You are on page 1of 128

AUSTENITE DECOMPOSITION OF L O W C A R B O N H I G H

STRENGTH STEELS DURING CONTINUOUS COOLING

By
Plamen Petkov

Dip. Eng., University of Chemical Technologies and Metallurgy


Sofia, Bulgaria, 1986

A THESIS SUBMITED IN PARTIAL F U L F I L L M E N T OF


THE REQUIREMENT FOR THE D E G R E E OF
M A S T E R OF APPLIED SCIENCE

in
THE F A C U L T Y OF G R A D U A T E STUDIES
D E P A R T M E N T OF M E T A L S A N D M A T E R I A L S ENGINEERING

We accept this thesis as conforming


to the required standard

THE UNIVERSITY OF BRITISH C O L U M B I A

March 2004
Plamen Petkov, 2004
Library Authorization

In presenting this thesis in partial fulfillment of the requirements for an advanced


degree at the University of British Columbia, I agree that the Library shall make it
freely available for reference and study. I further agree that permission for
extensive copying of this thesis for scholarly purposes may be granted by the
head of my department or by his or her representatives. It is understood that
copying or publication of this thesis for financiafgain shall not be allowed without
my written permission.

Plamen Petkov 07/04/2004


Name of Author (please print) Date (dd/mm/yyyy)

Title of Thesis: Austenite Decomposition of Low Carbon High Strength Steels


During Continuous Cooling

Degree: MASc. Year: 2004

Department of Metals and Materials Engineering


The University of British Columbia
Vancouver, BC Canada
ABSTRACT

The process of transformation on the run-out table of a hot strip mill critically

determines the final micro structure and thus, the hot rolled steel properties. The austenite

decomposition kinetics of a high strength, low alloy steel (HSLA-65) microalloyed with

niobium and a molybdenum bearing dual phase steel (DP 600) have been investigated and

quantified with dilatometry during continuous cooling using a Gleeble 3500 thermo-

mechanical simulator. The effects of different variables, such as austenite grain size and

cooling rate, were quantified. It has been shown that increasing the cooling rate and the

austenite grain size results in a decrease of transformation start temperatures. Evolution of

the resulting microstructure has been investigated; polygonal ferrite fraction and ferrite

grain size were determined. Accelerated cooling and smaller austenite grain size refine the

final ferrite grain size.

Further, austenite decomposition from pancaked austenite has been investigated.

Deformation affects austenite decomposition by introducing additional nucleation sites and

thus accelerating the nucleation rate. The retained strain showed a marked influence on the

increase in transformation start temperature, as well as the increase in ferrite fraction and

the ferrite grain refinement.

Two models were applied to describe the process of austenite-to-polygonal ferrite

transformation. The first model for the transformation kinetics is divided into two

submodels. The transformation start temperature is described with a combined nucleation

and early growth model. Subsequent growth can be described by the Avrami equation (or

J M A K model) which is adapted to non-isothermal transformation by using the Scheil

n
equation of additivity with an Avrami exponent n=0.9 and rate constant b which depends

exponentially on temperature; the effect of austenite grain size on transformation kinetics

can be captured with a suitable grain size exponent m. The second model is related to the

ferrite grain size which can be expressed as a function of initial austenite grain size and

transformation start temperature. A n effective grain size, d eff = d exp(--), was used to
y

incorporate the combined effect of austenite grain size and retained strain on the

transformation start, transformation kinetics and ferrite grain size models.

111
T A B L E OF CONTENTS

ABSTRACT ii

T A B L E OF CONTENTS iv

LIST OF FIGURES viii

LIST OF TABLES xiv

LIST OF SYMBOLS xv

ACKNOWLEDGMENT xviii

CHAPTER 1- INTRODUCTION 1

CHAPTER 2- L I T E R A T U R E R E V I E W 4

2.1 Resulting Microstructures from Austenite Decomposition 4

2.2 Allotriomorph Ferrite Reaction 7

2.2.1 Nucleation of Proeutectoid Ferrite 8

2.2.2 Growth 10

2.2.3 Impingement 13

2.3 Factors affecting Austenite Decomposition 14

2.3.1 Cooling Rate 14

2.3.2 Prior Austenite Microstructure 17

2.3.2.1 Austenite Grain Size 17

2.3.2.2 Effect of Retained Strain 17

2.3.2.3 Effective Grain Size 20

iv
2.3.3 Chemical Composition 21

2.3.3.1 Carbon 22

2.3.3.2 Manganese 22

2.3.3.3 Molybdenum 23

2.3.3.4 Niobium 25

2.4 Modeling of Austenite-to-Ferrite Transformation 27

2.4.1 Kinetics of Isothermal Transformation-JMAK Approach 27

2.4.2 Application of Isothermal Transformation Kinetics to Non-Isothermal

Conditions 28

2.4.3 Overall Transformation Model Applicable to Run-Out Table Simulated

Conditions 31

2.4.3.1 Transformation Start Temperature 31

2.4.3.2 J M A K Equation and Additivity Rule Applications 33

2.4.3.3 Polygonal Ferrite Grain Size 34

C H A P T E R 3- S C O P E A N D O B J E C T I V E S 36

C H A P T E R 4- E X P E R I M E N T A L 37

4.1 Materials 37

4.2 Experimental Equipment 38

4.3 Experimental Procedure 38

4.3.1 Austenite Grain Size 38

4.3.2 Continuous Cooling Transformation 40

4.3.3 Double Hit Tests 42

4.3.4 Deformation-Transformation Experiments 44

v
4.4 Microstructural Analysis and Techniques 45

4.5 Hardness Measurement 47

C H A P T E R 5- R E S U L T S 48

5.1 Initial Grain Size Conditions 48

5.2 CCT Results 51

5.2.1 Effect of Cooling Rate 51

5.2.2 Effect of Initial Austenite Grain Size 53

5.2.3 Effect of Chemical Composition : 55

5.2.4 Transformation Start and Finish Temperatures 57

5.3 Microstructural Evaluation 59

5.3.1 Resulting Microstructure in HSLA-65 Steel 59

5.3.2 Resulting Microstructure in DP 600 Steel 62

5.3.3 Polygonal Ferrite Grain Size in HSLA-65 and DP 600 Steels 64

5.3.4 Hardness 66

5.4 Effect of Retained Strain oh Continuous Cooling Transformation 68

5.4.1 Double Hit Compression Test 68

5.4.2 Effect of Retained Strain on Transformation Kinetics 71

5.4.3 Microstructure Resulting from Pancaked Austenite 75

5.4.3.1 DP 600 Steel 75

5.4.3.2 HSLA-65 Steel 78

5.4.4 Polygonal Ferrite Grain Size 80

5.4.5 Effect of Deformation on Hardness 82

C H A P T E R 6- M O D E L I N G 85

vi
6.1 Kinetics of Transformation from Austenite to Polygonal Ferrite 85

6.1.1 Transformation Start Temperature 85

6.1.2 Kinetics of Austenite-to-Polygonal Ferrite Transformation 89

6.2 Polygonal Ferrite Grain Size 97

C H A P T E R 7- C O N C L U S I O N S A N D F U T U R E W O R K 102

7.1 Conclusions 102

7.2 Future Work 103

REFERENCES 105

vii
LIST OF FIGURES

Figure Page

1.1 Schematic diagram of a hot strip mill 1

2.1 CCT diagram indicating the resulting microstructure 5

2.2 Fe - Fe C phase diagram showing equilibrium concentrations for


3

proeuctoid ferrite formation 7

2.3 Diffusion field of carbon during the growth of proeutectoid ferrite 8

2.4 Heterogeneous nucleation sites in fully recrystallized austenite 9

2.5 Thickening and lengthening of grain boundary allotriomorph 10

2.6 Typical plot of allotriomorph half-length and half-thickness as

a function of the square root of the reaction time 11

2.7 Schematic representation of continuous cooling transformation in

dual phase steel resulting in polygonal ferrite-martensite microstructure

in as-rolled product 16

2.8 The shape change of an austenite grain by reduction p 18

2.9 Ratio of grain boundary surface area as a function of rolling reduction 19

2.10 Effect of Mo on kinetics of austenite to ferrite transformation. 23

2.11 Effect of Mo on ratio of maximum to minimum cooling rates

producing martensite-ferrite microstructure on run-out table 25

2.12 Effect of Nb content on transformation start temperature 26

2.13 Sigmoidal transformation kinetics 28

2.14 Continuous cooling transformation represented as the sum

viii
of short-time isothermal steps 29

4.1 Schematic thermal path for obtaining the initial austenite grain sizes. 39

4.2 Schematic diagram of CCT test. 40

4.3 Schematic schedule of CCT tests. 40

4.4 Example of calculation of fraction transformed during continuous

cooling transformation. 42

4.5 Example of a double hit test for H S L A steel with d =49 um, Tdef=900 C
r
0

and load interruption at strain of 0.2. 43

4.6 Samples used for deformation and subsequent transformation. 44

4.7 Schematic schedule of deformation-continuous cooling transformation

tests. 45

5.1 Austenite grain growth of DP 600 steel (holding time of 120 seconds

at each temperature) and HSLA-65 steel (holding time of 60 seconds

at each temperature). 48

5.2 Austenite grain size of HSLA-65 steel with reheating temperature

1200 C and holding time of 60 seconds.


0
49

5.3 Austenite decomposition kinetics of DP 600 steel with austenite grain

size of 16 urn at different cooling rates. 52

5.4 Austenite decomposition kinetics of the HSLA-65 steel with austenite

grain size of 49 um at different cooling rates. 53

5.5 Comparison of austenite decomposition kinetics of DP 600 steel for

16, 24 and 32 um austenite grain size at a cooling rate of about 20C/s. 54

5.6 Comparison of austenite decomposition kinetics of HSLA-65 for

ix
16, 41 and 108 urn austenite grain size at a cooling rate of about 50C/s. 55

5.7 Comparison between austenite decomposition kinetics of DP 600 and

HSLA-Nb steels. 56

5.8 Transformation start temperature vs. cooling rate for DP 600 steel. 58

5.9 Transformation start temperature vs. cooling rate for HSLA-65 steel. 58

5.10 Microstructure after continuous cooling test for HSLA-65 having

d =49 um at cooling rates: a) 1 C/s; b) 18 C/s; c) 72 C/s.


r
0 0 0
59

5.11 Microstructure after continuous cooling test for HSLA-65 having

dy=19 um at cooling rates: a) lC/s; b) 13 C/s; c) 6 3 C/s.


0 0
60

5.12 Effect of austenite grain size and cooling rate on polygonal ferrite

fraction for HSLA-65 steel. 61

5.13 Microstructure after continuous cooling test for DP 600 steel having

d =16 um at cooling rates: a) 1 C/s; b) 23 C/s; c) 67 C/s.


r 62

5.14 Effect of austenite grain size and cooling rate on polygonal ferrite

fraction for DP 600 steel. 64

5.15 Polygonal ferrite grain size vs. cooling rate for DP 600 and

HSLA-65 steels. 65

5.16 Effect of cooling rate and austenite grain size on hardness

(HSLA-65 steel). 66

5.17 Transformation start temperature vs. hardness (HSLA-65 steel). 67

5.18 Double hit test for HSLA-65 steel (interruption time t. = 5 s,

d =19 um, strain rate=l/s, T


r rfe/ =900C). 68

5.19 Double hit test for DP 600 steel (interruption time t,. = 5 s, d =16 um,
T
strain rate=l/s, T def =900 C ).
u
69

5.20 Relative softening (measured by offset method) for DP 600 steel

(interruption time t .=2, 5, and 20 s, strain rate=l/s, prestrain=0.5 ).


( 70

5.21 Relative softening for DP 600 and HSLA-65 steels (interruption time

t .=2, 5, and 20 s; prestrain =0.2).


( 71

5.22 Effect of cooling rate on transformation kinetics for DP 600 steel

with constant deformation. 72

5.23 Effect of retained strain on transformation kinetics for HSLA-65 steel

(d =l9 um; cooling rate=60 C/s).


y
0
73

5.24 Effect of retained strain on transformation start temperature for

HSLA-65 steel (d =49 um).


y 74

5.25 Transformation start temperature as a function of cooling rates

and retained strain for DP 600 steel with d =32 um.


T 75

5.26 Effect of residual strain on polygonal ferrite evolution of DP 600 steel

with d =32 um: a) CR=49C/s, 8=0; b) CR=60C/s, s=0.2;


T

c) CR=60C/s, e=0.5;d) CR=62C/s, =0.7. 76

5.27 Polygonal ferrite fraction vs. cooling rates for DP 600 steel

with dy=32 um. 77

5.28 Polygonal ferrite fraction vs. cooling rates for DP 600 steel

with d =16 um.


T 78

5.29 Effect of retained strain on ferrite grain size of HSLA-65 with

dy=19 um, CR = 60C/s: a) s=0; b) =0.2; c) =0.5. 79

5.30 Effect of deformation on polygonal ferrite fraction for

xi
HSLA-65 (o>49 um). 80

5.31 Ferrite grain refinement in HSLA-65 steel with d = 19 um.


r 81

5.32 Ferrite grain refinement in DP 600 steel with d = 16 um.


T 82

5.33 Effect of deformation and cooling rate on hardness for DP 600 steel

with d = 32 (am.
T 83

5.34 Transformation start vs. hardness for DP 600 for strain and

strain-free cases. 84

6.1 Model of transformation start temperature for DP 600 steel. 86

6.2 Model of transformation start temperature for HSLA-65 steel. 87

6.3 Transformation start temperature model for DP 600 steel with

incorporated retained strain. 88

6.4 Transformation start temperature model for HSLA-65 steel including

deformation. 89

6.5 Expression of lnb as a linear function of temperature for HSLA-65

with n=0.9 and austenite grain of 19 um at different cooling rates. 91

6.6 Derivation of parameters for Umemoto equation. 92

6.7 Application of Umemoto grain size modified J M A K model

for DP 600 steel. 93

6.8 Application of Umemoto grain size modified J M A K model

for HSLA-65 steel. 94

6.9 Application of J M A K model incorporating the effective austenite

grain for DP 600 steel (d =32 um and cooling rate 5 C/s).


