example of a pure metal.

 Usually, as we heat a pure metal, it melts at single temperature called the melting point
(Tm). [Proviso, sufficient heat is available].
 Somehow, „strangely‟, the entire „lattice‟ collapses at a single temperature.
 However, in the cooling direction (i.e. on cooling the melt) freezing can occur at any
temperature below Tm.
 At Tm itself (i.e. at zero undercooling) there is no tendency for solidification to start.
 Though the solid state is energetically (in terms of G) favorable below the melting point,
freezing actually may not start „for long times‟ if we are just below the melting point.
 Heterogeneous nucleation sites may help the solidification process below melting point.
 Hence, there is a fixed melting point, but there is no fixed freezing point (even for a pure
metal).

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/volume/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*.

 G * 
 
 kT 
N *  N t e 

7
 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
by:  H 
  d

 '  s* e kT 

No. of critical sized Frequency with which they


Rate of nucleation = 
particles become supercritical
dN
I  G * 

 kT 
  H d 
 
dt N  Nt e
*    '  s* e kT 

No. of particles/volume in L  → 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

 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
T = Tm → G* =  → I = 0
Increasing T

 G* ↑  I ↓
 G*  H d 
  Note: G* is a function of T
 
I  Nt s*  e
T (K) →

 kT 
T↑ I ↑

T=0→I=0
0 I →

8
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).

ΔG 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) Lost Acircle 
 
Surface tension force balance   Cos       (1)

ΔG  (Vlens)Gv  (Alens)   ( Acircle)    ( Acircle)   (2)

Vlens = h2(3r-h)/3 Alens = 2rh h = (1-Cos)r rcircle = r Sin

9
 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 .

dG 2  4  
 
3
0 *
rhetero  Ghetero
*
  2  3Cos  Cos3
dr Gv 3 Gv2

  Ghetero
*

 
1 * 1 Just a function of 
Ghetero
*
 Ghomo 2  3Cos  Cos3  2  3Cos  Cos 3
4 *
Ghomo 4  the contact angle

Decreasing tendency to wet the substrate

Increasing
contact angle

Complete wetting
 = 0  f() = 0
Ghetero
*

*
Ghomo
1
4

 2  3Cos  Cos 3  f ( )   = 90  f() = ½ 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.

1
G*hetero (180o) = G*homo
no benefit
0.75
G*hetero (0o) = 0
G*hetero / G*homo →

no barrier to nucleation G*hetero (90o) = G*homo/2

0.5
    
Cos 
 
0.25

Complete wetting Partial wetting No wetting


0
0 30 60 90 120 150 180
 (degrees) →

10
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).      
Cos 
 Hence, nucleating agents are added to the melt (e.g. Ti for Al  
alloys, Zr for Mg alloys) for grain refinement.

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 Å) is used a heterogeneous nucleating agent in the production of
artificial diamonds (FCC, a = 3.57 Å) from graphite.

Examples of heterogeneous nucleation

 CO2 is dissolved into an aerated drink (say sprite) at high pressures (to give us
the fizz*).
 When a bottle of an aerated drink is opened it is under-pressurized (akin to
undercooling of a melt).
 CO2 would like evolve to bring the solubility corresponding to that for
atmospheric pressure. But, the “nucleation barrier” will not allow this to
happen immediately.
 Heterogeneous nucleation at straw and glass walls help this process.

Precipitates at triple points Bubble nucleation at straw walls

* Life with fizz and without fuss is fabulous !!

Precipitates at grain boundaries

Precipitate free zone adjacent to the


grain boundaries

Heterogeneous nucleation in AlMgZn alloy

11
Why does heterogeneous nucleation dominate? (aren‟t there more number of homogenous nucleation sites?)

 To understand the above questions, let us write the nucleation rate for both cases as a pre-
exponential 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.

 Ghomo
*   Ghetero
* 
   
 kT   kT 
I homo  I homo
0
e  
I hetero  I hetero
0
e  

= f(number of nucleation sites) = f(number of nucleation sites)


~ 1042 ~ 1026

BUT
the exponential term dominates

Ihetero > Ihomo

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 Growth of  phase


→
= Nucleation of  phase + till  is exhausted*

12
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.

dX 
Transforma tion rate  f(Nucleation rate, Growth rate) T   f (I , U )
r
dt
Tm
U Maximum of growth rate usually at higher
temperature than maximum of nucleation rate
Increasing T

Tr

I
T (K) →

0 I, U, Tr → [rate  sec1] * The U-T curve is an alternate way of stating the U-T curve

13
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).

Using „some‟ model


1.0
 π I U3 t 4  Saturation phase decreasing growth rate with time
 
 (region of law of diminishing returns)
Xβ  1  e  3 

0.5 Linear growth regime ~constant high growth rate


X →

Incubation period  slow growth (but with increasing growth rate with time)

t →
0

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 Time-
Temperature-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 Tm
Rate  f (T , t ) Small driving
force for nucleation
Nose of the „C-curve‟

Tr
T (K) →

Time for transformation


Replot
Tt
T (K) →

Sluggish growth

Tr (rate  sec1) → Tt (time  sec) →


0 0

14
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

volume fraction of  at t
 Fraction f 
final volume of 
transformed
 
99% = finish

1% = start
f  volume fraction of 
T (K) →

t (sec) → These are iso-fraction lines.

