You are on page 1of 36

CHAPTER 9

PHASE EQUILIBRIA
& PHASE DIAGRAMS
9.1

COMPONENTS AND PHASES

9.2

ONE-COMPONENT SYSTEMS

9.3

TWO-COMPONENT SYSTEMS

9.4

9.3.1

Solubility in the Solid State

9.3.2

Specification of Composition

9.5

9.6

9.7

BINARY ISOMORPHOUS SYSTEM


9.4.2

Determination of Phases Present

9.4.3

Determination of Phase Compositions

9.4.4

Determination of Phase Amounts

9.4.5

Solidification Under Equilibrium


Conditions

9.4.6

Non-Equilibrium Solidification

9.4.7

Mechanical Properties of
Isomorphous Alloys

9-1

BINARY EUTECTIC SYSTEM


9.5.1

Solidification of Eutectic Alloy

9.5.2

Solidification of Off-Eutectic Alloy

9.5.3

Alloys Without Eutectic Reactions

9.5.4

Mechanical Properties of Eutectic


Alloys

OTHER BINARY SYSTEMS


9.6.1

Invariant Reactions in Binary Systems

9.6.2

Complex Binary Phase Diagrams

IRON-IRON CARBIDE (FE-FE3C) SYSTEM


9.7.1

Transformations in Eutectoid Steels

9.7.2

Transformations in Hypoeutectoid Steels

9.7.3

Transformations in Hypereutectoid Steels

9.7.4

Mechanical Properties of Steels

9-2

9.1 COMPONENTS AND PHASES

Note that the phases in a material refer not only to the


gaseous, liquid and solid states; within the same material,

Most engineering materials are mixtures of elements

regions

of

different

composition,

different

crystal

and/or molecular compounds. Each element and/or

structures, or crystalline as opposed to amorphous, are

compound in the mixture is known as a component,

considered different phases, e.g. BCC iron is a distinct and

which is defined as a chemically distinct and essentially

separate phase from FCC iron.

indivisible substance; e.g. Fe, Si, NaCl, H2O.

Microstructure is the structure of a material on the

Material mixtures may be physical mixtures, in which the


components are unchanged chemically and remain as the
same identifiable entities. More often however, the
components interact chemically to form new constituents
or phases.

microscopic scale. It is characterized by the phases present,


their relative amounts, the composition and structure of each
phase, and size, shape and distribution of the phases.
The microstructure depends on the overall material
composition, the external temperature and pressure, and
thermal

A phase is a region of material that has the same


composition and structure throughout, and is separated

processing

history

of

the

material.

The

microstructure of a material has a profound influence on its


properties.

from the rest of material by a distinct interface. A phase


A phase diagram is a graphical representation of the

may contain one or more components.

stable phases present, and the ranges in composition,


temperature, and pressure over which the phases are

Some everyday examples of components and phases:


Material(s) present
Water
Water & ice
Salt & water
Excess salt & water

No. of components
1 (H2O)
1 (H2O)
2 (NaCl & H2O)
2 (NaCl & H2O)

No. of phases
1 (liquid)
2 (liquid & solid)
1 (salt solution)
2 (saturated solution &
solid salt)

stable.
A phase diagram describes the equilibrium (lowest energy)
state of a system: the compositions, structures and
amounts of the equilibrium (stable) phases do not vary
with time.

9-3

9-4

9.2 ONE-COMPONENT SYSTEMS

9.3 TWO-COMPONENT SYSTEMS

Unary (1 component) system; e.g. H2O (Fig. 9.2-1).

When

two

components

are

mixed,

they

may

be

completely soluble, partially soluble or insoluble in each


other, or react to form a new compound (Fig. 9.3-1).

Fig. 9.2-1 Schematic phase diagram for H2O.

The stable phase (whether ice, water or steam) depends on


the temperature and pressure.
Phase boundaries separate the 3 different regions over
which each phase is stable; the boundaries also indicate
conditions under which 2 phases coexist.
Under very restricted conditions, all 3 phases coexist (triple
point).

Fig. 9.3-1 Two component mixtures in the liquid state.

The phase diagram identifies the stable phase(s) at any


given pressure and temperature.
9-5

9-6

9.3.1

Solubility in the Solid State

Limited solubility: solute atoms dissolve in the solvent to

A solid solution is a homogeneous single phase formed by


the incorporation of solute atoms (either substituting for
solvent atoms or in interstitial sites) into a host crystal. The
original crystal structure is maintained, with solute atoms

formed,

the Periodic Table to prevent compound formation.

in

3. Valences must be the same.

which the solute atoms substitute

4. Crystal structures must be the same.

for the solvent atoms (Fig. 9.3-2).

9.3.2

Fig. 9.3-2 Substitutional solid solution.

the

interstitial

solvent
solid

atom,

an

solution

is

Specification of Composition

Composition is specified in weight percent (wt%) or atom

If the solute atom is much smaller


than

The necessary (but not sufficient) conditions for unlimited

2. Must belong to the same group or adjacent groups in

of similar size, a substitutional


is

interstitial atoms is always limited (rule 1 below).

1. Difference in atomic radii < 15%.

If the solute and solvent atoms are


solution,

solvent to form a separate new phase. The solubility of

solubility are expressed by the Hume-Rothery rules:

uniformly and randomly dispersed throughout.

solid

an extent only; excess solute may combine with the

percent (at%). In a system containing components A and


B, the composition of A is:

formed, in which the solute atoms

CA = weight (mass) of A x 100

reside in the interstitial spaces

total weight

between solvent atoms (Fig. 9.3-3).

=
!

Fig. 9.3-3 Interstitial solid solution.