T 95

6.10 Application of J M A K model incorporating the effective austenite

xn
grain size for HSLA-65 steel (d =19 um and cooling rate 60C/s).
y 96

6.11 Comparison between measured and predicted temperature for

50% transformation. 97

6.12 Changes in value of B vs. strain in HSLA-65 steel. 98

6.13 Comparison between the measured and predicted values of ferrite

grain size of DP 600 steel with applied deformation of 0.5. 99

6.14 Comparison between the measured and predicted values of ferrite

grain size of HSLA-65 steel. 100

6.15 Comparison between the measured and predicted values of ferrite

grain size for both steels. 101

xin
LIST OF TABLES

Table Page

2.1 Values of n and m for various transformation modes. 33

4.1 Chemical compositions (wt. %) of the steels investigated. 37

5.1 Schedule of the heat treatment and resulting austenite grain size. 50

5.2 Deformation-transformation test schedule. 71

6.1 Slopes and intercepts of lnb for DP 600 and HSLA-65 steels. 91

6.2 Parameters describing the polygonal ferrite grain size. 99

xiv
LIST OF S Y M B O L S

A C] Austenite-to-ferrite equilibrium temperature

a Symbol denoting ferrite

y Symbol denoting austenite

T Temperature

C 0 Concentration of carbon in bulk material

C a Equilibrium concentration of carbon in ferrite

C y Equilibrium concentration of carbon in austenite

co Parabolic rate constant for lengthening

y/ Parabolic rate constant for thickening

S Thickness of ferrite

t Time

D Diffusion coefficient of carbon in austenite

p Reduction

S gb Grain surface area before rolling

S b Grain surface area after reduction

s True strain

d a Ferrite grain size

d Austenite grain size

d eff Effective austenite grain size

xv
cp Cooling rate

X Fraction transformed

b Rate constant in J M A K equation

n Time exponent in J M A K equation

T Incubation time

M Number of preferred sites for corner nucleation

T N Nucleation temperature

T s Transformation start temperature

C* Limiting carbon concentration for nucleation

m Exponential constant in grain size modified J M A K equation

F Polygonal ferrite fraction

E Fitting parameter in ferrite grain model

B Function of initial microstructure in ferrite grain model

d (T)
m Measured dilation

d (T)
a Extrapolated dilation from the austenite region

d {T) Extrapolated dilation from the product region

S a Expansion coefficients of austenite

S p Expansion coefficients of product phases

a True flow stress

<r 0 Prestrain yield point

(7 r Reload yield point

o m Maximum stress immediately before unloading

xvi
X soft Fraction softened

Tnr Temperature of no recrystallization

T def Deformation temperature

xvii
ACKNOWLEDGMENT

I would like to thank Professor M . Militzer, who supervised me during the course of this

work, for his ideas, valuable discussions and permanent encouragement.

Many thanks to Fateh Fazeli and the mates from the research group for their advice and

discussions which helped me to finalize this work.

I would like to express my gratitude to NSERC, Stelco Ltd. and Dofasco Inc. for their

financial support and material supplies.

My family were big fans of mine during the years of this work, supporting and

encouraging me in anyway that they could. Thank you!

xviii
C H A P T E R 1- I N T R O D U C T I O N

Steel products are widely used for many applications, such as automotive, pipeline and

construction fabrication. Development of technology, increased requirement and

competition for improved mechanical properties has led to a fast invasion of advanced

high strength steels in industry. Low carbon content with small additions of alloying

elements and appropriate thermo-mechanical processing result in a successful

combination of increased weldability, strength, and toughness. To optimize the process

parameters and achieve maximum performance of the product, microstructural process

modelling plays a very important role reflecting the influence of the variable factors.

Sheet steel can be supplied as hot-rolled or as cold-rolled and annealed product. The

properties of hot-rolled material are a result of hot-strip rolling. A schematic diagram of a

hot-strip mill process is shown in Fig. 1.1.

R e h e a t furnace R o u g h i n g mill Finishing mill R u n o u t table Downcoiter

Fig. 1.1 Schematic diagram of a hot strip mill. [1]

The first process step in a typical hot strip mill involves reheating of a steel slab in a

furnace to approximately 1250C to provide easy deformation conditions. During

heating, growth of austenite grains occurs and the temperature and austenite grain size

identify the initial rolling condition. The roughing mill, which could be a reversing or a

1
tandem mill, is the next step which reduces the cross section of the material and the

microstructure evolution is characterized by work hardening, recovery and

recrystallization. Further reduction of the material at lower temperatures, in the interval

1100C-850C, occurs in the finishing mill where microstructure changes similar to

rough rolling are observed. Then the finish-rolled strip enters the run-out table where

accelerated cooling in the range 10C/s-150C/s is applied and decomposition of

austenite occurs. Finally, after the completion of transformation precipitation of carbides,

nitrides and/or carbo-nitrides may take place during coiling of the hot band.

Cooling on the run-out table plays an important role in the development of improved

mechanical properties by applying accelerated cooling or alternative complex cooling

strategies. Studying the transformation kinetics and final product microstructure gives

critical information for evaluating the material behaviour. The continuous cooling

transformation depends strongly on several factors, such as prior austenite grain size,

cooling rate, presence of retained strain and chemical composition. Typically, advanced

high strength steels are low carbon steels with alloying elements such as niobium,

vanadium, titanium, molybdenum and silicon which can be found in solution and/or in

the form of precipitates.

Laboratory studies of the microstructural evolution during cooling allow the

development of predictive models which describe the kinetics of austenite decomposition

which are important as a tool to improve product quality and reduce cost.

This thesis investigates two high strength steels: a Nb-microalloyed H S L A steel

(HSLA-65) and a Mo-bearing dual phase steel (DP 600). The kinetics of austenite

decomposition and microstructural evolution during continuous cooling is investigated

2
and applied to run-out table process conditions through modelling. The effect of initial

austenite grain size, cooling rate and retained strain are quantified. Submodels are applied

to the transformation start, the austenite-to-ferrite transformation kinetics, and to predict

ferrite grain size.

3
CHAPTER 2- LITERATURE REVIEW

2.1 Resulting Microstructures from Austenite Decomposition

Solid-solid phase transformation processes in metals and alloys can be divided into two

groups, originally distinguished from each other by different dependencies of reaction

velocity and amount of transformation on temperature and time [2], i.e.:

1) Diffusional transformation

2) Martensitic transformation

A CCT diagram of a steel with 0.06 wt. % C and 0.50 wt. % M n (austenization

temperature of 950 C) indicating the transformation products (F-ferrite, B-bainite, P-

pearlite, M-martensite) is illustrated in Fig.2.1. When a hypoeutectoid steel is allowed to

cool continuously with sufficiently slow cooling rate, the primary transformation product

may be polygonal ferrite which nucleates at austenite grain boundaries and grows toward

the grain center by rejecting carbon into the remaining austenite. Polygonal ferrite is

characterized by equiaxed grains with regular boundaries [3, 4].

Another type of ferrite that can be obtained is acicular ferrite. It has non-equiaxed

grains with a form of plates. According to Honeycombe and Bhadeshia [5], the plates of

acicular ferrite nucleate on small non-metallic inclusions in the austenite grain interior

and grow in many different directions from these nucleation sites. A possible criterion to

distinguish this microstructure from the polygonal ferrite is the aspect ratio. A transition

structure from polygonal to acicular ferrite has been recognized as quasipolygonal ferrite

with non-equiaxed grains having irregular, winding boundaries [4].

4
There is a tendency for the ferrite to grow from the austenite grain boundary as plates

at larger undercoolings [7]. So-called Widmanstaten ferrite side-plates form which

become finer with increasing undercooling and can grow directly from the austenite grain

boundaries (primary side-plates) or from crystals of another morphology but the same

phase (secondary side-plates) [8]. Two shape of the side-plates are observed: sideneedles

and sawteeth. The variety of ferrite morphologies was summarized by Dube et al. [9|.

When austenite with 0.8 wt. % C is cooled below the A temperature a eutectoid

transformation occurs. The resulting microstructure is pearlite which appears as lamellae

or sheets of cementite embedded in ferrite [7|. Pearlite nodules nucleate and grow with a

roughly constant radial velocity into the surrounding austenite; their size depends on the

magnitude of the undercooling.

5
Bainite is a non-lamellar two-phase product, a mixture of ferrite and carbide [8]. Upper

bainite consists of needles or laths of ferrite with cementite precipitates between the laths.

The ferrite laths nucleate and grow into austenite in a way similar to the growth of the

primary Widmanstaten side-plates [7]. At sufficiently low temperatures lower bainite

forms which is characterized by finer plates and carbide dispersion similar to tempered

martensite. The diffusion of carbon is slow in austenite and carbide precipitation may

occur in the ferrite plates.

Martensitic transformation occurs at very high undercoolings by the cooperative

movement of atoms less than one interatomic spacing. The martensitic microstructure

appears in forms of plates or laths embedded in the matrix [8].

The transformation processes of low carbon steels found in industrial conditions, i.e.

on the run-out table, are mainly attributed to diffusional transformation. After completion

of the polygonal ferrite formation, the remaining austenite is enriched by the carbon

rejected from the growing ferrite and decomposes into secondary transformation

products. Acicular ferrite has been observed to form in the remaining austenite at lower

transformation temperatures [3]. Pearlite is another secondary product which is observed

during continuous cooling transformation of low carbon steels [4]. Further, the bainitic or

martensitic secondary microstructures can be obtained i f lower transformation

temperatures are reached [3].

6
2.2 Allotriomorph Ferrite Reaction

The isothermal formation of proeutectoid ferrite from austenite occurs by a

diffusionally controlled reaction which consists of subsequent events:

Incubation -> Nucleation > Growth Impingement

Referring to the Fe-Fe^C phase diagram, Fig. 2.2, at temperatureT an alloy with
}

composition C 0 has equilibrium carbon solubility in ferrite and austenite denoted as

C and
a C,y respectively. To reach the equilibrium concentrations, the carbon

concentration in ferrite must decrease by rejecting the excess of carbon into the

remaining austenite.

Wt% Carbon

Fig.2.2 F e - F e C phase diagram showing equilibrium concentrations for proeuctoid ferrite formation
3

from austenite [8].

A diffusion field is established with local equilibrium at the ferrite:austenite interface.

The rate of advancement of the interface is controlled by the removal of carbon, i.e.

diffusion in austenite. Fig.2.3 shows this field at temperature 7j.

7
Co

^ -

a y

Distance

Fig. 2.3 Diffusion field of carbon during the growth of proeutectoid ferrite.

Broad reviews of the proeutectoid ferrite formation have been made by Aaronson [11]

and Reynolds et al. [12], summarizing the process in terms of morphology, kinetics,

nucleation and growth mechanisms.

2.2.1 Nucleation of Proeutectoid Ferrite

Many studies have been done on nucleation of proeutectoid ferrite [13-16]. In

general, two types of nucleation are possible during the phase transformation [13]:

homogeneous and heterogeneous. Proeutectoid ferrite nucleates heterogeneously in

recrystallized austenite; the most probable nucleation sites will be grain surface, edges

and corners, shown in Fig. 2.4.

8
Grain Boundary Grain Edge Grain Corner

Fig.2.4 Heterogeneous nucleation sites in fully recrystallized austenite. [13|

These sites are associated with lower activation energy for heterogeneous ferrite

nucleation which decreases from grain boundary and edges to grain corners. Increased

undercooling accelerates nucleation at less preferable sites on austenite grain boundaries.

Enomoto and Aaronson [14] proved experimentally that a spherical-cap based model

of the nucleation is practically not applicable due to large differences between the

calculated and measured values of nucleation rate. Further, they investigated the

nucleation kinetics of proeutectoid ferrite at austenite grain boundaries in low carbon

ferrous alloys and adopted the basis of classical heterogeneous nucleation theory to the

allotriomorph proeutectoid ferrite nucleation by using a pillbox shape model of nuclei

[15, 16]. The suggested model is associated with the maximum driving force for

nucleation and the minimum activation barrier. Pillbox-cum-spherical cap, which

represents partly coherent interfaces, is applied to the model, as well. It is associated with

larger undercooling when the nucleation shifts from the grain corners to the grain edges

and finally into the general grain area. Enomoto and Aaronson [14] extended their model

to the critical nucleation at grain edges assuming a critical nucleus shape of trigonal

prism.

9
Militzer et al. [17] applied the Enomoto and Aaronson model to conditions of

continuous cooling which is the basis of a model considering transformation start

temperature. Analysing the nucleation rates, they found that the nucleation temperature

does not depend on cooling rate and a nucleation model alone cannot account for the

observed transformation start temperatures.

2.2.2 Growth

Once ferrite has nucleated at boundary sites, growth of ferrite will proceed, being

driven by the difference in free energy between the austenite and ferrite phases. The

kinetics can be divided into thickening and lengthening (Fig.2.5):

Thickening

Fig.2.5 Thickening and lengthening of grain boundary allotriomorph [18].

Lengthening kinetics of proeutectoid ferrite provide results indicating that planar

interface-carbon-controlled and ledge mechanisms take place during the transformation.