15
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

16
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 = 4r2 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 4U 3t 2dt


2

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

dX dX
f    N n 4 U 3t 2 dt
1 X 1 X

X t t  4π N n U3 t 3 
dX
0 1  X  t 0 Nn 4U t dt
3 2  

Xβ  1  e  3 

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

17
 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


Pearlitic transformation
All of the parent phase is consumed by the product phase
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

18
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
f   r 3  Id    U (t   )  Id 
3

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

dX dX 4 4
f    r 3  Id    U (t   )  Id 
3

1 X 1 X 3 3

X  t  πIU t 
3 4
dX 4
     Note that X is both a function of I and
0 1  X  0 3   
3
 U (t  ) Id 
Xβ  1  e 
3  U. I & U are assumed constant

1.0
For a isothermal transformation

 π I U3 
   is a constant during isothermal transformation
 3 
0.5
X →

0
t →

19
APPLICATIONS
 of the concepts of nucleation & growth
 TTT/CCT diagrams

Phase Transformations in Steel

Precipitation

Solidification, Crystallization and Glass Transition

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
hyper-eutectoid 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).

20
We have already seen the Fe-Fe3C phase diagram (please have a second look!)

Fe-Cementite diagram

Peritectic L
L+→  Eutectic
1493ºC L →  + Fe3C

L+
0.1 %C  2.06 1147ºC

Eutectoid  + Fe3C
 →  + Fe3C
723ºC

0.025 %C  + Fe3C
T →

Fe %C → Fe3C
0.16 0.8 4.3 6.7

Pro
eutectoide
ferrite

Perlite
ferrite

21
 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.

800
TTT diagram for Eutectoid steel (0.8%C)
723C Eutectoid temperature
Eutectoid steel (0.8%C)
700
Pearlite
Above Eutectoid
600
temperature there is no
transformation „Nose‟ of „C‟ curve
 + Fe3C
500 Pearlite + Bainite
Important points to be
noted:
400 Bainite
 The x-axis is log scale.
T →

Austenite
 „Nose‟ of the „C‟ curve is in
~sec and just below TE
transformation times may be 300
~day.
 The starting phase has to Ms
. 200
 The ( + Fe3C) phase
field has more labels 100 Mf
included. Martensite
t (s) →
 There are horizontal
lines labeled Ms and Mf. 0.1 1 10 102 103 104 105

 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 (~225C, below the temperature marked Ms), then Austenite
transforms via a diffusionless transformation (involving shear) to a (hard) phase known as
Martensite (named after German scientist Martens). 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 M s and
Mf). 800
Eutectoid steel (0.8%C)
723C Eutectoid temperature
700
How are these TTT diagrams drawn? Pearlite
600
 Samples are quenched into a salt bath maintained at  + Fe3C
Pearlite + Bainite
various temperatures (practical version of the 500

„instantaneous quench‟) 400 Bainite


T →

Austenite
 The samples are then quenched from this bath to 300
room temperature after various times. Ms
200
 Phase fraction of transformed phase is determined by
100 Mf
optical metallography. Martensite
t (s) →
0.1 1 10 102 103 104 105

22
 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 and cool below Mf. Called the critical cooling rate.

 In the pearlitic region, at higher 800


Eutectoid steel (0.8%C)
temperature the interlamellar spacing 723C Eutectoid temperature
700
will be more (i.e. we obtain a coarse pearlite as Austenite Coarse
growth rate will predominate over nucleation rate). T1 Pearlite
600
 If we quench between Ms and Mf we Fine
will get a mixture of Martensite and 500 T2 Pearlite + Bainite
Austenite  (called retained
Austenite). 400 Bainite
T3
 Why does this retained Austenite not
transform?
Not an isothermal
300

200
 Ms
Austenite

transformation
100
Mf
T →

Martensite
0.1 1 10 102 103 104 105
t (s) →

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)

23
TTT Diagram: hypoeutectoid steel
 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
to below TE, we should get a mix of  + Fe3C  what is the meaning Hypo-Eutectoidof a „pro‟-eutectoid
steel
phase in a TTT diagram? (remember „pro-‟ implies „pre-‟).(Considered next)

Atlas of Isothermal Transformation and Cooling Transformation


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

C1

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


undercooling wrt A1 line (formation of pro-eutectoid )
(formation of  + Fe3C)

C1

24
 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 (for plain carbon steel)* is
that in the CCT case Bainite cannot form.

* For certain alloy steels Bainite can form on cooling as we shall see later.