CA = no. of moles of A x 100

Unlimited solubility: the components form a substitutional


solid solution when mixed in any amounts.
9-7

weight of A
x 100
weight of A + weight of B

!
!

wt%
wt%

at%

total no. of moles


no. of moles of A
=
x 100 at%
no. of moles of A + no. of moles of B
9-8

9.4 BINARY ISOMORPHOUS SYSTEM

The left edge of the phase diagram represents pure Cu


(100wt%Cu-0wt%Ni), the right edge pure Ni (0wt%Cu-

In this system, there is complete solubility (isomorphous) of


2 components (binary) over the entire composition range

100wt%Ni); at all other points, there are varying amounts


of Cu and Ni.

in both liquid and solid states; e.g. Cu-Ni (Fig. 9.4-1).


There are three distinct phase fields: liquid L, solid solution

!, and 2-phase ! + L. Above the liquidus, the liquid phase


is stable, while below the solidus, the solid phase is stable.
In between these boundaries, liquid and solid coexist.
Melting/solidification

in

pure

Cu

and

Ni

occurs

isothermally at unique melting temperatures. At all other


compositions

containing

both

Cu

and

Ni,

melting/solidification occurs over a temperature range.


The composition of an alloy is usually fixed, so the only
external variable is temperature; the effects of temperature
on the phases present, their compositions and their relative
amounts, are then of interest.
Within single-phase fields (L or !), the temperature may be
varied without changing the equilibrium phase or its
composition.

Fig. 9.4-1 The Cu-Ni phase diagram.

Since most materials are used under constant pressure (i.e.

In 2-phase fields (! + L), changes in temperature are

1 atm), the variables are now the temperature (vertical

accompanied by changes in the relative amounts and

axis) and composition of the alloy (horizontal axis).

compositions of ! and L (details later).


9-9

9-10

9.4.2

Determination of Phases Present

To find the composition of each phase in any 2-phase field:

To identify the phases present for an alloy of composition

1. Extend an isotherm (known as the tie line) at

C0 at temperature T0, locate the point on the phase

temperature T0, through the state point, and across the

diagram

(Fig. 9.4-1)

with the coordinates C0, T0 (this point of

interest is known as the state point) and see which phase


field (single phase L, !, or 2-phase ! + L) the point lies in.
For example, an alloy with composition 60wt%Ni40wt%Cu at 1100C (point A) is a single-phase solid
solution of Cu in Ni, denoted as !, while a 35wt%Ni65wt%Cu alloy at 1250C (point B) consists of both solid !
and liquid phases.

9.4.3

2-phase field.
2. Note the intersections between the tie line and the
phase boundaries on either side.
3. The compositions of the phases are given by the
horizontal coordinates (on the composition axis) of
these intersections.
For example, point B

Determination of Phase Compositions

(Fig 9.4-2)

is located in the 2-phase (! +

L) field; the compositions of the liquid L phase and the


solid ! phase are CL and C!, respectively.

Locate the state point on the phase diagram (Fig. 9.4-1).


If the point lies in a single-phase field (e.g. point A), the
phase composition is identical to the overall alloy
composition C0.
When the point lies in a 2-phase field (e.g. point B), the
phase compositions differ from each other and also from
the original overall alloy composition C0. The composition
of each phase depends on the temperature.
Fig. 9.4-2 Determining phase compositions in any 2-phase field.
9-11

9-12

9.4.4

Determination of Phase Amounts

Locate the state point on the phase diagram (Fig. 9.4-1).


If the point lies in a single-phase field (e.g. point A), the
amount of the single phase is 100%.
If the point lies in a 2-phase field (e.g. point B), the relative
amounts of the phases are calculated by the lever rule:
1. Draw a tie-line through the state point.
2. Locate the overall alloy composition C0 on the tie line.
3. The tie line may be thought of as a lever pivoted at
alloy composition C0

(Fig. 9.4-3 & 9.4-4).

The mass fraction of

Fig. 9.4-3 Determine phase amounts in any 2-phase field using the lever rule.

any phase is proportional to the length of the opposite


lever arm.
Mass fraction = Mass of phase = Opposite lever arm
Total mass

Total lever length

C " CL
!
Mass fraction
of ! phase, !! = M" = o

Mo

C# " C L

ML
= C" # C o
Mo
C" # C L

Mass fraction of L phase,!L = !

Fig. 9.4-4 Schematic illustration of the lever rule.

!
9-13

9-14

Worked Example

Using the lever rule:

Determine the phases, their compositions and relative amounts in a 40


wt%Ni-60wt%Cu alloy at 1300C, 1270C, 1250C, and 1200C.

At 1270C,
Mass fraction of L, L = C" # Co = 50 " 40 = 0.77 (77%)
C" # C L

50 " 37

Mass fraction of !, = Co " CL = 40 " 37 = 0.23 (23%)


!

C# " C L

50 " 37

(or simply, 1- L = 1- 0.77 = 0.23)

At 1250C (Fig. 9.4-6),

Fig. 9.4-5 Tie lines and phase compositions for


a 40 wt%Ni-60wt%Cu alloy at several temperatures.

Temperature

Phase

Composition
(wt%Ni)

Amount
(%)

1300

40

100

Read directly

77

Draw tie-line read from


intersection with liquidus (for
L) and solidus (for !)

37

1270

1250
1200

50

23

32

38

45

62

40

100

Fig. 9.4-6 Tie line at 1250C for determining phase amounts using lever rule.

Remarks

Use lever rule for phase


amounts (Fig. 9.4-6)

Mass fraction of L, L = C" # Co = 45 " 40 = 0.38 (38%)


C" # C L

45 " 32

Mass fraction of !, = Co " CL = 40 " 32 = 0.62 (62%)


!

C# " C L

45 " 32

(or simply, 1- L = 1- 0.38 = 0.62)

As above
Read directly
9-15

9-16

9.4.5

Solidification Under Equilibrium Conditions

Consider the cooling of an alloy of composition 35 wt%Ni65wt% Cu from 1300C (Fig. 9.4-7):
Point a: 100% liquid L
Phase:

L (35 wt% Ni = alloy composition, C0)

Point b: solidification starts at the liquidus temperature


Phases: L (35%Ni) + ! (46%Ni)
Amounts:<100% L + > 0% !
Point c: the proportion of ! increases, the compositions of
L and ! change with temperature, following the
liquidus and solidus respectively
Phases: L (32%Ni) + ! (43%Ni)
Amounts: 73% L

+ 27% !