Simonen et al. f 19] observed that lengthening occurs by the ledge mechanism. Their

model approach considers the radius of growing side-plates curvature. In a case of

10
diffusion controlled lengthening [20], a parabolic relation can be applied with a parabolic

rate constant, co.

Two possible mechanisms of thickening are by migration of planar incoherent

interfaces and by migration due to ledge movement in case of semicoherent plate-like

formations, respectively [7]. The rate for thickening [20, 21] can be expressed as a

parabolic function, similar to the lengthening, but with a different parabolic constant, if/.

Fig. 2.6 shows the lengthening and thickening in Fe-0.23 wt. % C at a transformation

temperature of 775 C. As it can be seen lengthening occurs much faster than thickening.

o I i I l I i I i I i I I
0 1 2 3 4 5 6
R e a c t i o n Time, sec." 2

Fig. 2.6 Typical plot of allotriomorph half-length and half-thickness as a function of the square root
of the reaction time. [20]

It was found [20, 22] that the average aspect ratio between half-thickness and half-

length in all alloyed and non-alloyed steels is 1/3. Since the lengthening of ferrite is much

11
faster than thickening, austenite grain boundaries are usually completely covered by

ferrite grains before appreciable growth by thickening has occurred [17]. In sufficiently

hardenable steels, particularly for higher undercoolings and larger austenite grains, the

observed microstructure may be allotriomorph ferrite as a thin layer on the grain

boundaries and bainite or martensite in the remaining volume of the grain. Further, the

ferrite film on austenite boundaries also constitutes the initial stages of transformation in

fully ferritic steels.

If the a - y interface of a grain boundary allotriomorph is considered to be incoherent,

a simple one-dimensional carbon diffusion model with a parabolic relationship can be

applied [23]:

(2.1)

where is the half thickness of ferrite, y/ is a parabolic rate constant of thickening

and t is time.

If a linear carbon gradient in the remaining austenite is assumed, the parabolic constant

y/ can be expressed as [8]:

(2.2)

where D is carbon diffusivity in austenite.

12
Three assumptions are met in this equation: (1) The interphase boundary is planar and

disordered; (2) The kinetics are volume diffusion controlled; (3) D is composition-

invariant.

More complex diffusion models which provide a correct solution for the diffusion

process in the remaining austenite would give a parabolic growth as well [21, 23].

2.2.3 Impingement

At the later stages of the transformation, the carbon diffusion fields from adjacent

ferrite grains, each being created by rejection of carbon into remaining austenite, overlap.

This process reduces the carbon gradient at the interface and thus slows down the carbon

diffusion. The phenomenon is called soft impingement; another type of impingement

which occurs when ferrite grains impinge on each other is known as hard impingement

[19]. The parabolic growth law does not hold anymore when either type of impingement

occurs.

Kamat [24] applied a finite-difference method to incorporate soft impingement to

carbon diffusion controlled ferrite growth. The model assumed planar and spherical

geometries, no carbon flux in the centre of the austenite grain, and local equilibrium at

the ferrite:austenite interface. The concentration profile of carbon in front of the growing

ferrite has been predicted. In the early stages of transformation, the planar model shows

very good agreement with the parabolic constants experimentally measured [25].

Discrepancy increases with increasing growth time, showing that soft impingement

occurred and growth rate decreased approaching zero. The spherical model showed

reasonable results for overall ferrite transformation.

13
2.3 Factors affecting austenite decomposition

2.3.1 Cooling Rate

Cooling rate is a factor that has a strong influence on the final microstructure in all

steels. By changing the cooling conditions during continuous cooling transformation of

low carbon, high strength steel, different transformation products can be obtained from

coarse polygonal ferrite to ferrite with significantly refined grains, or formation of

acicular ferrite and/or bainite

The cooling rate has a strong effect on the kinetics of austenite decomposition [10].

The transformation of austenite to low-temperature products is a thermally activated

process and thus the new product needs time for nucleation and growth. Increasing the

cooling rate, the available time at any given temperature for the transformation to initiate

is decreased, and the transformation start temperature decreases [10, 17, 26-28]. As a

result, the entire transformation is shifted to lower temperatures. On the other hand, rapid

decrease in temperature leads to activation of more nucleation sites due to the excess of

the free energy difference of the austenite and the forming ferrite, which are not available

at higher temperatures [7]. Thus, the result is a finer ferrite microstructure. However, at

very high cooling rates, because of reduced diffusivity, the transformation rate decreases

again which is reflected in the C-shape of CCT diagrams.

Speich and Scoonover [3] conducted a full examination of the effect of cooling rate on

the strength and the microstructure of HSLA-80 steel. At relatively high cooling rates,

they obtained low carbon martensite that is distinguishable by its high hardness and the

fine dislocation lath substructure. At lower cooling rates, the first transformation product

14
is acicular ferrite with no precipitates between the ferrite laths. A high dislocation density

is present in ferrite [4] but the dislocation cells are more irregular and more widely

spaced than in that of martensite. Growth is controlled by carbon diffusion away from the

edge and sides of growing plates, enriching the remaining austenite with more carbon. At

still lower cooling rates, polygonal ferrite is the prime product of transformation. The

temperature, at which polygonal ferrite forms, increases as the cooling rate decreases. As

a result, the polygonal ferrite grain size gradually increases with decreasing cooling rate.

The carbon-enriched austenite decomposes into secondary transformation products

[29]. Initially, the remaining austenite transforms to acicular ferrite or pearlite. As the

carbon content of the austenite increases at slower cooling rates because of the formation

of a greater amount of polygonal ferrite, transformation to acicular ferrite is suppressed.

The carbon-enriched austenite is stabilized until lower temperatures and transforms into

bainite or martensite. Presence of remaining austenite at room temperature is possible and

is referred to as a martensite-austenite constituent microstructure.

From an industrial viewpoint, it is possible to create appropriate conditions for

precipitation of carbides, nitrides or/and carbonitrides in H S L A steels during coiling

which lead to an increase in strength. Interrupted accelerated cooling was introduced [30,

31]. Coiling temperature can be manipulated by halt of the accelerated cooling which

gives the opportunity to control the precipitation. Maximum strengthening effect of

precipitates may be obtained.

The as-rolled dual phase steel concept is based on special continuous cooling

transformation characteristics which permit the steel to be processed on a conventional

hot strip mill to produce desired polygonal ferrite-martensite microstructure in the as-

15
rolled coiled sheet [32]. These characteristics related to the C C T diagram are: (1) an

elongated ferrite C-curve which promote formation of polygonal ferrite over a reasonable

wide range of cooling rates on the run-out table; (2) suppressed pearlite formation; (3)

high pearlite finish temperature to avoid perlite formation during coiling at temperatures

up to 620 C; (4) suppressed bainite formation to provide a temperature range at which

martensite forms before the steel to be coiled. A schematic continuous cooling behaviour

of a dual phase steel is shown in Fig. 2.7:

R o u g h i n g mill Finishing mill Runout table Coiler

f 0088888 llllll WttttHH

CX.;
E
F Ferrite
P Perlite
B Bainite
M Martensite M
Time

Fig. 2.7 Schematic representation of continuous cooling transformation in dual phase steel resulting
in polygonal ferrite-martensite microstructure in as-rolled product [33].

By controlling the cooling rate during the transformation, a final microstructure can be

obtained which determines the desired mechanical properties. In this way accelerated

cooling plays a critical role in the controlled thermo-mechanical treatment of steels.

16
2.3.2 Prior Austenite Microstructure

The microstructure of the material just before entering the run-out table is very

important for the transformation process. There are two factors to be considered:

austenite grain size and degree of pancaking (i.e. amount of retained strain).

2.3.2.1 Austenite Grain Size

Initial austenite grain size has an effect on transformation of polygonal ferrite in terms

of providing nucleation sites for heterogeneous nucleation on the grain surface.

Decreasing the austenite grain size leads to larger grain-boundary area per unit volume

and thus the probability of nucleation is increased. As a result, the transformation starts at

higher temperatures [10, 26, 34, 35]. The final microstructure will develop finer ferrite

grains because of: (1) more nucleation sites produce more ferrite grains; (2) intragranular

space provided by smaller austenite grains limits the growth of the new phase because of

earlier impingement of new grains, leading to significant grain refinement. It is observed

that with a decrease of the austenite grain size, the transformation rate increases due to

the increase in the ratio of the nucleation rate to the growth rate [36]. Large austenite

grains depress polygonal ferrite formation and increase the percentage of non-polygonal

microstructure [10].

2.3.2.2 Effect of Retained Strain

Deformation of steel at temperatures below the recrystallization stop temperature, T , nr

leads to formation of elongated (pancaked) austenite grains. As a result, more nucleation

sites for a new phase are introduced and transformation may occur at higher

17
temperatures. The larger number of nucleation sites are provided because of the change in

grain surface area and activation of intragranular nucleation sites, such as deformation

bands, sub-grain and twin boundaries [37]. A number of studies have been performed

[38-42] that show the strong effect of deformation on the transformation and provide

clear evidence that retained strain in austenite contributes to ferrite grain refinement.

There are three factors that contribute to the increase of ferrite nucleation rate in

pancaked austenite:

1) The increase in surface area.

The surface area per unit volume increases by grain elongation [43, 44]. Umemoto et

al. [45] assumed that the initial austenite grain has, for simplicity, a spherical shape and a

normalized radius of one. By applying deformation with reduction p, its shape becomes

ellipsoidal as shown in Fig.2.8:

Fig. 2.8 The shape change of an austenite grain by reduction p. [45]

The surface area of a grain before rolling is (after the assumption that r = 1):

S = An(rf = 4/zr(l) =An


2 (2.3)

and after the reduction p it changes to:

18
x {\-pf
p )sin
2
2 2
Ax^\-(2p- 6do\x + \tix (2.4)
J-l/l-p
\-x (\-p)<
2

The ratio of the surface area after and before reduction is q = *'Vlo . Fig. 2.9

/
S
8-b.

presents q as a function of the reduction. It can be seen that q increases about two times

by applying a reduction of about 0.7:


Strain c

Rolling reduction p

Fig. 2.9 Ratio of grain boundary surface area as a function of rolling reduction [45].

2) Increase of the nucleation rate per austenite grain surface area unit due to the

increase in the irregularity of grain boundaries.

When a specimen is deformed in the austenite condition, the smooth and flat grain

boundaries become wavy and irregular as a result of non-uniform slip mechanism

occurring inside the grain. As a consequence many small regions with high grain

19
boundary energy will be produced. Such regions may act as preferential sites for ferrite

nucleation [45].

3) Formation of additional intragranular nucleation sites.

When deformation is applied below the recrystallization temperature, deformation

bands are produced and they act as additional nucleation sites. Therefore, the density of

nucleation sites per unit volume is increased. Umemoto et al. [45] found that the average

fraction nucleated within the grains was 25% of the total transformed fraction. Also, they

concluded that the transformation rate is increased due to the increase of nucleation rate

but not in growth rate. Kozasu et al. [46] introduced an effective austenite interfacial area

that represents the overall interfacial area of the austenite grain boundaries and

deformation bands available for nucleation.

2.3.2.3 Effective Grain Size

A combined effect of austenite grain size and retained strain on transformation can be

expressed by introducing an effective grain size. Lacroix et al. [47] suggested a very

simple model to this approach. They assumed that the effective grain size could be

expressed as a deformed austenite grain with ellipsoid shape. Further, it is postulated that

the ferrite instantaneously covers all grain boundaries as a thin layer. The ferrite layer

growth occurs in the direction of the small axis. Thus, the ferrite growth rate is

determined by growth along to the small axis, d e x p ( - ) . Then, representing the length
y

of the small axis, an effective grain size can be introduced, such that:

d
cff = d
r exp(-e) = d (\-p)
y (2.5)

20
2.3.3 Chemical Composition

Interstitial and substitutional alloying elements added to pure iron can dramatically

change the range of stability of the phases which leads to a large variety of

microstructures and mechanical properties. For dual phase steels, carbon and manganese

are present as alloying elements with the addition of molybdenum, silicon, and

chromium. Nb, Ti, V are important microalloying elements in H S L A steels with a base

chemistry of carbon and manganese.

Three modes of transformation in presence of substitutional alloying elements were

suggested by Bradley and Aaronson [22]: 1) At small undercoolings an / : a interface

may migrate with sufficient partitioning of the alloying element between austenite and

ferrite and thus, the conditions are very close to equilibrium. In this case growth is

controlled by the diffusivity of the alloying element. 2) At higher supercooling the

interface a -y may still migrate with equilibrium interface compositions but without

partition of substitutional elements between the two phases. A "spike" is formed and

growth kinetics are controlled by the volume diffusion of carbon in austenite. 3) No

partition of alloying elements occurs and transformation is controlled exclusively by

carbon diffusion. The second and the third modes occur during typical cooling on the

run-out table.

Most substitutional elements retard the rate of diffusion-controlled transformation in

steels. The alloying effect on growth may be attributed to [48]:

1. Thermodynamic effect of alloying elements

2. Ternary diffusion interaction

21
3. "Solute drag" effect at the a-y interface due to the slowly diffusing alloying

element.

2.3.3.1 Carbon

Low carbon steels contain up to 0.25 wt. % C by classification. Increased carbon level

lowers the A temperature significantly and reduces the driving force for austenite

decomposition at any temperature. [49] The addition of carbon shifts the TTT diagram to

longer times, i.e. carbon is an austenite stabilizer, and facilitates the formation of non-

equilibrium microstructures such as bainite and martensite.

2.3.3.2 Manganese

Manganese is one of the most important alloying elements in low carbon steels. This

element is another austenite stabilizer which decreases the A temperature. Furthermore,

since the proeutectoid reaction involves rejection of manganese from ferrite, at relatively

small undercoolings, a positive manganese spike at the interface was observed [48] and

the growth of ferrite is retarded.