25
Continuous Cooling Transformation (CCT) Curves
Points to be noted:
 Using CR1: the  phase begins to transform to pearlite, but the transformation is not completed. The remaining 
transforms to Martensite on crossing the Mf line (point C).
 Using CR2: the  phase completely transforms to pearlite (after point E). Hence, there is no significance of the crossing of
the CR2 line of the Ms (point F) and Mf lines (point G).
800
 As before the x-axis is log Eutectoid steel (0.8%C)
scale. Eutectoid temperature
723
Start
 Bainite cannot form by
continuous cooling.
600 D Finish
 Constant rate cooling curves Pearlite
look like curves due to log E CCT curves
A
scale in x-axis. The higher 500
cooling rate curve has a higher
(negative) slope.
 Original TTT lines
400
T →
 As time is one of the axes, no
treatment curve can be drawn
where time decreases or 300
remains constant. Ms B F
200
Cooling curves CR1 CR2
Constant rate Mf
Constant 100 C G
T1 T2
Cooling rate Martensite
dT T1 > T2
T 0.1 1 10 102 103 104 105
dt CR1 CR2 t (s) →

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 steel (0.8%C)
Eutectoid temperature
723

600
Pearlite

500
 Original TTT lines
400
T →

300
Ms
200
Cooling curves
Constant rate Mf
100
T1 T2
Martensite

0.1 1 10 102 103 104 105


t (s) →

26
Common heat treatments involving cooling

 Common cooling heat treatment labels (with increasing cooling rate in steels) 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 (i.e. toughness).
800
Eutectoid steel (0.8%C)
723

600
Note: this is „Microstructure
Engineering‟ (changing properties
Water qu

without changing the composition) 500


Fu
No

ll a
T →

rm

400
nne
ali
ench

al
zin
g

300
Oi

When the „nose‟ of the CCT diagram is


lq

avoided by fast cooling (quench) then


uen

complete transformation to Martensite is Ms


200
ch

possible if cooled below Mf Coarse P

100 Mf
M M + P Fine P
P = Pearlite
0.1 1 10 102 103 104 105
M = Martensite t (s) →

27
What are the typical cooling rates of various processes?

Process Cooling rate (K/s) Comments


Furnace cooling (Annealing) 105 – 103 Typically for solid samples
Air Cooling 1 – 10 “
Oil Quenching* ~100 “
Water Quenching* ~500 “
Splat Quenching 105 For molten material
Melt-Spinning 106 – 108 “
Evaporation, sputtering 109 (expected) Gaseous state involved

* Depends on conditions discussed later

Pearlite
 →  + Fe3C

Lamellae of Pearlite in ~0.8% carbon steel

 Steel of eutectoid composition completely transforms to Pearlite. Pearlite is a


microconstituent consisting of alternating lamellae of ferrite and cementite. Though
cementite is brittle, this morphology is reasonably ductile.
 Pearlite forms by nucleation and growth processes (as explained later). This occurs
by heterogeneous nucleation at grain boundaries.
 The interlamellar spacing of pearlite is a function of the temperature of
transformation. The lower the temperature, higher will be the nucleation rate as
compared to growth rate → this implies that finer will the pearlitic spacing → and
higher will be the hardness.

28
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 Kurdyumov-
Sachs 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.

1 2 3 4

(100)C || (1 11)
Orientation Relation:
Kurdyumov-Sachs (010)C || (110) Branching mechanism
(001)C || (112)

Bainite
 →  + Fe3C**

Micrograph courtesy: Prof. Sandeep Sangal

 Bainite formed at high temperature (~ 350C) has a feathery appearance and is called
„Feathery Bainite‟ (i.e. refers to the morphology of the microstructure).
 Bainite formed at lower temperature (~ 275C) has a needle-like appearance and is called
„acicular Bainite‟.
 The process of formation of bainite involves nucleation and growth (akin to the formation
of pearlite).
 Typically this transformation is accompanied by surface distortions (especially in acicular
Bainite).
 ** Lower temperature → carbide could be ε 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.

29
More images of Bainite AFM image
0.8% C steel, the sample was quenched in a salt
bath having 400°C temperature and then it was
held for 2 hours.

Micrograph courtesy: Prof. Sandeep Sangal, Swati Sharma

Micrograph courtesy: Prof. Sandeep Sangal, Swati Sharma

What is the difference between


pearlite and bainite?

 Basically, both are micro-constituents


consisting of ferrite and carbide phases. It is
the morphology & distribution of these phases
which distinguishes one from the other.

Characteristic of Martensitic transformations


 Though originally described in steel, the name (Martensitic transformation) now applies to this
kind of displacive transformations (involving shear). The „essential reason‟ for the transformation
is that due to high undercooling there is high propensity for transformation, but due to the fast
cooling rate there is no time for the equilibrium diffusional transformation to take place. This
„pent up‟ driving force leads to the cooperative shear mechanism of Martensitic transformation.
 Shape of the Martensite formed in steel→ Lenticular (or thin parallel plates). The transformation
is 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

30
Martensite Change in Crystal Structure  ( FCC )  ' ( BCT )
Quench

0.8 %C 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
FCC Austenite alternate choice of Cell
the Matensitic transformation
c=a
~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*. If pure Fe (without „C‟) undergoes Bain distortion then
we obtain BCC Fe (the usual stable product at low temperatures).

C along the c-axis


obstructs the contraction

Tetragonal
Austenite to Martensite → ~4.3 % volume increase 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.