Point d: solidification is almost complete at the solidus


temperature
Phases: L (24%Ni) + ! (35%Ni)
Amounts: > 0% L + < 100% !
Point e: 100% solid !
Phase:

! (35%Ni = alloy composition, C0)

Fig. 9.4-7 The development of microstructure during the equilibrium


solidification of a 35 wt% Ni-65 wt% Cu alloy.
9-17

9-18

9.4.6

Non-Equilibrium Solidification

During solidification, the composition of ! and L changes


constantly according to the solidus and liquidus as the
temperature drops. These changes are accomplished by
diffusion. Since diffusion in solids is usually very slow, in
most practical situations, cooling is too fast for the
diffusing atoms to establish equilibrium phases and
compositions in the solid state (Fig. 9.4-8).
Point a: 100% liquid L
Phase: L (35%Ni = alloy composition, C0)
Point b: solidification starts at the liquidus temperature
Phases: L (35%Ni) + ! (46%Ni)
Point c: the diffusion in liquid L is assumed to be rapid
enough for the composition of L to follow the
liquidus; however, the composition of the new
layer of ! (40%Ni) formed is different from the
old ! (46%Ni) formed earlier at point b. Since
diffusion is too slow for the old ! to reach the
new ! composition of 40%Ni, the overall !
composition is higher than the solidus and does
not follow the solidus

Fig. 9.4-8 The development of microstructure during the non-equilibrium


solidification of a 35 wt% Ni-65 wt% Cu alloy.
A cored structure is obtained after solidification.

Phases: L (29%Ni) +

! (40%Ni!C!!46%Ni; overall C!"42%Ni)


9-19

9-20

Point d: because of the shift in the solidus, more liquid

9.4.7

Mechanical Properties of Isomorphous Alloys

remains than in the equilibrium case


Phases: L (24%Ni) +

Since isomorphous alloys form a single solid phase at all

! (35%Ni!C! !46%Ni; overall C!"38%Ni)

compositions, each component will experience solidsolution

strengthening

by

additions

of

the

other

component [note that grain-size strengthening is also possible in polycrystals].

Point e: solidification is almost complete


Phases: L (21%Ni) +

! (31%Ni!C!!46%Ni; overall C!"35%Ni)

At some intermediate composition, the strength of the


alloy will be a maximum, but its ductility will exhibit the
opposite trend (Fig. 9.4-9).

Point f: 100% solid !


Phase: # with cored structure as a result of
segregation (i.e. the non-equilibrium
variation in composition within a phase)
(31%Ni!C!!46%Ni; overall C!"35%Ni"Co)
The shift of the solidus from the equilibrium depends on
the cooling rate. Faster cooling results in greater deviation.

Fig. 9.4-9
For the copper-nickel system,
(a) solid solution strengthening
with alloying addition, but
(b) opposite trend in ductility.

The component with the higher melting point segregates


at the centre of the solid while regions between the grains
are rich in the lower-melting point component. This
coring phenomenon causes hot shortness, where regions
around the grain boundaries melt before the equilibrium
solidus temperature of the alloy is reached.
9-21

9-22

9.5 BINARY EUTECTIC SYSTEM

The solvus separates the single phase solid regions (! or ")

In this system, the two components are only partially


soluble in the solid state; e.g. Pb-Sn (Fig. 9.5-1)

from the 2-phase ! + " solid region. The solvus corresponds


to the solubility limit of Pb in Sn and vice versa.
At the eutectic point, E

(Fig. 9.5-1),

!, " and L coexist in

equilibrium. The eutectic point is invariant; i.e. it exists only


at a specific temperature and alloy composition that
cannot be varied. The eutectic point is similar to the triple
point in one-component systems.
When an alloy of eutectic composition, CE, is cooled through
the eutectic temperature, a eutectic reaction occurs:
L

cooling

""""""""#

!"
""""
heating

!+"

During the eutectic reaction, melting/solidification occurs


isothermally at one temperature instead of over a range of
temperatures, as seen in other alloy compositions.
An alloy of eutectic composition, CE,, melts/solidifies at a

Fig. 9.5-1 The Pb-Sn phase diagram.

temperature lower than the melting points of either


Sn is soluble in Pb up to maximum 18.3 wt% at 183C,
while Pb is soluble in Sn up to 2.2 wt%, forming singlephase solid solutions ! and " respectively.

component. The eutectic isotherm represents the lowest


temperature at which the liquid phase exists.
The composition of a hypoeutectic alloy is lower than the

For all other intermediate alloy compositions, a 2-phase


mixture of ! + " exists.
9-23

eutectic composition, CE,, while that of a hypereutectic


alloy is higher than the eutectic composition.
9-24

9.5.1

Solidification of Eutectic Alloy

During the eutectic reaction, Pb and Sn atoms must be

Consider the cooling of an alloy of eutectic composition


61.9 wt% Sn (Fig. 9.5-2):

redistributed via diffusion to form simultaneously a high Pblow Sn ! phase and a low Pb-high Sn " phase.
Alternating layers (lamellae) of ! and " are formed because
such a structure requires Pb and Sn atoms to diffuse only
over relatively short distances (Figs. 9.5-3 and 9.5-4).

Fig. 9.5-3 Simultaneous diffusion of Pb


and Sn atoms to form eutectic lamellae.

Fig. 9.5-4 The eutectic lamellae


structure in the Pb-Sn system.

Fig. 9.5-2 The development of microstructure in a eutectic alloy.

Point h: 100% liquid L


Phase: L (61.9%Sn = eutectic composition, CE)
Eutectic point: eutectic reaction
L (61.9%Sn)

!183C
!!!!"

! (18.3%Sn) + " (97.8%Sn)

Fig. 9.5-5 Schematic


illustration of various
eutectic structures: (a)
lamellar, (b) rodlike,
(c) globular, and
(d) acicular.