Manganese affects the steel microstructure by: (1) grain refinement due to depression

of the y -> a transformation temperature; (2) modification of the transformed

microstructure [50]. For instance, the dominant resulting microstructure in H S L A steels

on the run-out table is polygonal ferrite with a few percent of pearlite. A n increase in the

manganese-carbon ratio under reduced carbon content gives rise to acicular ferrite or an

ultra-low carbon bainitic structure. The final microstructure can be related to the

transformation temperature that is strongly dependent on the steel chemistry.

22
2.3.3.3 Molybdenum

Kinsman and Aaronson [25] conducted extensive investigations on the effect of

molybdenum on isothermal transformation kinetics. They used a thermoionic emission

microscope, which allowed them to investigate the nucleation and growth in plain, M n ,

and Mo-bearing low carbon steels at elevated temperatures to describe the effect of

alloying elements. Also, they were capable of measuring the effect of alloying elements

on nucleation and growth rate separately.

Molybdenum-bearing steel shows very interesting behaviour, where growth kinetics

form "a nose" at 800 C and "a deep bay" at about 600 C (Fig. 2.10):

1700

1500 I
to
1300 2L
E
CD

1100 *~

900
1 10 10 2
10 3
10 4

Reaction Time, seconds

Fig. 2.10 Effect of M o on kinetics of austenite to ferrite transformation. [25]

The possible explanation of this behaviour is related to a "solute drag effect" similar to

that observed for grain growth and recrystallization. Molybdenum shows tendency to

segregate to the disordered austenite:ferrite boundaries, where it cannot be dragged along

with the boundaries by volume diffusion [51]. Mo atoms can diffuse for short distance

along the austenite:ferrite interface before completing their transfer to ferrite, or simply

23
serve as "immobile pinning points" around which the boundary must bend before it can

break away. On the other side, molybdenum is a strong carbide-former and tends to react

with carbon which concentrations are high at a :y interface. The formation o f M o C 2

rows and fibres were observed [52] at moving austenite:ferrite boundaries. The presence

of the deep bay at 600 C corresponds to the state when the solute drag and/or carbide

effects are at or near its maximum effectiveness.

An extended study on the effect of alloying elements on continuous cooling

transformation in an as rolled dual phase steel was conducted by Coldren and Eldis [53].

Their examinations show that molybdenum has a pronounced effect on retardation of

pearlite formation. Mo-bearing steels with composition 0.05% C, 1.21%Si, 0.50% Cr,

and different levels of M o were investigated. As the Mo content increases from 0% to

0.50%o, the pearlite nose in C C T diagram shifts markedly from 10 to 10000 seconds.

They introduced a ratio of maximum to minimum acceptable run-out table cooling rates

(C7? max ICR )


min for obtaining good dual phase structure, i.e. about 90% of polygonal

ferrite and 10 % martensite, and no pearlite. Changing the amount of Mo, they found the

relationship shown in Fig. 2.11. As it can be seen, the working range increases from 0.4

to approximately 30 of C7? max /C7? by addition of 0.50% wt Mo. They concluded that
min

the optimum content of M o is about 0.38%o because a ratio (C7? max / CR )


min of 10 is

sufficient for typical mill conditions.

24
-J I I
1.0 1.5 2.0 %'Si
AUOr CONTENT, WT. %

Fig. 2.11 Effect of M o on ratio of maximum to minimum cooling rates producing ferrite-martensite
microstructure on run-out table [53].

2.3.3.4 Niobium

Many authors have investigated the effect of niobium on the austenite decomposition

[37, 54-58]. Fig. 2.12 shows the effect of Nb concentration on transformation start

temperature in steels with a base composition of 0.08 wt.% C, 1.39 wt.% Mn, and 0.36

wt.% Si, an austenization temperature of 950 C and a cooling rate of 1 C/s.

25
900

850 -

o
a> 800
3
as

E
o
600 I i i < 1 1 i
2 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

Nb Content (wt %)

Fig. 2.12 Effect of Nb content on transformation start temperature. [53]

The observed effect is the delay of the onset of the proeutectoid ferrite transformation

with addition of Nb and maximum retardation occurs at some intermediate Nb content of

approximately 0.04 wt.% [54, 55]. The observed minimum in the transformation

temperature can be attributed to the effects of competing mechanisms. At low Nb

additions, below 0.04 wt.%, part of the niobium may form fine matrix precipitates during

cooling but a large portion remains in solution. Thus, niobium can segregate readily to

austenite grain boundaries [56, 57] where retardation of ferrite nucleation occurs by

lowering of the local driving force for nucleation and/or preventing carbon to diffuse

from ferrite nuclei (Nb has strong affinity to carbon). However, i f the Nb content is

increased beyond 0.04% wt., a rapid formation and coarsening of the Nb(C,N)

precipitates occur at the austenite grain boundaries. Ohmori [59] found that coarse grain

boundary Nb precipitates could be potent nucleation sites for ferrite formation. These

particles have a large mismatch with respect to austenite and provide high energy

interfaces suitable for ferrite nucleation. The formation of large Nb(C,N) particles also

26
locally reduces Nb and carbon contents in solution. As a result, the strong solute drag

effect of Nb is diminished and ferrite formation is promoted.

Summarizing the effect of alloying elements, it can be said that they in general depress

the transformation from austenite to low-temperature products that lead to two important

features: refinement of polygonal ferrite grains and promotion of non-polygonal ferrite

structures.

2.4 Modeling of Austenite-to-Ferrite Transformation

2.4.1 Kinetics of Isothermal Transformation- J M A K Approach

The isothermal kinetics of diffusional phase transformations was initially described by

Kolmogorov [60], Johnson and Mehl [61], and Avrami [62-64]. A general form that

describes a phase transformation kinetics can be given by:

X = l-exp(-bt") (2.6)

known as Johnson-Mehl-Avrami-Kolmogorov (JMAK) or simply the Avrami equation.

The value of b is related to the magnitude of the nucleation and growth rate and it is

temperature dependent, while n is a factor that depends on the nucleation conditions and

the shape of the growing particle. Fig. 2.13 shows an example for sigmoidal

transformation kinetics which includes nucleation and growth of a new phase:

27
o
E
o
tn

c
T>
ut

Time (s)

Fig. 2.13 Sigmoidal transformation kinetics [10].

2.4.2 Application of Isothermal Transformation Kinetics to Non-Isothermal

Conditions

Observation of isothermal austenite decomposition simplifies the interpretations since

the transformation temperature, as a fundamental parameter, is a constant. However,

continuous cooling transformations are widely used in industrial conditions where the

temperature decreases with time. Thus, there is a need to translate data from isothermal

kinetics, which is much more extensively studied, to non-isothermal treatment paths.

Continuous cooling transformation can be expressed as a sum of short-time isothermal

steps as shown in Fig. 2.14:

28
Ah A Ij A t ,

Fig. 2.14 Continuous cooling transformation shown as the sum of short-time isothermal steps [65|.

The approach of applying isothermal transformation kinetics to non-isothermal

conditions was initially proposed by Scheil [66] who used a method to describe

nucleation during cooling based on the incubation period associated with isothermal

transformation. A s s u m i n g t h a t t h e t i m e n e e d e d f o r i n c u b a t i o n at t e m p e r a t u r e T is z(T)

a n d t h e t i m e s t e p i s e q u a l t o dt, at e a c h t e m p e r a t u r e a f r a c t i o n o f i n c u b a t i o n t i m e e q u a l to

dt I r(T) w i l l be c o n s u m e d . T h u s , for the c o n s u m p t i o n o f the i n c u b a t i o n time o v e r a range

of temperatures d u r i n g c o n t i n u o u s c o o l i n g , t h e i n c r e m e n t a l f r a c t i o n s c o n s u m e d at each

t e m p e r a t u r e c o u l d b e s u m m e d . W h e n t h i s s u m e q u a l s o n e , i.e.:

(2.7)

incubation will be complete.

Cahn [67] generalized the additivity principle for isokinetic reactions (i.e. the

n u c l e a t i o n rate is p r o p o r t i o n a l to the g r o w t h rate) b y p o s t u l a t i n g that a r e a c t i o n is a d d i t i v e

w h e n e v e r the rate is a f u n c t i o n o n l y o f f r a c t i o n t r a n s f o r m e d a n d t e m p e r a t u r e :

29
(2.8)

In contrast, Christian [2] suggested that additivity is only fulfilled when the

transformation rate can be expressed with two independent functions, of the temperature

and fraction transformed, respectively, in a form of:

^ = (2.9)

where %(T) is a function of temperature and (X) a is a function of fraction transformed.

More recently Lusk and Jou [68] confirmed that the solution given by Christian is

required for a reaction to be additive.

The Avrami equation can be adapted to the principle of additivity by using Eq. 2.6 and

taking the derivative with respect to time:

=fc*r
dt f "-'
-l
i
1 A
(l-X)
l n1( l - X )
W

Eq. 2.10 satisfies the criterion for additivity provided that b=f(T) and =const.

30
2.4.3 Overall Transformation Model Applicable to Run-Out Table Simulated

Conditions

2.4.3.1 Transformation Start Temperature

The transformation start model is associated with nucleation and early growth of

ferrite at the austenite grain boundaries.

A simple spherical geometry of ferrite, associated with steady-state growth can be

given by:

dRf C-C 1 0

L = D- - (2.11)
dt C -C R y a f

where R f is the growing ferrite particle, D is the diffusion coefficient of carbon in

austenite, C is the average carbon bulk content, C and C are the carbon equilibrium
0 a y

concentrations in ferrite and austenite, respectively. Integration for continuous cooling,

T = T - (pt, yields:
N

1
1 \ ? (2.12)
C C
D d T
R
f

Grain coiners are initially preferred for nucleation at small undercooling. The number,

M , of preferred sites for corner nucleation was estimated [17] to equal approximately

two per grain. During continuous cooling, growth of these early formed nuclei determines

how much austenite boundary area is available for additional nucleation at higher

31
undercoolings. Nucleation temperature, T , N at grain corners can be estimated from

transformation start for slow controlled cooling when sufficient time is available to obtain

nucleation and early growth at virtually the same temperature. As analyzed by Militzer et

al. [17], T can be assumed to be independent of cooling rate.


N

During continuous cooling, as the temperature decreases, a shift of the nucleation sites

occurs from preferred grain corners to edges and grain boundary area. The transformation

start is associated with nucleation site saturation, i.e. it is determined as the temperature

at which the entire austenite grain boundary is covered by a layer of ferrite. Nucleation

cannot take place at those boundary areas which are already covered by ferrite, nor in the

vicinity of growing ferrite where the increase in carbon concentration in the austenite

results in a reduction of the driving force for nucleation. Thus, the nucleation rate

becomes a function of the distance from the growing ferrite [69]. A limiting carbon

concentration is introduced, i.e. nucleation can only occur in areas where the carbon

concentration remains below a specific concentration, C*, which defines the nucleation

limit radius, r* around the growing ferrite:

When M (r*)
p
2
= d,
2
further nucleation becomes impossible and for a constant cooling

rate, cp, T can be estimated from:


s

(2.14)
r

32
Equation 2.11 shows the combined effect of cooling rate and austenite grain size on the

transformation start.

2.4.3.2 J M A K Equation and Additivity Rule Applications

The transformation kinetics depends strongly on the initial grain size because the

nucleation occurs usually heterogeneously at grain boundary, edges or corners. Umemoto

et al. [36] considered this feature and modified the Avrami equation (Eq. 2.6) in the form

of:

X = 1 - exp (2.15)

where b and n have the same meaning as in Eq. 2.6, and m is a fitting parameter

dependent on the growth geometry. The values of m and n for different transformation

products are summarized in Table 2.1.

Table2.1 Values of n and m for various transformation modes [70].

Transformation n m Nucleation

Pearlite 4 2 Edge Nucleation


Nucleation and Growth
Ferrite 1 1 . Surface Nucleation near Site Saturation

Bainite 4 0.6 Grain Boundary, Inside Grain, Nucleation and


Growth

Typical values of n for ferrite formation in low carbon steels are in the range of 0.8-1.2

which can be found in literature [10, 26, 35, 71, 72]. A wide variety of functions were

33
used to describe the kinetics parameter b. A linear function of Inb vs. temperature was

used [10] where two parameters were sufficient to describe the transformation rate.

Nakata and Militzer [35] used two linear functions of Inb to characterize separately the

earlier and later stages of transformation kinetics in an H S L A steel. A more complex

function in a polynomial form was applied by Hawbolt et al. [72].

Many studies have been conducted adapting the additivity principle for the y > a

transformation. Among them are those of Hollomon et al. [73], Manning et al. [74], and

more recently further extended studies of the problem [17, 71, 72]. Leblond and Devaux

[75] extended the approach in terms of formation of two and more phases as products of

austenite decomposition.

2.4.3.3 Polygonal Ferrite Grain Size

The ferrite grain size depends strongly on the start of transformation when site

saturation occurs and no additional nucleation can take place below T . No significant
s

ferrite grain growth has been observed after completion of transformation [10, 76].

Subsequently, the number of ferrite grains, and thus the resulting grain size, is related to

the number of additional nucleation sites at grain edges and grain boundaries created

during continuous cooling in the initial transformation stages.

The model used for d a can be expressed as a function of transformation start

temperature. The equation is based on the approach proposed by Suehiro [77]:

X
Fexp B (2.16)
5 J

34
where d a is ferrite equivalent area diameter (EQAD) in jum , F is the final polygonal

ferrite fraction, measured by optical metallography, T is the transformation start


s

temperature in K , E is a fitting parameter, and B is a function of the initial austenite

microstructure. In strain-free cases, B can be expressed as [26]:

B = Sd q
r (2.1/

where S and q are fitting parameters.