31
How is invariant plane condition obtained?
 Experimentally it is observed that the habit plane remains „invariant‟ (Fig. A).
 Bain distortion will lead to the rotation of the habit plane (Fig.B).
 If a shear is applied then the habit plane can be made invariant (Fig.C).
 But, this shear will lead to the distortion of the lattice (Fig.D).
 This implies that Slip or Twinning also must occur to keep the lattice undistorted (Fig.E)

Fig.D

Fig.A
Fig.C
Fig.B

But shear will distort the lattice!

Slip Twinning

The average shape remains undistorted

Fig.E

Martensitic microstructure
Martensite in 0.6%C steel

Examples of Martensitic transformation in other systems


As pointed out before, the term Martensitic transformation applies to other systems (i.e. non-
ferrous) as well. This includes non-ferrous alloys, ceramics and inorganic compounds.

[*] MS (C) Hardness (HV)


ZrO2 927 1000
Fe–31Ni–0.23C wt% 190 300
Fe–34Ni–0.22C wt% 269 250
Fe–3Mn–2Si–0.4C wt% 220 600
Cu–15Al 20 200

* Source: http://www.msm.cam.ac.uk/phase-trans/2000/C9/lectures45.pdf

32
Summary of characteristics of Martensitic transformation
 The transformation occurs by shear without need for diffusion (and hence it takes place
without a change in the composition).
 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. If the Mf temperature is below RT (and MS being above RT) and
the sample is quenched in a medium at RT, then there will be some amount of retained
Austenite).

 Hardness of martensite is a function of


the carbon content (in fact the carbon is 60
primarily responsible for the hardness of

Hardness (Rc) →
martensite even in alloy steels)
→ but high hardness steel is very brittle 40 Hardness of Martensite as a
as martensite is brittle. function of Carbon content
 Steel is reheated to increase its ductility
→ this process is called TEMPERING.
20
% Carbon →
0.2 0.4 0.6

Microstructure Engineering in Action !

 What is the hardness of 0.8% carbon steel? → There is no answer to this question !
 The hardness depends on the microstructure of the steel, which is determined by
the heat treatment process used.
 The hardness of 0.8% carbon can be as low as 16 Rc or as high as 65 Rc.
 So we have two handles on the hardness: the composition and the microstructure.
 As we shall see later, often we will aim for a graded microstructure (e.g. in case
carburized steel, where the surface is made hard, with a touch interior).

Properties of 0.8% C steel


Constituent Hardness (Rc) Tensile strength (UTS) (MN/m2)
Coarse pearlite 16 710
Fine pearlite 30 990
Bainite 45 1470
Martensite 65 -
Martensite tempered at 250C 55 1990

33
Alloying of steel

 Plain carbon steel is inexpensive and has the „appeal of simplicity‟. But, plain
carbon steel suffers from  Low hardenability,  Loss of hardness on tempering,
 Low corrosion and oxidation resistance,  Loss of strength at high temperatures.
 Hence, there is a strong reason to alloy plain carbon steel with various other
substitutional alloying elements (like Cr, Mn, Ni, W, Mo, etc.). Many of the alloys
may contain multiple alloying elements. These elements are expected to give a
combination of „improved‟ properties.
 Alloying elements added can: (i) segregate/phase separate or (ii) form a solid
solution or (iii) form a compound. (A combination of these is also possible).

Segregation / phase separation


Interstitial
Element Added Solid solution
Substitutional
Compound (new crystal structure)

 Most of the alloying elements move the TTT diagrams to the right (except Co moves the
TTT curve to the left*). In plain carbon steel of say 0.8% C, during (say) pearlitic
transformation the uniform carbon concentration has to split into regions with high C (Fe3C)
and low C (ferrite). In alloy steels in addition to the carbon „redistribution‟, the alloying
element(s) have to redistribute as well. This diffusion (of the alloying elements) takes additional time
and hence the TTT diagram (also the CCT) diagram will „move to the right‟ (i.e. will take
longer times for the transformations).
 This (moving of TTT/CCT to the right) has important consequences: (i) a „less-severe quench‟ can be
employed to obtain martensite by avoiding the nose of the TTT/CCT diagram (which implies
lower residual thermal stresses in the material/component), (ii) larger cross-sections can be through
hardened (i.e. at least 50% martensite can be obtained in the centre of (say) a circular cross-
section).
 A variety of alloy steels have been developed with an amazing set of properties and
underlying „mechanics‟. These include: Maraging steels, TRIP (transformation induced
plasticity) steels, TWIP (twinning induced plasticity), high strength low alloy (HSLA) steels,
etc.

• Alter temperature at which the transformations occurs.


Alloying
• Alter solubility of C in  or  Iron.
elements
• Alter the rate of various reactions.

* I don‟t know the reason for this !