Point i: 100% solid (overall composition = CE = 61.9%Sn)


Phases: eutectic ! (18.3%Sn) + " (97.8%Sn) lamellae
9-25

9-26

9.5.2
Consider

Solidification of Off-Eutectic Alloy


the

cooling

of

an

alloy

of

hypoeutectic

Point l: Phases: L (61.9%Sn) + ! (18.3%Sn)


Just below the eutectic temperature, the remaining L
undergoes the eutectic reaction and transforms to the

composition 40 wt% Sn (Fig. 9.5-6):

eutectic ! + " lamellae.


Point m: 100% solid (overall composition = 40%Sn)
Phases: proeutectic ! (18.3%Sn) +
eutectic ! (18.3%Sn) + " (97.8%Sn) lamellae
The ! that forms prior to the eutectic reaction is known as
proeutectic or primary !, to distinguish it from the
eutectic ! lamellae formed during the eutectic reaction.
To determine the relative amount of proeutectic !, the lever
rule is applied in the 2-phase ! + L region at point l, just
above the eutectic temperature, using the composition of

! at one end of the tie-line and the eutectic composition,

Fig. 9.5-6 The development of microstructure in a hypoeutectic alloy.

CE, at the other end.

Point j: 100% liquid L

To determine the relative amount of eutectic ! (i.e. !

Phase: L (40%Sn = overall composition)


Point k: ! solidifies and increases in proportion as the alloy
cools; the compositions of L and ! increase with
decreasing temperature, following the liquidus
and solidus respectively

mixed in the eutectic ! + " lamellae), the lever rule is first


applied in the 2-phase ! + " region at point m, just below
the eutectic temperature, using the composition of ! at
one end of the tie-line and the composition of " at the
other end. This gives the total amount of !; therefore,

Phases: L (47%Sn) + ! (16%Sn)

eutectic ! = total ! proeutectic !.


9-27

9-28

9.5.3

Worked Example

Alloys Without Eutectic Reactions

Determine the phases, their compositions and relative amounts in a 30


wt%Sn-70wt%Pb alloy at 300C, 200C, 184C and 182C and 0C.

Only alloys with compositions exceeding either maximum


solid solubility limits at the eutectic temperature; i.e. 18.3
wt% Sn < C0 < 97.8 wt% Sn, undergo the eutectic reaction
during solidification.
Alloys within the solubility limit at room temperature form
a single-phase solid solution, with microstructure and
solidification characteristics identical to the isomorphous
alloys (Fig. 9.5-7).

Fig. 9.5-7
Solidification in an alloy
of composition C1, which
lies within the roomtemperature solubility limit
This alloy does not undergo
a eutectic reaction.

Total ! @ 182C = 86%, proeutectic ! (from 184C) = 74%;

$ eutectic ! @ 182C = 86-74 = 12%


9-29

9-30

Alloys exceeding the solubility limit at room-temperature,


but within the maximum solubility limit at the eutectic
temperature forms a two-phase microstructure that has a
different morphology (size/shape/distribution) from the
two-phase microstructure containing the eutectic.

Consider the cooling of an alloy of composition C2 wt% Sn


(Fig. 9.5-8):

Point d: 100% liquid L


Phase: L (C2%Sn = overall composition)
Point e: ! solidifies and increases in proportion as the alloy
cools; the compositions of L and ! increase with
decreasing temperature, following the liquidus
and solidus respectively
Phases: L + !
Point f: 100% solid ! (C2%Sn = overall composition)
Point g: solid solubility of ! is exceeded upon crossing the
solvus; " particles precipitate within !. " does not
form lamellae with !, but is dispersed within a
matrix of !
100% solid (overall composition = C2%Sn)
Phases: ! + " (non-lamellae dispersion)
Because the solubility of ! drops with temperature, the
mass fraction of " increases, with the " precipitates
growing in size as the alloy cools. The compositions of !
and " will change with decreasing temperature, following

Fig. 9.5-8 Solidification and precipitation in an alloy of composition C2.


This alloy does not undergo a eutectic reaction.

their respective solvus.


9-31

9-32

9.5.4

Mechanical Properties of Eutectic Alloys

The boundary between the two phases (also known as


interphase boundary) is an obstacle to dislocation motion
the greater the number of boundaries, the greater the
strengthening effect.
The maximum number of interphase boundaries in a
slowly

solidified

eutectic

alloy

occurs

when

the

microstructure is wholly eutectic; thus, the larger the


amount of eutectic, the stronger the alloy (Fig. 9.5-9).

Fig. 9.5-9 The effect of composition and strengthening mechanism


on the strength of lead-tin alloys.
9-33

9-34

9.6 OTHER BINARY SYSTEMS


9.6.1

Invariant Reactions in Binary Systems

Fig. 9.6-2 The Cu-Zn phase diagram.

Fig. 9.6-1 Some important three-phase reactions in binary systems. All invariant
reactions are reversible if cooling/heating is carried out under equilibrium conditions.

9.6.2

Intermediate solid solutions exist over a range of


compositions; i.e., ", $, %, &

(Fig. 9.6-2).

Intermediate com-

pounds have fixed compositions, shown as straight lines

Complex Binary Phase Diagrams

on the phase diagram; e.g. Mg2Ni and MgNi2 (Fig. 9.6-3).

The solid solutions ! and # are called terminal solid


solutions because they appear at the ends (terminus) of
the phase diagram (Fig. 9.6-2).
The components in a binary system may react chemically
to form intermediate phases, which are single phases
formed at alloy compositions away from the ends of the
phase diagram.

Fig. 9.6-3 The Mg-Ni phase diagram.