35
C H A P T E R 3-SCOPE A N D O B J E C T I V E S

The knowledge of relationships between the variable conditions on the run-out table

and the relevant microstructure is important for determination of the final mechanical

properties of the material. The goal of the present research is to apply models for the

austenite decomposition kinetics and the resulting microstructure, applying similar to

run-out table conditions, to two high strength steels; a Nb microalloyed H S L A steel and a

Mo-bearing DP steel.

To achieve this objective, the study is divided into two parts:

1. Experimental:

a) Quantifying the effect of initial austenite grain size, cooling rate, and retained

strain on the austenite decomposition kinetics.

b) Determination of the resulting microstructure, including the quantification of

polygonal ferrite fraction and ferrite grain size.

c) Hardness measurements.

2. Application of models previously developed to describe the transformation kinetics

in both steels:

a) Transformation start temperature

b) J M A K model adopting additivity to describe polygonal ferrite growth

c) Polygonal ferrite grain size

36
C H A P T E R 4- E X P E R I M E N T A L

4.1 Materials

The materials investigated are a high strength low alloy-Nb steel (HSLA-65) and dual

phase Mo-bearing steel (DP 600). The chemical composition (wt. %) of each steel is

given in Table 4.1. Dofasco Inc. and Stelco Ltd. supplied these materials as transfer bars

(product at the end of the roughing mill), which were used to fabricate the samples for

laboratory testing.

Table 4.1 Chemical compositions (wt. %) of the steels investigated.

Grade C Mn P S Si Mo Nb Ti Al V N

HSLA- 0.062 1.240 0.007 0.004 0.051 0.008 0.063 0.002 0.040 0.001 0.007
Nb
DP- 0.060 1.860 0.015 0.004 0.077 0.155 - 0.011 0.043 - 0.007
Mo

T H E R M O C A L C software was applied to determine the A for each steel which are

822 C and 840 C for DP 600 and HSLA-65, respectively.

Austenite grain growth samples were parallelograms, with a height of 3 mm, a width of

5 mm and a length of 6 mm. Continuous cooling test samples have been prepared with a

tubular shape having a wall thickness of 1 mm, outer diameter of 8 mm, and length of 20

mm. The samples used for double hit compression tests have been machined as solid

cylinders with diameter of 10 mm and length of 15 mm.

37
4.2 Experimental Equipment

The Gleeble 3500 thermo-mechanical simulator has been used to perform all tests.

The temperature has been controlled by using Chromel-Alumel thermocouple (type K ,

d=0.25 mm) that has been spot-welded to the outer surface and at a middle point of the

length of each specimen. Electrical resistance has been applied for heating, soaking, and

cooling with feedback control. Natural cooling (heating has been shut off) or gas (He)

quenching has been applied for higher cooling rates. To avoid oxidation and to minimize

decarburization the chamber has been evacuated to a pressure under 3 mTorr and then

back filled with argon gas.

Low and high-force jaws were utilized to perform tests for phase transformation and/or

deformation.

During the experiments the built-in software QuickSim has performed the feedback

control of the parameters. The temperature, time, stroke, force, and dilatometer responses

were recorded in each transformation and/or deformation test for analysis purposes.

4.3 Experimental Procedure

4.3.1 Austenite Grain Size

Fig. 4.1 shows the schematic thermal path used to obtain the three initial austenite grain

sizes for each steel where T is temperature and t is holding time.

38
Hold-Ti,ti

Time

Fig. 4.1 Schematic thermal path for obtaining the initial austenite grain sizes.

First, austenizing test were performed to establish suitable reheating conditions for

subsequent transformation tests to quantify the effect of austenite grain size on

transformation. It is necessary to obtain at least three initial austenite grain sizes for each

steel. Further, the initial microstructure must be uniform since an abnormal austenite

grain microstructure would initiate ferrite formation at different temperatures and the

kinetics will be different for the different austenite grain sizes. The variables that can be

used to obtain a range of grain sizes are heating rate, soaking time and temperature. The

samples were cooled to 900 C with a cooling rate of 10C/s and were held at this

temperature for 30 seconds to facilitate the subsequent water or fast gas quenching to

room temperature.

39
4.3.2 Continuous Cooling Transformation

The volumetric changes during the continuous cooling transformation were measured

by a crosswise dilatometer located at the temperature measurement cross-section plane,

Fig. 4.2.

Specimen
Thermocouple
Spring

Helium gas_ Helium


used to quench gas out

C r o s s w i s e strain
measurement
C u electrical
Anvil
contacts

Fig. 4.2 Schematic diagram of C C T test. [10]

A schematic schedule of CCT tests is given in Fig. 4.3:

Hold-Ti,ti

Time
Fig. 4.3 Schematic schedule of C C T tests.

40
The samples were austenized to obtain desired austenite grain size, as determined in

the tests before. Investigated cooling rates ranged from 1 C/s to 250 C/s. Lower cooling

rates (<15C/s) have been obtained by controlling the resistance heating. The cooling

rate of about 15C/s was obtained by turning the power off, i.e. air cooling. Higher

cooling rates (>15C/s) have been obtained by applying helium gas introduced through

the inner sample surface. Varying the He flow rates leads to a range of cooling rates up to

250C/s.

The cooling rates have been calculated in the temperature range of A 20 C.

Quantification of the phase transformation is based on the difference of the austenite

and product molar volume. Austenite is FCC, i.e. a closed-packed structure and ferrite is

BCC with a lower density. The sample volume increases during the phase transformation.

Thus, this phenomenon can be recorded and presented by dilation response vs.

temperature. Applying the Lever rule [55] we obtain:

d (T)-d (T)
X(T) = ^ ^ - -^- a
(4.1)
d(T)-d (T) a

where X(T) is the fraction transformed,

d (T) is measured dilation,


m

d (T) = I + S T
a a a is the extrapolated dilation from the austenite region

d (T) = / + ST
p is the extrapolated dilation from the product region.

41
with thermal expansion coefficients S a and S for austenite and product phases

respectively.

A schematic illustration of determining the fraction transformed during continuous

cooling is given in Fig. 4.4. From these curves both the kinetics of the austenite-to-ferrite

transformation and subsequent transformations can be obtained.

Fig. 4.4 Example of calculation of fraction transformed during continuous cooling transformation.

4.3.3 Double Hit Test

Double hit tests are required to establish retained strain conditions in the austenite

region where recrystallization is sufficiently slow and the applied strains can be retained,

i.e. deformation under no-recrystallization conditions.

A C-strain device, located at the temperature measurement cross-section plane, has

been used to control the strain and strain rate during the deformation experiments. Double

hit tests have been performed using interruption at a selected strain (0.2 and 0.5) and for

42
different times, before again deformed at the same temperature and strain rate. Fig. 4.5

illustrates an example of double hit test. The yield points at the first hit and reloading, and

the stress immediately before unloading can be used to determine the degree of static

softening [78, 79].

Fig. 4.5 Example of a double hit test for H S L A steel with dy=49 um Tder=900 C and load interruption
at strain of 0.2.

The calculated fraction softened can be expressed as:

cr - cr
X
soft - (4.2)

where a m is the maximum flow stress immediately before unloading. <y and <r are the
r 0

yield points for reloading and prestraining, respectively, determined by the offset method,

in which yield point corresponds to a plastic strain of 0.2%.

43
4.3.4 Deformation -Transformation Experiments

The effect of deformation on transformation kinetics has been investigated by applying

two series of initial austenite grain size and two strains - 0.2 and 0.5 at constant

deformation temperature and strain rates for each steel. The samples needed for

deformation-transformation tests have a complex geometry as shown in Fig. 4.6. The

thermocouple was welded and the C-strain device was located at the mid plane of the

small diameter (6 mm). The applied cooling rates were in the range from 1 C/s to about

60C/s (natural cooling- power off). Gas cooling was not applied due to the sample

geometry; a solid body introduces a temperature gradient at the higher cooling rates

larger than 60C/s. The cooling rates have been calculated in the temperature range of

A e + 2 0 C , as for the regular CCT tests.

fc
M10

E
CO
I ....
1 I
10mm

1 300 m m

F i g . 4.6 Samples used f o r d e f o r m a t i o n a n d subsequent t r a n s f o r m a t i o n .

44
The schematic schedule employed for the deformation-transformation tests is given in

Fig. 4.7:

Hold-Ti, ti

Time
Fig. 4.7 Schematic schedule of deformation-continuous cooling transformation tests.

4.4 Microstructural Analysis and Techniques

Microstructural examination has been performed following every thermal or thermo-

mechanical test.

The samples for austenite grain size determination, continuous cooling transformation,

and deformation-continuous cooling transformation were cut off at the mid-plane where

the thermocouple was welded. A n aluminium oxide cut-off disk and water-cooling were

employed. One half of every specimen was cold moulded in epoxy resin in a position

accessible for processing 'the measured cross section. Grinding and polishing was

performed using a Buehler (Phoenix 400) grinding and polishing machine. By means of

three subsequent grades of silicon carbide sandpaper disks, 180, 600, and 1200 grid, the

45
grinding was performed. Cloth and diamond suspension of 6 urn and 1 um were used for

polishing

The prior austenite grain boundaries have been revealed in the DP 600 and H S L A 65

steels by applying an etching technique based on picric acid (2g saturated picric acid, 1

mL HC1, 1 mg dodecylbenzinesulfonic acid (wetting agent), 100 mL water) [80] at an

elevated temperature up to 60 C. The samples were immersed in the solution for 2 to 10

min until a good contrast was obtained between the matrix and prior austenite grain

boundaries. A n additional heat treatment of the HSLA-65 steel was introduced by

tempering it at 500 C for 24 hours to produce impurity segregation to grain boundaries

that leads to improvement of the subsequent contrast during etching. The austenite grain

size was determined by using Jeffries' method described in A S T M E l 12-88 standard

[81]. In this standard every whole grain is counted once and each partial one, cut by the

observed field edge, as a half grain. At least 50 grains have to be present in one observed

field. In addition, for every specimen at least 300 grains were counted. The number of

austenite grains per area and subsequent calculation of mean austenite grain area and the

mean equivalent area diameter, E Q A D , were performed. To facilitate the measurements,

Image Tool software was employed.

The polygonal ferrite fraction and grain size have been determined using a similar

technique to that of austenite grain size. After the grinding and polishing, described

above, the samples have been immersed in 2% nital solution until the best contrast has

been obtained between the ferrite matrix and grain boundaries. Image Tool software has

been used to determine the area of the second phase grains. Subtracting the second phase

area from the whole image area, the polygonal ferrite fraction has been obtained.

46
Maximum aspect ratio of 5 was used to separate polygonal from non-polygonal fraction

in the mixed structures, i.e. a ferrite grain with an aspect ratio of less than 5 was

considered to be polygonal. In case where polygonal ferrite is limited to ferrite plates at

grain boundaries, this method may lead to underestimation of the fraction of polygonal

grains. Polygonal ferrite grain size has been calculated similarly to that of austenite grain

size. Using the mean polygonal ferrite grain area, excluding secondary product grain

area, Jeffries' method has been applied to determine the E Q A D .

4.5 Hardness measurements

The hardness measurements have been performed with a microhardness tester

(Micromet 3 of Buehler Ltd.). Hardness measurement tests were conducted on each

transformation and deformation-transformation sample in the mid-plane where the

microstructural analysis was performed. The fields of measurements were randomly

chosen avoiding the area in the vicinity of the sample surface (due to oxidation and

decarborization). Twelve measurements were performed on each sample applying a load

of 200 g on the penetrating prism.

47
C H A P T E R 5- R E S U L T S

5.1 Initial Austenite Grain Size Conditions

A number of heat treatment cycles were performed for HSLA-65 and DP 600 steels to

obtain selected uniform microstructure with an appropriate austenite grain size as starting

conditions for subsequent transformation studies.

The austenite grain growth behaviour of the two steels at elevated temperature is

shown in Fig. 5.1:

300

0 H T , , ! 1 r - ^ 1
900 950 1000 1050 1100 1150 1200 1250

T e m p e r a t u r e [C]

Fig. 5.1 Austenite grain growth of DP 600 steel (holding time of 120 seconds at each temperature) and
HSLA-65 steel (holding time of 60 seconds at each temperature).

The HSLA-65 steel has shown uniform grain growth up to 1 150 C and a holding time
0

of 60 seconds. This behaviour can be related to the presence of Nb(C, N) precipitates at

temperatures below 1150C and is associated with normal, strongly pinned grain growth.

48
In the interval 1150 C-1200 C Nb precipitates rapidly dissolve, the pinning effect on

austenite grain boundaries is not so active and abnormal growth presumably occurs [10,

80, 82]. The microstructure at 1200C was uniform but accelerated grain growth occurs,

Fig. 5.2. Above this temperature very coarse uniform austenite microstructure was

observed as a result of lightly pinned growth.

Fig. 5.2 Austenite grain size of HSLA-65 steel with reheating temperature 1200 C and holding time
of 60 seconds.

The DP 600 steel has been investigated for searching of three appropriate initial

austenite grain microstructures as well. Austenite grain growth was uniform for

temperatures below 1100C and holding times of up to 780 seconds. Above this

temperature, even for holding times of less than 20 seconds and heating rates as high as

100 C/s, abnormal grain growth was observed. This behaviour is presumably associated

with the dissolution of A1N precipitates. The presence of undissolved fine A1N

precipitates, which are stable at temperatures below 1100 C, explains the inhibition of
0

49
austenite grain growth at this lower temperature interval. Interestingly, even holding at

1100C for 780 seconds has shown very little increase in austenite grain size and the

grain size distribution remains uniform. This can be considered as the critical temperature

above which very fast abnormal grain growth was observed, associated with fast

dissolution of A1N particles. In the case of A1N, the effect of temperature is stronger than

the effect of holding time, i.e. very rapid dissolution of A1N with associated abnormal

grain growth occurs when a temperature above 1100 C is reached.