34
 The role of alloying elements is summarized in the figure below. (Note that not all alloying elements can
do play all the roles!).
 It is clear that alloying elements can provide a variety of benefits.
 Some elements which are not intentionally added, but are accidentally present (i.e. could not
be removed during extraction) are called impurities. Examples of (especially) deleterious
impurities in steel are P (causes cold shortness) and S (causes hot shortness/sulphur
embrittlement/red shortness).
 Hot shortness is caused by the formation of iron sulphide (FeS) along grain boundaries
(where S segregates). This FeS melts at low temperatures causing grains to separate (at grain
boundaries). This problem can be alleviated by the addition of Mn which forms MnS (with a
higher melting point).
 P present in quantities greater than 0.12% causes cold shortness (i.e. lack of ductility during
cold working*).
• Increase (↑) hardenability.
• Provide a fine distribution of alloy carbides during tempering.
• Increase (↑) resistance to softening on tempering.
Alloying • Increase (↑) corrosion and oxidation resistance.
elements • Increase (↑) 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).
* A formal definition of hot and cold working will be dealt with when we discuss Recovery_Recrystallization_Grain_Growth

Sample elements and their role

 The topic of the role of alloying elements in steel is a book by itself. Some points
are given below.
 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.

35
Effect of alloying elements
 Alloying elements increase hardenability but the major contribution to hardness comes from
Carbon.
 When it comes to the stabilization of a phase in steel on the addition of alloying elements
the thumb rule is that ‘like stabilizes like’. CCP elements stablize CCP phases and BCC
stablize BCC phases. This increase in stability is reflected as an expansion of a phase field.
 Mn & Ni which are FCC (CCP) metals are Austenite stabilizers.
 Cr (BCC) is Ferrite stabilizer. This implies that increasing Cr will lead to a shrinkage of the
 phase field.

Temperature →
0.35% Mn

Mn +0.1% C
180 on
rb
Ca
Brinell Hardness →

of Mn
on
iti 6.5% Mn
dd Cr + 0.1%C C (%) →
A
140
0 0.4 0.8 1.2 1.6
n of Carbon
Additio
100 Cr Temperature → Outline of the  phase field

Alloying Element (%) →


60
v

15% Cr 12% Cr
0 2 4 6 8 10
5% Cr 0% Cr

C (%) →
0 0.4 0.8 1.2 1.6

TTT diagram for alloy steels


 In alloy steels the „C‟ curves for the Bainitic and Pearlitic transformation are separate.
 This implies that we can obtain Bainite even by continuous cooling.

 The diagram on the right is more accurate and


the ferrite arm can be seen at high
temperatures.
 The nose of the pearlitic transformation is at
longer times, but that of the bainitic
transformation is at shorter times.
800

700
Austenite Pearlite
600
Ni-Cr-Mo low alloy steel
500
T →

400
TTT diagram of low alloy steel
300 Bainite (0.42% C, 0.78% Mn, 1.79% Ni,
Ms 0.80% Cr, 0.33% Mo) U.S.S. Carilloy Steels,
200 United States Steel Corporation, Pittsburgh, 1948)

Mf
100
Martensite t →
~1 min

36
Effect of carbon content and heat treatment on properties of steel
1000

900 0.4% C - Slow cooled


0.8% C - Slow cooled
800
0.8% C - quenched
Engineering Stress (s) [MPa]

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
400 Slowly cooled- 0.8% C
Slowly cooled- 1.0% C

Vikers Hardness
350

300

250
Hardness
200

150
0.5 0.6 0.7 0.8 0.9 1 1.1
C%

Precipitation Hardening
The study of precipitation hardening is ‘enough’ to understand ‘virtually’ the whole of physical metallurgy.

To understand the processes and benefits we need to understand:


 Vacancies, dislocations and crystal structures.
 G-X curves, activation energy and diffusion.
 Interfaces and strain energy.
 Phase diagrams and TTT diagrams.
 Quenching, aging and metastability.
 & perhaps more!

37
Precipitation Hardening

 The presence of dislocations weakens the crystal → leading to easy plastic


deformation.
 Putting „hindrance‟ to dislocation motion increases the strength of the crystal.
 Fine precipitates dispersed in the matrix provides (one) such an impediment.
 „Al‟ is a marvelous material (light weight, good oxidation resistance due to the
tough adherent oxide layer on the surface, relatively inexpensive, etc.) but suffers
from a low strength of 100 MPa*.
 The strength of Duralumin (Al+4% Cu+other alloying elements) with proper heat
treatment can reach 500 MPa !
 In this section we will discuss the heat treatment process which will give us such a
enhanced strength.

* We have already noted that strength (e.g. yield strength, UTS, fracture strength) is a microstructure sensitive property and this value is just an
„indicator‟.

Philosophy behind the process steps in Precipitation Hardening


 If a high temperature solid solution is slowly cooled, then coarse (large sized) equilibrium
precipitates are produced. These precipitates have a large distance between them. These
precipitates have incoherent boundaries with the matrix (incoherent precipitates).
 Such (coarse) precipitates, which have a large inter-precipitate distance, are „not the best‟ in
terms of the increase in the hardness.
 Hence, we device a 3 step process to obtain a fine distribution of precipitates, which have a
low inter-precipitate distance, to obtain a good increase in hardness.