9-35

9-36

9.7 IRON-IRON CARBIDE (FE-FE3C) SYSTEM

Peritectic reaction at 1493C (important for casting):

%+L

cooling

""""""""#

!"
""""
heating

Eutectic reaction at 1147C (important for cast irons):


L

cooling

""""""""#

!"""""
heating

$ + Fe3C

Eutectoid reaction at 727C (important for the heat


treatment of steels):
cooling

""""""""#

!"""""
heating

! + Fe3C

Ferrite (!): solid solution of carbon in BCC iron (max.


solubility = 0.022 wt% C at 727C); soft and ductile.
Austenite ($): solid solution of carbon in FCC iron (max.
solubility = 2.14 wt% C at 1147C); tough and ductile.
Fig. 9.7-1 Fe-Fe3C phase diagram.

Cementite (Fe3C): intermediate compound of 4 C and 12


Fe atoms (6.67 wt%C); orthorhombic crystal structure with

This system forms the basis for steels and cast irons.
Fe3C is an intermediate compound of iron and carbon, and
is represented by a straight line at the right terminus of the

metallic-covalent bonding, hard and brittle.


Pearlite: 2-phase mixture of alternating layers (lamellae) of
ferrite and cementite (! + Fe3C) formed simultaneously

phase diagram.
Although Fe3C forms one terminus of the phase diagram,
by convention and for convenience, the composition is still
measured in terms of wt% C; 6.67 wt% C = 100 wt% Fe3C.
9-37

during the eutectoid reaction.


The terms ferrite, austenite, cementite and pearlite are
used in the Fe-Fe3C system only.
9-38

9.7.1

Transformations in Eutectoid Steels

Consider the cooling of a eutectoid steel (0.76 wt% C)


from 800C (Fig. 9.7-2).

During the eutectoid reaction, C atoms must be


redistributed by diffusion such that $ (0.76%C) is
transformed to low-C ! (0.022%C) and high-C Fe3C
(6.67%C) simultaneously.
Alternating layers (lamellae) of ! and Fe3C are formed
because such a structure requires C atoms to diffuse only
over relatively short distances (Figs. 9.7-3 and 9.7-4).

Fig. 9.7-3 Schematic


representation of
pearlite formation
(arrows indicate
direction of
C diffusion).

Fig. 9.7-2 The development of microstructure in a eutectoid steel.

Point a: 100% $ (0.76%C)


Eutectoid point: eutectoid reaction:

$ (0.76%C)

727C
!!
!!!"

! (0.022%C) + Fe3C (6.67%C)

Fig. 9.7-4 Pearlite in


eutectoid steel, consisting
of alternating layers of #
(light phase) and Fe3C (thin
layers that look dark).

Point b: 100% pearlite


eutectoid ! (0.022%C) + Fe3C (6.67%C) lamellae
9-39

9-40

9.7.2

Transformations in Hypoeutectoid Steels

Consider the cooling of a hypoeutectoid steel of

9.7.3

Transformations in Hypereutectoid Steels

Consider the cooling of a hypereutectoid steel of


composition C1 (> 0.76 wt% C) from 915C.

composition C0 (< 0.76 wt% C) from 875C.


Point c: all $ (compositionC0)

Point g: all $ (composition C1)

Point d: proeutectoid !

Point h: proeutectoid Fe3C

forms along $ boundaries

forms along $ boundaries

and increases in pro-

and increases in

portion as the steel

proportion as the steel

cools; the compositions

cools; the composition of

of ! and $ increase with

Fe3C remains constant but

decreasing temperature,

that of $ decreases with

according to the solvus.

temperature, according to

Phases: ! (0.022%C) +

the solvus.

$ (0.3%C)

Phases: $ (1.0%C) +
Fe3C (6.67%C)

Point e: Phases: ! (0.022%C) + $ (0.76%C)


Just below the eutectoid temperature, the remaining $

Just below the eutectoid temperature, the remaining $

(0.76%C) undergoes the eutectoid reaction and transforms

(0.76%C) undergoes the eutectoid reaction and transforms

to the eutectoid ! + Fe3C pearlite.

to the eutectoid ! + Fe3C pearlite.

Point f: Phases: proeutectoid ! (0.022%C) +


eutectoid ! (0.022%C) +
Fe3C (6.67%C)

Point i: Phases: proeutectoid Fe3C (6.67%C) +


eutectoid ! (0.022%C) +

pearlite

Fe3C (6.67%C)
9-41

pearlite
9-42

9.7.4

Mechanical Properties of Steels

The interphase boundary between ferrite and cementite is


an obstacle to dislocation movement the greater the
number of boundaries, the greater the strengthening.
In addition, cementite, which contains complex metalliccovalent bonding, does not undergo plastic deformation at
room temperatures, such that dislocation movement can
occur only within the ferrite phase. Therefore, the more
cementite a steel contains (i.e. the higher the carbon
content), the higher its strength

(Fig. 9.7-7).

Fig. 9.7-7 (a) Variation of strength and hardness, (b) ductility and impact energy
with carbon content for plain carbon steels with fine pearlite structure.
9-43

CHAPTER 10

10.1

KINETICS OF PHASE
TRANSFORMATIONS

Phase

PHASE TRANSFORMATIONS
transformations

are

not

instantaneous.

The

transformation from one phase to another involves a


change in composition, crystal structure and/or the
number of phases present. This is achieved through the

10.1

10.2

10.3

rearrangement of atoms (often via diffusion), which

PHASE TRANSFORMATIONS

requires a finite amount of time.

10.1.1

The Driving Force for


Transformation

10.1.2

Nucleation

10.1.3

Growth

diagrams are obtained only when the temperature is

10.1.4

Kinetics of Phase
Transformation

changed extremely slowly, such that a state of equilibrium

The microstructures characterized by the equilibrium phase

is maintained at all times. The phase diagrams do not


indicate the time required for equilibrium microstructures

ISOTHERMAL TRANSFORMATION
DIAGRAMS
10.2.1

Transformations in Eutectoid
Steel

10.2.2

Hypoeutectoid and
Hypereutectoid Steels

to develop.
Under non-equilibrium conditions, the resultant microstructure depends on the rate of heating/cooling and the
actual temperature at which phase transformations occur.