The schedules selected for application to C C T and deformation-transformation tests

are shown in Table 5.1.

Table 5.1 Schedule of the heat treatment and resulting austenite grain size.

Heating Holding Holding Cooling Holding Holding


Steel Rate Temp.T, Time t, Rate Temp.T 2 Time 1 Quench2 - EQAD
[C/s] [C] [si [C/s] [C] [s]
HSLA- 5 950 90 - 950 - Water 16 19
65
5 1050 60 10 950 30 Water 41 49

5 1150 60 10 950 30 Water 108 130

DP 5 950 150 - 950 - He 13 16


600
5 1050 120 10 950 30 He 20 24

5 1100 780 10 950 30 He 27 32

The included heat treatment cycles have resulted in uniform austenite microstructures

with grain sizes (EQAD) of 16, 41 and 108 um for the HSLA-65 steel and 13, 20 and 27

50
um for the DP 600 steel respectively. Mean equivalent area diameter is converted to the

corresponding volumetric diameter by d vol =\.2dEQAD [1].

Differences in the E Q A D of the austenite grains from different fields of measurement

were observed. The etching procedure, inhomogeneous composition, etc. could be the

reason. The overall experimental error of austenite grain size measurement was estimated

to be approximately 10% for both steels.

5.2 C C T Results

5.2.1 Effect of Cooling Rate

The continuous cooling austenite decomposition kinetics of DP 600 steel with \6/um

austenite grain size at different cooling rates are shown in Fig. 5.3. The C C T curves

clearly show three stages of the process; relatively slow transformation rates in the

beginning followed by fast growth of the resulting phase and again slower rates at the

end. The resulting transformation kinetics can be explained by the following factors:

nucleation, growth, and impingement. At the beginning of transformation, nucleation

takes a leading place to early site saturation. The result is a gradually increasing

transformation rate. The process continues with rapid growth due to the steep carbon

gradient in the austenite at the a : y interface. Finally, at about 75% transformed fraction,

the transformation rate slows down significantly due to the impingement. As a result,

austenite decomposition curves display a sigmoidal shape.

51
1

o
0.8
E 1

c 0.6

O
O
O
O
\ A
A
(0 A A

O A
A
o A
c 0.4 rc/s

o A
o A


I

o
A
A
o 5C/s
o A
U A
0.2 44C/s o A

A 122C/s

227C/s A

0 T

450 500 550 600 650 700 750

Temperature [C]

Fig. 5.3 Austenite decomposition kinetics of D P 600 steel with austenite grain size of 16 um at
different cooling rates.

As illustrated in Fig. 5.3, increasing the cooling rate leads to a visible shift to lower

temperature ranges for the austenite decomposition. Since the diffusional transformation

is a process that requires time for nucleation and growth, increasing the cooling rate

decreases the time available at any given temperature to accomplish the two stages and

subsequently leads to lower temperatures of transformation. The trend is observed for the

whole range of austenite grain sizes.

The kinetics of austenite decomposition and the effect of cooling rate on HSLA-Nb

steel with d = 49//mare shown in Fig. 5.4. Similar to the DP 600 steel, increasing the
y

cooling rate leads to lower temperatures of transformation.

52
1
. % A A
A m
a o 0
A A
*A A

A
A O
A
A.
O 0.8 A
A O A. A

A
.
A O *
E
i_


A
A
O

o A O
O
A

(A A A
C 0.6 A H ' A
A
re A H

La A
h-
c 0.4 A "
'o \
A
o A rc/s A
O
O A
"4-" A
o 2.5C/s O A
o
C3
A
A
i_
0.2 18C/s A O
A
LL A 72C/S A
A
191X/S O A


. A
A 0
A
A
0
500 550 600 650 700 750 800

Temperature [C]

Fig. 5.4 Austenite decomposition kinetics of the HSLA-65 steel with austenite grain size of 49 um at
different cooling rates.

5.2.2 Effect of Initial Austenite Grain Size

Initial austenite grain size has a strong effect on continuous cooling transformation

kinetics.

Fig. 5.5 shows the transformation kinetics for the DP 600 steel with austenite grain

sizes of 16, 24 and 32 jum and a cooling rate of about 20C/s. It can be seen from the

graph that the increase in austenite grain size leads to lower temperatures of

transformation.

53
1

73
O
0.8
E
CO
0.6 -
c
1

I-
c 0.4 - o
o A d =16p-m A
D
Q
u A
y

re O dY=24nm A
0.2 - A
dy=32p.m ^ A-

A
0
500 550 600 650 700 750

Temperature [C]

Fig. 5.5 Comparison of austenite decomposition kinetics of DP 600 steel for 16, 24 and 32^m
austenite grain size at a cooling rate of about 20 C/s.
0

A similar trend was observed for the HSLA-65 steel for austenite grain sizes of 19, 49

and 130 . Fig. 5.6 shows the austenite decomposition kinetics for a cooling rate of

about 50 C/s.

54
1
AA A
A A A

AA
E 0.8 H AA
A
o A.
A
0.6 o
c o
A
A
CO o A
o A
e o
o 0.4 - oO A
A dy=19p-m A
o O dy=49|J.m
A
CO A
0.2 - Q
O A
d =130u.m
Y
Q A
Q A
O
0 A

500 550 600 650 700 750


Temperature [C]

Fig. 5.6 Comparison of austenite decomposition kinetics of H S L A - 6 5 for 19, 49 and 130 jam austenite
grain size at a cooling rate of about 5 0 C/s.
0

This effect is primarily related to the influence of grain boundary area where

nucleation of ferrite occurs. Increasing the austenite grain size decreases the grain

boundary area per unit volume. Further, growing distances are longer as the austenite

grain size increases. As a result, the transformation is delayed to lower temperatures.

5.2.3 Effect of Chemical Composition

The influence of chemical composition is shown in Fig. 5.7. The graph is plotted as

fraction transformed vs. undercooling. The comparison is presented for two cooling rates

for one austenite grain size for each steel. The DP 600 steel transforms at larger

undercooling than HSLA-65 steel, even though the DP 600 steel has a slightly smaller

initial austenite grain size.

55
SoO_oCki
D

o o
o
0.8 oo HSLA-65
o . .
d =19 p m
o
y

0.6
o 24C/s
o O 63C/s
o
c 0.4 o
o O I
D P 600
U d =16 um

Si
y

2 0.2 23C/s
67C/s

50 100 150 200 250 300

Undercooling (Ae,-T) [C]

Fig.5. 7 Comparison between austenite decomposition kinetics of D P 600 and H S L A - N b steels.

To initiate the austenite decomposition in DP 600, an undercooling of about 125 C

was needed while for the HSLA-65 steel it was about 93 C of undercooling is required

for the cooling rate of approximately 24 C/s. The difference in the chemical composition

results in this behaviour and the presence of molybdenum and niobium in solution for the

DP 600 and the HSLA-65 steels respectively defines the kinetics of the decomposition.

Both elements have a strong affinity to segregate to austenite grain boundaries and the

a : y interface. Obviously, Nb in the HSLA-65 steel shows less effect on transformation

kinetics presumably due to niobium precipitates which diminish the effect on lowering of

the transformation temperatures.

Several tests were reproduced to evaluate the experimental accuracy of determining the

transformation temperatures. The variation was estimated to be 6 C .

56
5.2.4 Transformation Start and Finish Temperatures

Transformation start temperature, T , has been determined as the temperature of 5%


s

transformation. The effect of initial austenite grain size and cooling rate on T is shown
s

in Fig. 5.8 and Fig. 5.9 for the three initial conditions for the DP 600 and HSLA-65

steels. As seen in the figures, increasing the cooling rates depresses the transformation

start temperature. For a given cooling rate, a smaller austenite grain size, leads to a higher

T . Larger grains provide less available nucleation sites thereby leading to the delay of
s

transformation start.

Transformation finish temperature is defined as the temperature at which 95%

transformation is attained. The general observation is that with increasing cooling rate,

the transformation finish temperature decreases for a given austenite grain size; for a

given cooling rate transformation finish temperature decreases with increasing austenite

grain size. This behaviour is similar to that of the transformation start temperature.

57
Fig. 5.8 Transformation start temperature vs. cooling rate for DP 600 steel.

Fig. 5.9 Transformation start temperature vs. cooling rate for HSLA-65 steel.

58
5.3 Microstructural evaluation

5.3.1 Resulting microstructure in HSLA-65 Steel

Resulting microstructures of the HSLA-65 steel for an austenite grain size of 19 um

and cooling rates of 1, 13 and 63 C/s are shown in Fig. 5.10.


0

Fig. 5.10 Microstructure after continuous cooling test for HSLA-65 having dy=19 u.m at cooling rates
of: a) 1 C/s; b) 13 C/s; c) 63 C/s.
0 0 0

As it can be seen from the micrograph, a predominantly polygonal ferrite structure is

formed up to a cooling rate of 153 C/s, the highest cooling rate investigated. The

relatively small austenite grain provides a large number of nucleation sites on grain

59
boundaries and stabilizes the polygonal ferrite formation. However, with the increase in

the cooling rate, transition to quasipolygonal ferrite was observed, Fig. 5.10 c).

Fig. 5.11 shows the microstructure of the HSLA-65 steel with a prior austenite grain

size of 49 um, obtained by applying cooling rates of 1, 18, and 72 C/s. Unlike the

microstructure for the 19 um austenite grain size, it can be seen from the micrographs

that with increasing cooling rate the polygonal ferrite fraction decreases. For 1 C/s the 0

polygonal ferrite fraction is 85%, dramatically decreases to 38% for 18C/s, and

disappears for cooling rates higher than 72 C/s.

Fig. 5.11 Microstructure after continuous cooling test for H S L A - 6 5 having oV=49 |im at cooling rates
of: a) 1 C/s; b) 1 8 C/s; c) 7 2 C/s.
0 0 0

60
A complex microstructure with polygonal ferrite and elongated grains embedded in a

dark constituent is observed in Fig. 5.11 a). This microstructure can be related to

formation of the polygonal ferrite at higher temperatures at austenite grain boundaries. As

the temperature decreases nucleation sites, such as non-metallic inclusions, in the prior

austenite grain interior become more active. Thus, formation of the acicular

microstructure is possible. Further, at lower temperatures the remaining austenite

transforms presumably into bainite or martensite.

The polygonal ferrite fraction decreases very rapidly with increasing cooling rate for

austenite grain size of 130 um. A very small fraction of polygonal ferrite (56%) was

obtained for a cooling rate of lC/s and completely disappears for cooling rates larger

than 18 C/s.

1 10 100 1000
Cooling Rate [C/s]

Fig. 5.12 Effect of austenite grain size and cooling rate on polygonal ferrite fraction for HSLA-65
steel.

61
The summarized effect of cooling rates and austenite grain sizes is given in Fig. 5.12.

As the graph indicates, the effect of cooling rate for transition from polygonal to non-

polygonal structure is stronger for larger austenite grain sizes.

5.3.2 Resulting Microstructure in DP 600 Steel

Fig 5.13 shows the microstructure obtained with the austenite grain size of 16 um and

cooling rates of 1, 23, and 67C/s.

Fig. 5.13 Microstructure after continuous cooling test for DP 600 steel having a\=16 nm at cooling
rates: a) 1 C/s; b) 23 C/s; c) 67 C/s.
0 0 0

62
At lower cooling rates, 1 and 5 C/s, an amount of about 87% polygonal ferrite fraction

was obtained for all employed austenite grain sizes of 16, 24 and 32 um. The non-

polygonal fraction dramatically increases at air cooling (23 C/s). Further increasing in

the cooling rate decreases significantly the polygonal ferrite fraction for all austenite

grain sizes. For sufficiently low transformation temperatures the formation of polygonal

ferrite is terminated and non-polygonal transformation products form (Widmanstatten

ferrite, bainite, etc.). DP 600 steel transforms at lower temperatures compared to those of

the HSLA-65 steel and thus, this effect is more pronounced.

The combined effect of the cooling rate and the austenite grain size on the polygonal

ferrite fraction is plotted in Fig. 5.14. It can be seen from the graph that the larger the

austenite grain size, the less polygonal ferrite fraction is present at a given cooling rate.

For slower cooling rates (1 and 5 C/s), however, there is no significant effect of austenite

grain size on polygonal ferrite fraction, presumably because of the narrow austenite grain

size range. Further increasing of the cooling rate leads to an increase of a non-polygonal

ferrite microstructure and the effect is stronger for the largest austenite grain size.

63
o 0H . 1 1
0 20 40 60
C o o l i n g Rate [C/s]

Fig. 5.14 Effect of austenite grain size and cooling rate on polygonal ferrite fraction for DP 600 steel.

5.3.3 Polygonal Ferrite Grain Size in HSLA-65 and D P 600 Steels

Fig. 5.15 shows the effect of cooling rate and austenite grain size on the resulting

polygonal ferrite grain size in both steels. Increasing the cooling rate and decreasing the

austenite grain size lead to a refinement of the ferrite grain size.

64
A HSLA-65
o dy=19 u:m
d=49 |im
y

DP 600
d=16 (J,m
r

d=32 jam
y

T 1

1 10 100 1000
Cooling Rate [C/s]

Fig. 5.15 Polygonal ferrite grain size vs. cooling rate for DP 600 and HSLA-65 steels.