 ( FCC )   ( FCC )   CuAl2 (Tetragonal) 


  slowcool    
  4 % Cu    0.5 % Cu    52 % Cu 
 550 o C   RT   
+      RT 

 →+ *Also refer section on Double Ended Frank-


 Slow equilibrium cooling gives rise to coarse  Read Source in the chapter on plasticity:
precipitates which is not good in impeding max = Gb/L
dislocation motion.*
4 % Cu

Multi-step process used to


Coarse incoherent obtain a fine distribution of
precipitates, with large inter- precipitates (with small
precipitate distance inter-precipitate distance)

38
Al-Cu phase diagram: the sloping solvus line and the design of heat treatments

 The Al-Cu system is a model system to understand precipitation hardening (typical


composition chosen is Al-4 wt.% Cu).
 Primary requirement (for precipitation hardening) is the presence of a sloping solvus line
(i.e. high solubility at high temperatures and decreasing solubility with decreasing
temperature). In the Al rich end, compositions marked with a shaded box can only be used
for precipitation hardening.

Sloping Solvus line:


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

Al
Cu

Heat treatment steps to obtain a fine distribution of precipitates


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

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


  Assume that we start with a material having coarse
B
equilibrium precipitates (which has been obtained by prior slow cooling of the
sample).

A  A: We heat the sample to the single phase region () in


C
the phase diagram (550C).
 B: We quench (fast cooling) the sample in water to
+ obtain a metastable supersaturated solid solution (the amount
of Cu in the sample is more than that allowed at room temperature according to the
phase diagram).

4 % Cu  C: We reheat the sample to relatively low temperature


(~180C/200C) get a fine distribution of precipitates. We
have noted before that at „low‟ temperatures nucleation is dominant over growth.

A Heat (to 550oC) → solid solution 


Supersaturated solution
B Quench (to RT) →
Increased vacancy concentration
C Age (reheat to 200oC) → fine precipitates

39
Effect of Aging Temperature
 Higher the aging temperature, the less the time it takes to obtain peak hardness.
 Lower the temperature of aging, higher the peak hardness obtained.
 This implies that we have to optimize between time spent in aging and hardness required.
This optimization is very important in an industrial production scenario.
 For the Al-4wt.%Cu alloy the typical aging time chosen is 180C.

Schematic curves →
Hardness →

Real experimental curves 100oC


are in later slides

Note that the peak hardness


obtained is considerably less

180oC 20oC
Like room temperature (natural) aging.
Peak hardness not yet obtained at this temperature.

Log(t) →
Note: Schematic curves shown- real curves considered later

180oC Peak-aged
Hardness →

Dispersion of Coarsening
fine precipitates of precipitates
(closely spaced) with increased
inter-precipitate spacing

Underaged Overaged
Not zero of

Log(t) →
hardness scale

Region of precipitation
Region of solid solution hardening
strengthening→ Hardness is higher than that of Al (but little/some solid solution
(no precipitation hardening) strengthening)

40
A complex set of events are happening in parallel/sequentially during the aging process.
These are related to:
 The size and the inter-precipitate distance → the size of the precipitates increases and during
the coarsening stage the inter-precipitate distance increases,
 The nature of the precipitate → metastable phases transform to stable phases,
 The nature of the interface → the interface goes from coherent to semi-coherent to
incoherent (with increasing size).
 When the precipitate is coherent, plastic deformation can take place by precipitate shearing.
The CRSS increase has a r1/2 dependence. The by-pass mechanism (Orowan bowing
mechanism) will be operative when the precipitates are incoherent (with 1/r dependence).
This change in mechanism of deformation gives the typical peak in the H-t plot.
The path (mechanism)
Hardness →

offering the lease resistance


will be operative

CRSS Increase →
Precipitate
shearing
1
Log(t) → 1 
r 2 r Precipitate
By-pass
Increasing size of precipitates & increasing inter-precipitate spacing
Particle radius (r) →
Interface goes from coherent to semi-coherent to incoherent
Click here to know more about interfaces
Precipitate goes from GP zone → ‟‟ → ‟ → 

GP Zones (Guinier–Preston zones)

 GP zones are Cu rich zones fully coherent with the matrix and can be observed only in a
transmission electron microscope. The strain field associated with GP zones is can be
imaged in a bright field image (TEM).
 This is a state in the borderline between a precipitate and a „segregation‟.
 GP zones have a low interfacial energy with the Al-Cu alloy matrix.
 (It is to be noted that the 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 (G*) is much less.
 These zones are typically 2 atomic layers thick, 10nm in diameter and with a spacing of
~10nm (in the Al-Cu system).
 The zones seem to be homogenously nucleated (excess vacancies seem to play an important
role in their nucleation→ we have already noted that the quenching process gives rise to an
excess concentration of vacancies).