CONTINUOUS COOLING
TRANSFORMATION DIAGRAMS
10.3.1

These

factors

also

determine

the

rate

of

phase

transformation (and hence, the time required).

Continuous Cooling in Steels

The desired microstructure may be therefore be tailored by


controlling the rate (kinetics) of phase transformations.

10-1

10-2

10.1.1

The Driving Force for Transformation

At the equilibrium transformation temperature, TE = 0C,


the free energy of ice and water are equal (Gice = Gwater "

The thermodynamic state of a system is defined by the


Gibbs free energy, G, which is a measure of the internal

!G = 0). A change from water to ice or vice versa would


not lower G, so no spontaneous change occurs.

energy, as well as the randomness in the system.


Equilibrium is a state of no spontaneous change, at which
G is a minimum.

When T < TE, Gice < Gwater, so ice is the stable form. The
system can lower its free energy by changing water into
ice (Gice - Gwater < 0 " !G < 0). The decrease in free energy,

Phase transformations are only possible if such changes


lower G, i.e. !G < 0.

!G, becomes larger and larger as T falls further and further


below TE. TE T = !T is known as the undercooling.
!G may be thought of as the driving force for the

Consider the solidification of water into ice (Fig. 10.1-1).

transformation. The greater the degree of undercooling,


!T, the larger the magnitude of !G, and the higher the
driving force.
Generally, a phase transformation begins with the
nucleation of the new phase within the parent phase,
followed by the growth of the new phase.

Fig. 10.1-1 Gibbs free energy as a function of temperature for ice and water.
10-3

10-4

10.1.2

Nucleation

However, the random clustering of atoms requires the


local diffusion of atoms, the rate of which decreases with

Consider the nucleation of a solid in a liquid during


solidification. The atoms in a liquid are in a state of
continual random movement. From time to time, a small
group of atoms will, purely by chance, come together to
form a tiny crystal nucleus (Fig. 10.1-2).

temperature.
The net nucleation rate is therefore a balance between the
ease of nucleation and atomic mobility

(Fig. 10.1-3).

The

maximum nucleation rate occurs in the temperature range


where the driving force for nucleation and diffusion are
both significant.

Fig. 10.1-2 The random clustering of neighbouring atoms to form a crystal nucleus.

The nucleus must be of critical size, r*, or larger, in order


to remain stable and grow; nuclei smaller than r* would
simply redissolve back into the liquid.
An increase in the degree of undercooling, !T, decreases
the critical nucleus size, r*, required. Since there is a higher
probability of atoms randomly clustering in small groups
rather than large ones, nucleation becomes easier and the
nucleation rate is faster with larger !T.

Fig. 10.1-3 Variation of the net nucleation rate with temperature.

Nucleation occurs preferentially at sites such as walls of


containers or suspended impurities in a liquid. In solid-tosolid phase transformations, preferential nucleation sites
include grain boundaries, dislocations, phase boundaries,
and the surfaces of impurities and precipitates.

10-5

10-6

10.1.3

Growth

The progress of a phase transformation with time at any

Once a nucleus of critical size or larger forms, spontaneous


and sustained growth of the nucleus occurs

(Fig. 10.1-4).

given temperature is described by a sigmoidal (S-shaped)


curve (Fig. 10.1-6).

Growth involves the transport of atoms to the nucleus and


the rearrangement of these atoms into the crystal structure
of the nucleus. These processes are diffusion-controlled,

Fig. 10.1-6 The fraction of


transformed material as a
function of time at any
given temperature.

and the growth rate increases with temperature.


Fig. 10.1-4 Growth
of a stable nucleus.

10.1.4

Kinetics of Phase Transformation

The overall rate (or kinetics) of phase transformation


depends on both nucleation and growth rates (Fig. 10.1-5).

Characteristics of phase transformation:


1. An incubation period is required for nucleation. There is
no measurable phase transformation during this period.
2. Transformation is slow initially as nuclei form.
3. Once nucleated, the new phase begins to grow at the
expense of the parent phase, and there is a rapid
increase in the amount of new phase present.
4. The growth rate of the new phase decreases eventually

Fig. 10.1-5 (a) Variation of rate of phase transformation with temperature.


1 ).
(b) Corresponding time-temperature-transformation curve or C-curve (Note: time, t # Rate

because of the depletion of solute atoms or physical

10-7

10-8

impingement of the growing phase.

10.2

I SOTHERMAL T RANSFORMATION D IAGRAMS

Equilibrium phase diagrams define only the microstructures


that develop under equilibrium conditions. In practical
situations, the rate of heating/cooling and the actual
temperature of transformation (undercooling) determine
the resultant microstructure, which may be different from
that characterized by the equilibrium phase diagrams.

10.2.1

Transformations in Eutectoid Steel

Recall the eutectoid reaction in steel:

!727C
!!!!"

" + Fe3C

The upper (high-temperature) portion of the TTT diagram


for eutectoid steels (0.76 wt% C)

(Figs. 10.2-2 & 10.2-3)

shows

the decomposition of austenite to pearlite with time, when


the steel is quenched (cooled rapidly) from the austenite

Isothermal transformation (IT) diagrams, or timetemperature-transformation (TTT) diagrams, show the


progress of transformation with time, and the final

phase field and held isothermally at various temperatures


below the eutectoid transformation temperature.

microstructure (including non-equilibrium phases).


TTT diagrams are derived from a series of experimental
sigmoidal curves at different temperatures

(Fig. 10.2-1);

each

diagram is valid for one alloy composition only.

Fig. 10.2-1
Construction of a TTT
diagram from a series
ofexperimental
sigmoidal curves at
different temperatures.

Fig. 10.2-2 Isothermal transformation of austenite to pearlite


as a function of time and temperature.
10-9

10-10

Above the eutectoid temperature, only austenite will exist;

Since the transformation to pearlite during the eutectoid

transformation from austenite to pearlite will occur only if

reaction involves redistribution of carbon atoms via

the steel is cooled below the eutectoid temperature.

diffusion, the morphology (size and shape) of pearlite

The solid curves define the onset and completion of the


transformation of austenite to pearlite, while the dashed
curve represents 50% completion (Fig. 10.2-3).

depends on the actual temperature of transformation.