The number of available nucleation sites is an important factor for grain refinement

during continuous cooling transformation. Considering the austenite grain boundary area

as an available nucleation site, noticeable nucleation will start first at grain corners where

the nucleation energy barrier is lowest and the probability for nucleation is highest.

Decreasing the temperature with continuous cooling, more nucleation sites are available

because of the increased driving pressure of transformation and thus less favourable sites

(grain edges and surface) can act as nucleation sites. Creating more nuclei leads to the

refinement of the final microstructure.

In addition, the effect of austenite grain size also plays an important role. The smaller

is the austenite grain size, the more grain boundary area per unit volume is available for

nucleation. More nucleation sites are available on grain boundaries resulting in smaller

ferrite grain sizes for a given cooling condition.

65
5.3.4 Hardness

The measured hardness of the three initial austenite grains of HSLA-65 steel is shown

in Fig. 5.16.

260

220 H

>
X 180 4
A
dy=19 fim
140 4 o dy=49 \im
dy=130 \im
100
0 50 100 150 200 250

Cooling Rate [C/s]

Fig. 5.16 Effect of cooling rate and austenite grain size on hardness (HSLA-65).

Increasing the austenite grain size and/or cooling rate lead to decreasing of the

transformation temperature and results in refinement of the polygonal ferrite grains or an

increase in non-polygonal fraction. Consequently, an increase in the hardness is observed.

As an example, it can be seen from the graph that increasing the austenite grain size from

19 to 130 um the hardness increases from 165 to 220 H V for a cooling rate of about

65 C/s and increasing the cooling rate from 1 to 190 C/s increases the hardness from

175 to 225 H V for d=49 um.

66
The combined effects of cooling rate and austenite grain size on hardness can be

rationalized using the transformation temperature. Fig. 5.17 shows hardness vs.

transformation start temperature for the HSLA-65 steel. It can be seen that with an

increasing transformation start temperature the hardness decreases. On the other hand, the

related final microstructure is strongly related to the transformation start temperature.

Similar trends of hardness were observed for the DP 600 steel.

260

220 H

> 180 H

A dy=19 |am
140 - o dy=49 Jim
dy=130 U-m

100
600 650 700 750 800

T s [ C ]

Fig. 5.17 Transformation start temperature vs. hardness (HSLA-65 steel).

67
5.4 Effect of Retained Strain on Continuous Cooling Transformation

5.4.1 Double Hit Compression Test

To investigate static softening, a series of double hit tests were performed. It was

necessary to find conditions which guarantee the presence of retained strain in the

austenite region in order to study its effect on transformation.

A double hit test for the HSLA-65 steel, with prestrain of 0.5, is shown in Fig. 5.18.

Very little softening, i.e. about 14%, is observed during unloading which can be

attributed to the presence of niobium in solution or as precipitates. The solute drag effect

or precipitate pinning affects the mobility of the austenite grain boundary causing

retardation of recrystallization.

250 T

200 -

Strain

Fig. 5.18 Double hit test for HSLA-65 steel (interruption time t . = 5 s, dy=19 um, strain rate=l/s,
(

E=0.5).

68
Fig. 5.19 shows an example of the softening behaviour of the DP 600 steel with

prestrain of 0.5. As it can be seen from the figure, the relative softening during the

interruption time is significant. Thus, a decrease in deformation temperature is the

solution to delay softening sufficiently as requirement for the transformation tests.

250 T

200 -

50 -

0-
0 0.2 0.4 0.6 0.8
Strain

Fig. 5.19 Double hit test for DP 600 steel (interruption time t .= 5 s, dy=16 um, strain rate=l/s, e=0.5).
(

The initial conditions for investigation of the effect of deformation on austenite

decomposition were based on the material softening before the second hit, i.e. during the

interruption time. Fig. 5.20 shows the softening behaviour in the DP 600 steel with

austenite grain sizes of 16 and 32 um. As it can be seen, decreasing the temperature of

deformation to 850 C reduces the rate of softening for the DP 600 steel with d =16 um at T

the shorter interruption times where negligible softening is recorded. Obviously, the

lower deformation temperature was more appropriate for the subsequent deformation-

69
transformation tests due to lower softening rates. For the austenite grain size of 32 um the

softening is very little even for increased interruption times of up to 20 seconds.

0 A l ; . 1 1 ; 1
0 5 10 15 20 25

Interruption Time [s]

Fig. 5.20 Relative softening (measured by offset method) for D P 600 steel (interruption time t .=2, 5,
and 20 s; strain rate=l/s, prestrain=0.5).

Several double hit tests were performed with a prestrain of 0.2 for both steels and

Tdef=900C for the HSLA-65 steel and Tdef=850C for the D P 600 steel with an

interruption time ranging from 2 to 20 seconds. As Fig.5.21 illustrates, observed

softening in both steels is about 20% which indicates usually that only recovery takes

place [79].

70
100
o D P 600
80 d =16 |J,m
y

Tdef=850C
A HSLA-65
60
c d =49 u,m
T

'E Tdef=900C
o 40
o
CO
20

0
0 5 10 15 20 25
Interruption Time [s]

Fig. 5.21 Relative softening for D P 600 and H S L A - 6 5 steels (interruption time t . =2, 5, and 20 s;
(

prestrain =0.2).

5.4.2 Effect of Retained Strain on Transformation Kinetics

Table 5.2 shows the matrix of deformation transformation tests that were performed in

order to investigate the effect of deformation on austenite decomposition kinetics,

resulting microstructure, and hardness:

Table 5.2 Deformation-transformation test schedule.

Steel dy Tdef Strain Strain Rate Cooling


[jam] [C] [s ]
]
Rate [C/s]
HSLA-65 19 900 0.2 1 10,20,30, 66
900 0.5 1 10, 30, 60
49 900 0.2 0.1 1, 5, 10, 66
900 0.5 0.1 1, 5, 10, 65
DP 600 16 850 0.2 1 1, 5, 30, 58
850 0.5 1 1, 5, 29, 60
32 850 0.2 1 1, 5, 29, 59
850 0.5 1 1, 5, 30, 60

71
Comparing the kinetics at constant strain, e.g. 0.5, the behaviour is similar to that of

undeformed conditions - increasing the cooling rate shifts the transformation to lower

temperatures as it is shown in Fig. 5.22 for DP 600 steel:

T3
o 0.8
0 A

\
w 0.6
^
A
A
O. A

c 0.4 O A
O Cooling Rate L O A
'+
u A 1C/S \ 9. A

2 0.2 O 5C/s
O ^
O A

60C/s
o A

0 o A

500 550 600 650 700 750

Temperature [C]

Fig. 5.22 Effect of cooling rate on transformation kinetics for D P 600 steel with constant deformation
(E=0.5, df=16 nm).

The same trends were observed for strain of 0.2.

The effect of deformation on transformation kinetics for the HSLA-65 steel with

austenite grain size of 19 um is shown in Fig. 5.23 for a cooling rate of about 60 C/s. As

it can be seen, the transformation occurs at higher temperature with increasing strains.

Similar observations were made for other cooling rates and austenite grain sizes.

72
oo
TJ ******
E 0.8 o
o A
0.6 A
03 8=0 A
O

o 8=0.2
o AA
c 0.4 oA
o
A 8=0.5 O A
os 0.2 OA
OA
o.
0
550 600 650 700 750 800
Temperature [C]

Fig. 5.23 Effect of retained strain on transformation kinetics for HSLA-65 steel (dT=19 |im; cooling
rate=60C/s).

An applied strain of 0.2 led to an increase in the transformation start temperature in

HSLA-65 steel with a prior austenite grain size of 49 um, as shown in Fig. 5.24. Further,

increasing the strain to 0.5 leads to additional increase in the transformation start. This

effect is stronger for higher cooling rates, e.g. at a cooling rate of about 70 C/s, the

transformation start is increased by about 40 C.

73
800

760 4

2- 720 H
(A

680 H

640
0 20 40 60 80
Cooling Rate [C/s]

Fig. 5.24 Effect of retained strain on transformation start temperature for HSLA-65 steel (dy=49 um).

Transformation start temperature as a function of strain and cooling rate for the DP 600

steel is shown in Fig. 5.25. Increased cooling rate and decreased retained strain leads to a

decrease in transformation start temperature. The effect, as it can be seen from the graphs

in Fig.5.26, is stronger for the higher cooling rates, similarly to the behaviour in the

HSLA-65 steel.

74
800
8=0
760 H
A 6=0.2

o
in
^ 680 A

640 H

600
0 20 40 60 80
Cooling Rate [C/s]

Fig. 5.25 Transformation start temperature as a function of cooling rates and retained strain for
DP 600 steel with dy=32 urn

5.4.3 Microstructure Resulting from Pancaked Austenite

5.4.3.1 DP 600 Steel

A significant influence of retained strain on the microstructure is observed for the two

steels. Fig. 5.26 shows the resulting microstructure of DP 600 steel with an austenite

grain size of 32 um and a cooling rate of 60 C/s. For the non-deformation condition,

even at a somewhat lower cooling rate (49 C/s) the dominant microstructure is non-

polygonal. Introducing a small amount of deformation increases the polygonal ferrite

fraction. Introducing a strain of 0.5 results in a significant increase of the polygonal

ferrite formation and for a strain of 0.7 the microstructure is almost completely polygonal

ferrite.

75
Fig. 5.26 Effect of residual strain on polygonal ferrite evolution of D P 600 steel with df=32 um:
a) CR=49 C/s, e=0; b) CR=60 C/s, e=0.2; c) CR=60 C/s, e=0.5; d) CR=62 C/s, e=0.7.
0 0 0 0

The effect of retained strain on the polygonal ferrite fraction is more pronounced as the

cooling rate increases. Fig. 5.27 represents an overview of the polygonal ferrite fraction

dependence on retained strain and cooling rates for the DP 600 steel with an austenite

grain size of 32 um. For slower cooling rates (1 and 5C/s) no significant changes were

observed in the amount of polygonal ferrite fraction, since even without retained strain a

predominantly polygonal ferrite microstructure forms (ferrite fraction of approximately

83%). A strain of 0.2 is not enough to create sufficiently more nucleation sites to have a

76
sufficiently larger polygonal ferrite fraction in the final microstructure. However,

applying a strain of 0.5 leads to an increase in nucleation sites and results is a

predominantly polygonal microstructure also at higher cooling rates.

For smaller austenite grains of 16 um, as summarised in Fig. 5.28, an applied strain of

0.2 is sufficient to increase the polygonal ferrite formation at higher cooling rates, i.e. the

polygonal ferrite fractions are 80% and higher for cooling rates of 30 and 60C/s. For

slower cooling rates there are no significant changes; the polygonal ferrite fraction is

about 88% independent of the strain.

77
5.4.3.2 HSLA-65 Steel

No significant effect of retained strain is observed on the amount of polygonal ferrite

in the HSLA-65 steel with an austenite grain size of 19 urn, as shown in Fig. 5.29. The

fraction for all experiments remains above 94%. However, with the increase of the strain

transition from quasipolygonal to completely polygonal microstructure occurs.

78
Fig. 5.29 Effect of retained strain on ferrite grain size of HSLA-65 with dy=19 um, C R K 6 0 C / S :
a) e=0; b) e=0.2; c) e=0.5.

For the larger austenite grain of 49 um in the HSLA-65 steel, retained strain facilitates

formation of polygonal ferrite as shown in Fig. 5.30. For instance, increasing the

deformation level from 0 to 0.5 results in an increase of polygonal ferrite fraction from

8% to 51% at a cooling rate of approximately 60 C/s. The effect is stronger as the

cooling rate increases. The trend observed for the larger austenite grains in HSLA-65

steel is similar to that of DP 600 steel with an austenite grain size of 32 um; more stored

79
energy (at strains above 0.2) is needed to increase significantly the polygonal ferrite

fraction.

5.4.4 Polygonal Ferrite Grain Size

The polygonal ferrite grain refinement for the HSLA-65 steel with an austenite grain

size of 19 um is summarized in Fig. 5.31. Increasing the cooling rate and/or the strain

result in a decrease of the ferrite grain size. It can be seen from the figure that the grain

size obtained with a cooling rate of about 60 C/s is reduced to 5 um when a strain of 0.5

has been applied.

80
Q_ 4 -I , , _, , , 1
0 10 20 30 40 50 60 70
C o o l i n g Rate [C/s]

Fig. 5.31 Ferrite grain refinement in H S L A - 6 5 steel with dy=19 um.

The effect of retained strain and cooling rate on ferrite grain sizes in the DP 600 steel

is shown in Fig. 5.32. Increasing the strain from 0 to 0.5 resulted in the refinement of the

ferrite grains. For higher cooling rates, increasing the retained strain to 0.5 refines the

ferrite grain size to about 4 um.

The experimental error of polygonal ferrite grain size measurement was estimated to be

approximately 10%.

81
0 20 40 60
Cooling Rate [C/s]

Fig. 5.32 Ferrite grain refinement in DP 600 steel with dy=16 nm.

5.4.5 Effect of Deformation on Hardness

Increasing the amount of retained strain results in a decrease in hardness. Accelerated

cooling leads to increased hardness in case of constant strain. Fig. 5.33 demonstrates the

influence of deformation and cooling rate on hardness in the DP 600 steel samples with

an initial austenite grain size of 32 um. The trend can be explained (as in the case of the

effect of the austenite grain size in non-deformation conditions on the hardness) with the

transformation start temperature. Increasing strain leads to higher transformation

temperature and the formation of more polygonal ferrite, i.e. a softer microstrucure.