Disc shaped in 3D

Section of GP zone parallel to (200) plane

41
More views on GP zones
Selected area diffraction (SAD) pattern,
showing streaks arising from the zones.

020

200

[001]

Bright field TEM micrograph of an Al-4% Cu


alloy (solutionized and aged)  GP zones.
(Note the scale of the zones)

Nucleation, Growth and Coarsening of precipitates


 At a given undercooling, nucleation is an event „uphill‟ in G. This takes place if the size of
the product phase (in this case a precipitate) exceed a minimum critical size (r*).
Precipitates smaller than the critical size will tend to dissolve.
 During growth the preformed nuclei grow „downhill‟ in G. Growth will proceed till the
solute concentration reaches the equilibrium value (i.e. the supersaturated solid solution
attains a solubility as dictated by the phase diagram).
 If we restrict ourselves to a microstructural level equilibrium, then the system should be
frozen in this state. However, the interfaces between the precipitate and the matrix „costs‟
energy (are in a higher state of energy as compared to the crystal interior).
 Coarsening is driven by an decrease in interfacial free energy (per unit volume). This
constitutes the global reason, but for a process to take place the global and local reason both
have to be satisfied (we will see the local reason later). During coarsening the large precipitates will
grow at the expense of the smaller ones, leading to an increase in the average precipitate
size.

 During coarsening there is: (i) a decrease (↓) in the number of precipitates, while
(ii) there is an increase (↑) in inter-precipitate spacing.
 This leads to a reduced hindrance to motion of dislocations (as max = Gb/L). This implies
coarsening leads to a decrease in hardness.

42
The metastable and stable phases in the Al-4wt.% Cu alloy
 We have already noted that GP zones can be considered to be a precursor to a true precipitate.
 The sequence of precipitates obtained while aging depends on the temperature. At low
temperatures (as compared to the solvus line) the activation energy available is less and this
implies that the system cannot „throw out‟ (precipitate) the equilibrium  (CuAl2) phase.
 If aging is done at temperatures marked shaded green in the phase diagram (next page), then
the sequence of precipitates obtained is:
GPZ (Discs) '' (Discs)  ' (Plates)   (CuAl2)
 In the sequence above the morphology of the precipitate is also mentioned in brackets. The
morphology is determined by the interfacial energy and (more importantly) the strain energy.
 In this phase diagram additionally information has been superposed onto the phase diagram
(i.e. ‘lines of metastability’ have been overlaid). ” & ' are metastable precipitates.
 If aging is carried out at higher temperatures then the '' precipitate is obtained directly. At even
higher temperatures of aging (brown/maroon) region of the phase diagram the ' precipitate is
obtained. Finally, at even higher temperatures of aging (the yellow region) the  phase is
obtained straightaway.
 In summary the phase obtained while aging depends on the activation energy available, which
depends on the temperature of aging. Hence (given this constraint), the system may „throw out‟
metastable phases, which do not give maximum G benefit. Refer set of upcoming Figs. .

Extended Al-Cu phase diagram indicating lines of metastability

 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

43
Schematics of activation energy barriers for the transformations in the A-Cu system and the „G‟ benefit

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
Schematic plot („equilibrium‟) process

 The free energy vs composition plots give us an idea of the phases obtained during aging (low
temperature, say 180C) and „G‟ benefit involved. The common tangent construction gives the
composition of the phases in “equilibrium*”.
 GP zones (and ‟‟) have relatively higher free energy benefit, but the common tangent
construction shows that there is a overall lowering of „G‟. The G-X curve for GP zones can be
considered as an extension of that for the α phase.
 Note that the α matrix gets poorer in Cu during the precipitation sequence (α1 to α4).

Successive lowering if free


energy of the system  α0 is the composition of the original
supersaturated matrix (Al-4wt.%Cu).
 The „G‟ is lowered stepwise during the
transformation sequence:
(G0  G1) to (G3  G4).

Phase Transformations in
Metals and Alloys, D.A. Porter
and K.E. Easterling,Chapman Schematic diagram showing the lowering of the Gibbs
& Hall, London, 1992.
free energy of the system on sequential transformation:
GP zones → ‟‟ → ‟ → 
Say Al-4wt.%Cu

* I hope by now we understand the meaning of the term equilibrium in this context !

44
 The precipitation sequence can be
followed by (schematic) TTT
diagrams as well.

Q&A How does the strain energy determine the morphology of the precipitate?

 For an elastically harder precipitate (which is


usually the case), the elastic energy is least for
a oblate spheroid shaped precipitate (~like a
disc). It is highest for a spherical precipitate
and intermediate for a prolate spheroid shaped
precipitate (needle shaped).

F.R.N. Nabarro, Proceedings of the Royal Society A 175 (1940) 519.

This is a „brain raker‟ of a page: Crystallography, orientation relationship, interface characteristics, etc. information is
included about the ,  &  phases

Distorted FCC
(001) '' || (001)
10 nmthick ,100 nm diameter
UC composition Al6Cu2 = Al3Cu [100] '' ||[100]

 ''
Becomes incoherent '
as ppt. grows
(001) ' || (001)
[100] ' || [100] Tetragonal
UC composition Al4Cu2 = Al2Cu


BCT, I4/mcm (140),
a = 6.06Å, c = 4.87Å, tI12
UC composition Al8Cu4 = Al2Cu

Phase Transformations in Metals and Alloys, D.A. Porter and K.E. Easterling, Chapman & Hall, London, 1992.

45
„Practical‟ „aging curves‟ can be very different from the schematic ones considered before.

‟‟ at start

Points to be noted:
 In low T aging (130C) The aging curves have more detail than the single peak as discussed
schematically before.
 In low T aging (130C) the full sequence of precipitation is observed (GPZ  ''  ').
 At high T aging (190C) '' 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 190C aging treatment as compared to the
130C 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.