Near the eutectoid temperature, slow nucleation (few
grains) but fast diffusion (carbon atoms diffuse over long

To the left of the start curve, only austenite (unstable) will


be present; to the right of the finish curve, only pearlite
will exist. In between, the austenite is in the process of

distances quickly) produces coarse pearlite, consisting of


thick

layers

of

ferrite

and

cementite

with

large

interlamellar spacing (Fig 10.2-4a).


As undercooling increases, incubation becomes faster and

transforming to pearlite, thus both will be present.

transformation to pearlite occurs earlier. Fast nucleation


(many grains) but slow diffusion (diffuse short distances
only) yields fine pearlite, with thin layers of ferrite and
cementite, and small interlamellar spacing (Fig 10.2-4b).

Fig. 10.2-4
(a) Coarse pearlite;
(b) fine pearlite.

Fig. 10.2-3 The isothermal transformation of austenite to pearlite,


showing the development of microstructures.
10-11

10-12

Further undercooling produces bainite

(Fig. 10.2-5),

which is

The distribution and morphology (size and shape) of the

an even finer distribution of ferrite and cementite than in

ferrite and cementite phases in pearlite and bainite differ

fine pearlite. Unlike pearlite, which is lamellar, bainite

due to differences in the temperature at which their

consists of extremely fine, elongated cementite particles

transformations occur, which affect the nucleation and

between ferrite plates or needles (Fig. 10.2-6).

growth kinetics of ferrite and cementite. However, both


pearlite and bainite are still physical mixtures of ferrite and
cementite, as indicated in the equilibrium phase diagram.
Once a portion of austenite has transformed to pearlite or
bainite, the pearlite or bainite remains stable upon further
Fig. 10.2-5 The TTT
diagram for eutectoid
steel showing the
transformation of
austenite to pearlite and
austenite to bainite.

cooling

(Fig. 10.2-7).

It is not possible for the pearlite to

transform directly into bainite, or vice versa, without first


reheating to form austenite. Only austenite, which is
unstable below the eutectoid temperature, is able to
transform to other, more stable, phases upon cooling.

Fig. 10.2-7 Coarse pearlite remains stable upon cooling.

Fig. 10.2-6 Bainite formed at (a) 450C, and (b) 260C.


10-13

10-14

Immediately

prior

to

the

eutectoid

transformation,

Martensite has a BCT (body-centred tetragonal) structure

austenite contains 0.76 wt% C, with carbon atoms

because carbon atoms remain trapped in the octahedral

dissolved in the octahedral interstitial sites of the FCC

interstitial positions of the original FCC austenite and distort

austenite lattice. The transformation of austenite to pearlite

the BCC structure of ferrite into BCT (Figs. 10.2-9 and

10.2-10).

or bainite requires the diffusion of carbon atoms to form


low-carbon ferrite (0.022 wt% C) and high-carbon
cementite (6.67 wt% C) simultaneously.
When undercooling becomes so great (isothermal transformation temperature below Ms) that carbon diffusion
cannot occur, the diffusive transformation to ferrite and
cementite is suppressed, and a diffusionless transformation
to non-eqiuilibrium martensite takes place (Fig. 10.2-8).

Fig. 10.2-9 The relationship between FCC


austenite and BCT martensite (c>a).

Fig. 10.2-10 The BCT unit cell of


martensite produced by the
distortion of the BCC unit cell by
interstitial carbon atoms.

Since carbon atoms are trapped in the BCT structure,


martensite has the same composition as its parent
austenite (0.76 wt% C in the case of eutectoid steel).
Unlike the transformation to pearlite and bainite, the
amount of austenite that transforms to martensite depends
solely on temperature, not time. Martensite begins to form
only when austenite is quenched to temperature Ms, while
the M50 isotherm indicates the temperature at which 50%
of austenite will transform to martensite (Fig. 10.2-11).

Fig. 10.2-8 The complete TTT diagram for eutectoid steel.


10-15

10-16

The Ms, M50, M90 (or MF) temperatures are not fixed, but

Martensite is the hardest phase in steel

(Fig. 10.2-12)

as a result

decrease with increases in the carbon content of the steel.

of the lattice distortion caused by the trapped carbon

In some steels, complete (100%) transformation to

atoms. However, it tends to be brittle.

martensite may require a subzero quench to below MF


10.2-11);

(Fig.

quenching to temperatures between Ms and MF in

these steels result in untransformed austenite remaining


amidst the martensite as retained austenite.
Fig. 10.2-12
Hardness of
martensite and
pearlite as a function
of carbon content.

Bainite is generally harder and stronger than pearlite


10.2-13)

due to its much finer distribution

(Fig.

(Sec. 6.3-5)

of

cementite particles within ferrite.


Fig. 10.2-11 The amount of martensite formed when steel is quenched
to room temperature depends on Ms and MF.

Martensite is metastable; i.e. it is stable with time at room


temperature, but upon reheating (in a process known as
tempering), martensite will decompose to the even more

Fig. 10.2-13 Hardness and


strength of bainite and
pearlite as a function of
isothermal transformation
temperature.

stable phases of ferrite and cementite, known as tempered


martensite.
10-17

10-18

Worked Example

(a) At 350C, austenite transforms to bainite, beginning after about

Determine the final microstructure (in terms of microconstituents

10 s, and is completed at about 500 s. Therefore, by 104 s, the

present and approximate percentages) of a small specimen of

specimen is 100% bainite, which does not transform further, even

eutectoid steel that has been heated to its austenite phase, and then

though the final quenching passes through the martensite region

subjected to the following treatments.

of the TTT diagram.