Precipitation hardening in other systems

 Precipitation hardening is used in many other systems. The sequence of precipiatation in


other systems may not be as elaborate as that in the Al-Cu system.

Base Alloy Precipitation Sequence


Al 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 Cu-Be GPZ (Discs) '   (CuBe)
Cu-Co GPZ (Spheres)  (Plates, Co)
Fe Fe-C -carbide (Discs) Fe3C (Plates)
Fe-N '' (Discs) Fe4N (Plates)
Ni Ni-Cr-Ti-Al ' (Cubes/Spheres)

Precipitation Sequence in some precipitation hardening systems


(Morphology/compound stoichiometry are given in brackets)

46
Particle/precipitate Coarsening The local reason for coarsening
 There will be a range of precipitate sizes due to time of nucleation and rate of growth.
 Smaller precipitates are highly curved as compared to larger precipitates.
 The G of the precipitate depends on the curvature of the surface (higher the curvature of the
surface, higher is its G). Hence, we have curvature dependent G-composition curves.
 As the curvature increases, the solute concentration (XB) in the matrix adjacent to the
precipitate increases. This implies that the matrix adjacent to the smaller precipitates is
richer in the solute (called the Gibbs-Thomson effect). Shown in next slide.
 This leads to concentration gradients in the matrix → solute diffuses from near the small
precipitates towards the large precipitates.
 To maintain equilibrium, the smaller precipitates dissolve a little and release some solute.
This further increases the curvature of the precipitate and thus enriching the adjacent matrix
with the solute.

 This leads to the shrinkage of small


precipitates and the growth of large
particles. Something like “rich get
richer and poor get poorer”!
 This implies that with increasing
Gibbs-Thomson effect
time: (i) total number of precipitates
decrease, while (ii) the mean radius
(ravg) increases with time.

G-X curves to understand Gibbs-Thompson effect


 Smaller precipitates are have a higher interface curvature as compared to large ones. As the
curvature increases, more unsaturated are going to be the bonds on the surface, which further
implies a higher „G‟ for the precipitate. Vary large precipitates can be assumed to have a flat
interface.
 By the common tangent construction we see that the smaller precipitates (say β with radius r2)
have a higher concentration of solute adjacent to it, as compared to larger precipitates (say β
with radius r1). This sets up the concentration gradient and leads to diffusion of solute (Cu in the case of Al-Cu system).

G-X curve
for the Gibbs-Thomson effect
matrix
(solid
solution)

Three different G-XB


curves for 3 different
curvatures of the
precipiate
(curvature dependent
G-X curves).

The composition of the matrix


adjacent to the 3 precipitates is different

47
How to avoid coarsening of precipitates
 Many industrially important precipitation hardening systems need to operate at high
temperatures (e.g. Ni based superalloys). This will lead to coarsening of the precipitates and
hence to loss of strength.
 In a diffusion controlled process if the initial average size of the precipitates is r0 and after
time ‘t’ it is ravg, the following equation is followed: 3
ravg  r0  kt k  D  X e
3

 r0 → ravg at t = 0
Rate Volume diffusion rate  D → Diffusivity
ravg controlling  Xe → XB (r = )
factor
Interface diffusion rate
r0
t
 ‘k’ is a constant which depends on Diffusivity (D), interfacial free energy () and equilibrium
concentration of the solute (Xe) .
 D is a exponential function of temperature  coarsening increases rapidly with T.
 Linear ravg3 vs ‘t’ relation may break down due to short-circuit paths for diffusion (like grain
boundary diffusion). dravg k
 The rate of coarsening can be written as:  2
dt ravg
 This implies that smaller precipitates coarsen more rapidly!

 Hence, it is clear that precipitation hardening systems employed for high-temperature


applications must avoid coarsening by having a low value for: , Xe or D. (If the process is
diffusion controlled).

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

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 X e.
 The term DXe 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.

48
Q&A How to use the nucleation rate (I) and growth rate (U) curves to make fine
grained materials?
 At low temperatures (T of operation less 0.4Tm) fine grain size bestows superior mechanical
properties to the material (Hall-Petch relation). The reverse is true at high temperature.
 When solidifying from the molten state, a high nucleation rate along with a slow growth rate
can lead to a fine grain size.
 To obtain fine grain size we can: (i) allow the transformation to take place at low
temperatures or (ii) cool fast or (iii) use heterogeneous nucleating agents. The first option is
comparatively difficult in practice (so typically samples are cooled fast to obtain a fine grain
size).
 At ↑ Cooling rates the sample spends lesser time at temperatures near Tm , where the peak of
growth rate (U) lies. This implies that the sample spends more time close to the region (in T) of
↑ nucleation rate. Thus more nuclei form which grow little. Tm
U
 At the other extreme lies the situation where we want to
grow single crystals. I.e. a single nucleus grows to
complete the solidification.

T (K) →
 Single crystals can be grown by pulling a seed crystal out
of the melt. In this case the seed crystal plays the role of
the nucleus (and nucleation in the melt is avoided). I

I, U →
0

49

Вам также может понравиться