4

(a) Rapidly cool to 350C, hold for 10 s, and quench to room


temperature.
(b) Rapidly cool to 250C, hold for 100 s, and quench to room
temperature.
(c) Rapidly cool to 650C, hold for 20 s, rapidly cool to 400C, hold
for 103 s, and quench to room temperature.

(b) At 250C, after 100 s, bainite transformation has not yet begun, so
the specimen is still 100% austenite. Quenching through the
martensite region of the TTT diagram, more and more austenite
transforms to

martensite

as temperature

drops. At room

temperature, most of this transformation is completed (past the


M90 line), so that the final microstructure is roughly 100%
martensite.

(c) At 650C, austenite transforms to pearlite, beginning after about 7


s. After 20 s, only about 50% of the specimen has transformed to
pearlite, which does not transform further upon cooling.
For the remaining 50% austenite, rapid cooling to 400C is fast
enough such that virtually no transformation occurs during
cooling, even though cooling occurs through pearlite and bainite
regions of the TTT diagram.
At 400C, timing for the remaining 50% austenite is reset to zero,
because at each temperature, the kinetics (nucleation/growth
behavioiur with time) of austenite transformation is unique.
Austenite transforms to bainite, beginning after about 4 s, and is
completed at about 100 s. Therefore by 103s, the remaining 50%
of the specimen is 100% bainite, which does not transform further
upon cooling. Final microstructure: 50% pearlite + 50% bainite.
10-19

10-20

10.2.2

Hypoeutectoid and Hypereutectoid Steels

Since the transformation of austenite to the proeutectoid


phases requires the diffusion of carbon atoms, the amount

When hyporeutectoid and hypereutectoid steels are

of proeutectoid phase that forms depends on the

cooled, austenite first transforms to proeutectoid ferrite

undercooling. More proeutectoid phase is formed when

and

the isothermal transformation temperature is high.

cementite,

respectively,

before

the

eutectoid

transformation to pearlite. The formation of proeutectoid


ferrite or cementite adds an extra curve to the TTT
diagram above the pearlite nose (Figs 10.2-14 and 10.2-15).

When the undercooling is so large that bainite or


martensite is obtained, no proeutectoid phase is formed.

Fig. 10.2-14 TTT diagram for a hypoeutectoid steel (0.5 wt% C)


and its corresponding portion of the Fe-Fe3C phase diagram,
showing an extra curve due to proeutectoid ferrite formation.

Fig. 10.2-15 TTT diagram for a hypereutectoid steel (1.13 wt% C).

10-21

10-22

10.3

10.3.1

CONTINUOUS COOLING
T R A N S F O R M A T I O N D I A G RA M S

Continuous Cooling in Steels

Because the temperature is constantly decreasing when


Strictly speaking, TTT diagrams are valid only for

the steel is cooled continuously, CCT curves are shifted to

isothermal transformations, in which the material is

longer times and lower temperatures when compared to

assumed to be instantaneously quenched from a single-

TTT diagrams (Fig. 10.3-1).

phase field to a specific temperature and held isothermally


for various times to produce different amounts of the
transformed products.
In practice, transformations occur during continuous
cooling to room temperature, in which the temperature is
constantly changing as the transformation progresses.
Therefore, it is more accurate to use continuous-coolingtransformation (CCT) diagrams.
CCT diagrams are derived from a series of experimental
sigmoidal curves at different cooling rates. Like the TTT
diagrams, each CCT diagram is valid for one alloy
composition only.

Fig. 10.3-1 Comparison of TTT and CCT diagrams for eutectoid steel.

There is no longer any bainite formation when cooling


continuously at any given constant cooling rate because all
the austenite will have transformed to pearlite by the time
bainite transformation becomes possible.
10-23

10-24

The transformation from austenite to pearlite occurs over a


range of temperatures rather than at a single temperature.
Slow cooling produces coarse pearlite while moderately
fast cooling produces fine pearlite (Fig. 10.3-2).

There is a range of cooling rates

(between 35 to 140C/s in Fig. 10.3-3)

over which both pearlite and martensite are produced.


When cooling within this range, there is insufficient time to
complete the austenite-to-pearlite transformation (since
the cooling curves does not reach the pearlite finish curve
on the right). The remaining austenite that does not
transform to pearlite at the higher temperatures will begin
to transform to martensite when the Ms temperature is
reached.

Fig. 10.3-2 Different cooling rates produces different microstructures.

There is a critical cooling rate

(140C in Fig. 10.3-3)

that will just

miss the nose at which the pearlite transformation


begins. It represents the minimum cooling rate that will
avoid pearlite formation and produce a fully martensitic

Fig. 10.3-3 Critical cooling rate for a fully martensitic structure.

structure (if cooled below MF).


10-25

10-26

Worked Example
Determine the final microstructure of a small specimen of eutectoid
steel that has been cooled from its austenite phase to room
temperature at the following rates: (a) 1C/s, (b) 20C/s, (c) 50C/s,
and (d) 175C/s.
(a) From Fig. 10.3-3 1C/s lies to the far right of the CCT diagram,
where

nucleation

is

slow,

but

growth

is

fast;

the

final

microstructure is likely to be coarse pearlite.


(b) 20C/s lies closer to the cooling curve of 35C/s, which is the
fastest cooling rate possible in order to still obtain 100% pearlite.
For a faster cooling rate, nucleation is faster, but growth is slower;
the final microstructure is likely to be fine pearlite.
(c) 50/s lies in the region of split transformation on the CCT diagram
at high temperatures, pearlite nucleation and growth occurs, but
the rapid drop in temperature means that the transformation from
austenite to pearlite cannot be completed, because diffusion
becomes too slow at lower temperatures. The remaining austenite
transforms to martensite when cooled through the martensite
region of the CCT diagram. Final microstructure is a mixture of
fine pearlite and martensite.
(d) 175C/s is faster than the critical cooling rate of 135C/s, at which
there is no time for diffusion to occur, so the final microstructure is
approximately 100% martensite, assuming MF (or at least M90) is
reached.

10-27

You might also like