82
240

220

200 H

^ 180

160
lit D
6=0
1 0 8=0.2
A 6=0.5
120
20 40 60 80

C o o l i n g Rate [ C / s ]

Fig5.33 Effect of deformation and cooling rate on hardness for DP 600 steel with dy= 32 nm.

A similar trend was observed for the HSLA-65 steel.

Fig. 5.34 shows the relationship between the transformation start temperature and

hardness for the DP 600 steel with austenite grain sizes of 16 and 32 um for strain and

strain-free cases. It can be see that the increase in transformation start, as a result of

applied deformation, leads to softer microstructure. Similar trend is observed for the

HSLA-65 steel.

83
84
CHAPTER 6- MODELING

6.1 Kinetics of Transformation from Austenite to Polygonal Ferrite

6.1.1 Transformation Start Temperature

An application of Eq. 2.11 (Section 2.4.3.1, p.32) is given in Fig. 6.1 where the

experimentally obtained transformation start temperature, T , and the model for the DP
s

600 steel are compared. For M a number of two is applied. The values of diffusivity,

which are temperature and carbon content dependent, have been calculated by using

Agren approach [83]. T H E R M O C A L C software has been used to calculate the carbon

(ortho-) equilibrium concentration in ferrite, C , and austenite, C . The parameters T


a y N

and C * are used as fitting parameters. T = 7 3 2 C has been estimated from the
N

transformation start for the cooling rate of lC/s. The second parameter which is the

critical carbon content,C*, shows the best fit at 1.5C . As it can be seen, for the largest
0

austenite grain size and highest cooling rates where a complex microstructure was

obtained, deviation from the model was observed. This can be explained using a more

complex model as proposed by Militzer at al. [26] but does not appear to be significant

for the run-out table cooling conditions where the formation of more than 80% ferrite is

the target.

85
300

q>dy 2
[Cs-Vm ]2

Fig. 6.1 Model of transformation start temperature for DP 600 steel.

The prediction of transformation start temperature for the HSLA-65 steel is shown in

Fig. 6.2. Similar to the transformation start model for the DP 600 steel, the fitting

parameters, T = 8 1 9 C and C* = 1.8C , were adjusted to measured values. As it can be


N 0

seen, a reasonable description of transformation start can be obtained except for the larger

austenite grain size, similarly as discussed for the DP 600 steel. It is more relevant to

compare AT = A - T
N N for the two steels. For the DP 600 steel, ATN is larger

presumably due to a strong effect of molybdenum on the delay of the transformation.

86
300 7

Fig. 6.2 Model of transformation start temperature for HSLA-65 steel.

The incorporation of the effect of retained strain into the model for the DP 600 steel

can be seen in Fig. 6.3. The magnitude of C * and T was kept constant with the same
N

value as in the case of no-strain conditions. To include the effect of deformation, an

effective grain size is introduced in the form of: d eff = d exp(-e) as described in
y

literature review [47]. The model is then in reasonable agreement with the experimental

data.

87
cpd [Cs-Vm ]
eff
2

Fig. 6.3 Transformation start temperature for DP 600 steel with incorporated retained strain.

Fig. 6.4 shows the effect of deformation on the applied transformation start model for

the HSLA-65 steel. Similarly to the DP 600 steel reasonable results were obtained.

88
(pd eff [C/s- Lim ]
1 2

Fig. 6.4 Transformation start temperature model for H S L A - 6 5 steel including deformation.

6.1.2 Kinetics of Austenite-to-Polygonal Ferrite Transformation

The model to describe the formation of polygonal ferrite was developed based on those

transformation data which were associated with microstructures that have at least 85%

polygonal ferrite fraction. To test the applicability of the model for larger austenite grain

size, a few cases with polygonal ferrite fractions as low as 75% have been considered.

The polygonal ferrite fraction obtained at continuous cooling was normalized to the

equilibrium fraction at each temperature increment. T H E R M O C A L C software was used

to find the equilibrium data for each steel and the following fitting equations for (ortho-)

equilibrium ferrite fraction can be used:

x
e q . = 0-9(1 - exp(-0.02(^ - T))) for HSLA-65 steel

and

89
for DP 600 steel
0.92

X eq .=l-exp(-exp(-3.54(4 -r))3

where T is the temperature.

The fraction of polygonal ferrite transformed can be described with the J M A K model

(Eq. 2.6) and adopting the additivity rule. Eq. 2.10 can be transformed into:

6.1

which can be used to determine the experimental values of b for a given n.

Initially an analysis of the CCT test was performed in which n was varied in the range

0.8-1.2 to find an n value for which b is independent of cooling rate. For both of the

steels the best fit is obtained for n= 0.9 which was also reported for a series of low carbon

and Nb-bearing steels in previous work [10, 26]. The actual cooling path was used to

determine b as a function of temperature. As an example, Fig. 6.5 illustrates the

experimental values of \nb vs. temperature for the HSLA-65 steel with =0.9 and an

austenite grain size of 19/um . As in previous work [10, 26], it can be assumed as a first

approximation that Xnb is a linear function of the temperature and the whole series of

cooling rates for a given austenite grain size can be described by:

\nb = AT + Y (6.2)

90
1H

-1 H
c 3C/s
-2 10C/s
o 24C/s
lnb=-0.018T+12.5
-3 47C/s
o 63C/s
-4H 124C/s

153C/s
-5

Fig.6.5 Expression of Inb as a linear function of temperature for H S L A - 6 5 steel with n=0.9 and
austenite grain of 19 um at different cooling rates.

Table 6.1 summarizes the values for the slopes A and the intercepts Y obtained for

different CCT tests in both steels.

Table 6.1 Slopes and intercepts of Inb for D P 600 and H S L A - 6 5 steels.

Steel dy^m] A Y
DP 600 16 11.93
24 -0.020 11.41
32 11.05
HSLA-65 19 -0.018 12.50
49 9.50

The slope A is independent of austenite grain size but the intercept Y decreases with

increasing d . To account for this in the J M A K equation, the grain size modified

91
Umemoto equation (Eq. 2.12, Section 2A3.2) was applied to experimental results

including the effect of austenite grain size. Based on Eq.2.12 it can be written:

(6.3)

or

In6 = In6, -m\nd y (6.4)

From this relationship it was possible to determine m and In 6, by using the 6-data

(intercept) presented in Table 6.1, as shown in. Fig. 6.6:

Fig. 6.6 Derivation of parameters for Umemoto equation.

Thus,ln& can be expressed as:

In b = -0.0187 + 20.8 - 2 . 9 In d y for HSLA-65 steel (6.5)

92
and

lnfe = -0.02r + 15.4-1.31nrf r for DP 600 steel. (6.6)

The applications of these relationships for lnfrare shown in Fig. 6.7 and 6.8 for both

steels. Reasonable agreement of measured and predicted transformation kinetics is

observed. For all cases, except for the HSLA-65 steel with d =\9 y /urn where the model

was applied for the entire range of cooling rates, the model was applied for lower cooling

rates because the polygonal ferrite was predominantly present at slow cooling ranges.

Fig.6.7 Application of Umemoto grain size modified J M A K model for DP 600 steel.

93
1

I-

0
580 620 660 700 740 780
Temperature [C]

Fig.6.8 Application of Umemoto grain size modified J M A K model for HSLA-65 steel.

The next step for using the J M A K equation is to incorporate the role of retained strain.

Applying the approach of Lacroix et al. [47], an effective grain size is introduced, i.e.

d eff - d exp(-,?), to obtain a new relationship for In b :

\nb = AT + Y-m\n[d (-e)]


y (6.7)

The model predictions for DP 600 steel {d =32//m) at a cooling rate of 5C/s are
y

shown in Fig. 6.9. It can be seen that there is good agreement between the experimental

data and the model. Similar results were observed for all other cases except for austenite

grain size of 16//mand the lowest cooling rate of lC/s where the predicted

transformation rate was higher than the experimental one. This can presumably be

94
explained by the restoration of the deformed microstructure due to the lack of precipitates

and subsequent static recovery and recrystallization occurring. This is consistent with the

softening kinetics shown in Fig. 5.18 for the austenite grain size of\6pm. There is about

40% softening for a holding time of 5 seconds, whereas for 32 pm holds of 20 seconds

yields to softening of less than 20%. Thus, this series should be excluded from any

further analysis. More experiments are necessary to investigate and confirm the

magnitude of the restoration between the end of deformation finish and transformation

start.

550 600 650 700 750

Temperature [C]

Fig.6.9 Application of J M A K model incorporating the effective austenite grain for DP 600 steel
(dy=32 \i.m and cooling rate 5 C/s).

Fig. 6.10 compares the model predictions with the experimental data for HSLA-65 steel

having an austenite grain size of \9pm, at a cooling rate of about 60C/s, and different

95
levels of retained strain; these parameters are typical for industrial run-out table

operations. Again excellent agreement of predicted with measured kinetics is observed.

Fig.6.11 shows the comparison between the measured and predicted temperatures at

which 50% fraction transformed is observed. The data include the entire range of

predominant polygonal ferrite microstructures, including those obtained from pancaked

austenite and excluding the data for the DP 600 with austenite grain size of 16/um,

applied strain of 0.2 and 0.5 and cooling rate of lC/s (as mentioned before). For both

steels excellent agreement within 6 C is obtained.


0

96
600 650 700 750 800
Tso [C]-Measured

Fig. 6.11 Comparison between measured and predicted temperature for 50% transformation.

6.2 Polygonal Ferrite Grain Size

The ferrite grain size model has been applied to those cases with microstructures that

consist of at least 85% polygonal ferrite. For the strain-free cases, the data are fairly

limited. In the HSLA-65 steel the resulting microstructure was polygonal ferrite only for

an austenite grain size of 19 um even at very high cooling rates, the model could be

applied for the whole range of cooling rates. The data for the three austenite grain sizes of

the DP 600 steel were limited to the cooling rates of 5 C/s and lower. In all other cases,
0

transition to non-polygonal microstructure occurred in early transformation stages and the

model could not be applied.

97
The approach of retained strain case is similar to that of the strain-free, i.e. Eq. 2.13.

The calculation of B (Eq. 2.14) showed a decrease with increasing of the strain. The

changes of B vs. retained strain are shown in Fig. 6.12 for the HSLA-65 steel:

CQ

0 0.1 0.2 0.3 0.4 0.5 0.6


Strain

Fig. 6.12 Changes in value of B vs. strain in HSLA-65 steel.

To express the values of B as a function of the combined effect of the austenite grain size

and deformation, again the effective grain size [47] can be used such that B takes the

form:

B = E[d exp(-f)]
r 6.12

For the DP steel the approach was applied for strain of 0.5 where more experimental data

were available. The fitting parameters g,E , and E are summarized in Table 6.2.

98
Table 6.2 Parameters describing the polygonal ferrite grain size.

Steel q E

DP 600 0.019 45.2 41000

HSLA-65 0.056 30.2 31000

The ferrite grain size model with the influence of deformation is compared with the

experimental results for DP 600 and HSLA-65 steels in Fig. 6.13 and Fig.6.14; good

agreement is achieved for both steels.

660 680 700 720 740 760


Transformation Start Temperature [C]

Fig. 6.13 Comparison between the measured and predicted values of ferrite grain size of DP 600 steel
with applied deformation of 0.5.

99
Fig. 6.14 Comparison between the measured and predicted values of ferrite grain size of HSLA-65
steel.

A comprehensive comparison between the predicted and the experimental grain size is

given in Fig.6.15. Good reproduction was in particular obtained for the smaller ferrite

grain size range which is desirable in the industrially produced hot band.

100
2 4 6 8 10 12

Measured EQAD [fim]

Fig. 6.15 Comparison between the measured and predicted values of ferrite grain size for both steels.

101
C H A P T E R 7- C O N C L U S I O N S A N D F U T U R E W O R K

7.1 Conclusions

1. The austenite grain size and cooling rate have shown great influence on austenite

decomposition in both steels. Increasing the cooling rate for a given austenite grain size

results in decreasing of transformation start temperature with subsequent effect on the

final microstructure, i.e. refinement of the ferrite grain size. Beyond a certain cooling

rate, formation of non-polygonal structure starts and its fraction increases with cooling

rate. The increase in the austenite grain size reduces the transformation start temperature

and leads to a more non-polygonal microstructure.

2. Chemical composition strongly affects the kinetics of austenite decomposition.

Molybdenum in DP 600 steel shows a greater influence in retardation the transformation

than Nb in HSLA-65 steel for the same thermal conditions, presumably because of the

stronger solute drag effect of Mo. However, not all Nb has been in solution during the

transformation tests which may have decreased its efficiency in retardation of the

transformation.

3. The effect of deformation in the interval between T and A


nr leads to an increase in

nucleation rate of allotriomorph ferrite and an associated increase in transformation start

temperature. As a result, an increase of polygonal ferrite fraction and additional ferrite

grain refinement occur at elevated cooling rates. This effect is stronger for larger

austenite grain sizes.

4. The effect of processing variables shows that the hardness of both steels increases

with the increase of austenite grain size and cooling rate and with the decrease of the

102
strain attributed to the increase in non-polygonal structures which show a higher

hardness. This effect can be rationalized with transformation temperature, e.g. a

decreasing transformation start temperature leads to higher hardness values.

5. The transformation start temperature of allotriomorph ferrite was described with a

fundamental model based on nucleation and early growth. A n application of effective

grain size was used to incorporate the effect of retained strain in austenite in the model.

6. The J M A K model has been used to describe continuous cooling transformation

kinetics by adopting the rule of additivity. A grain-size modified Avrami equation has

been used to incorporate the effect of austenite grain size on transformation rate. For the

case of retained strain, the effective grain size approach has been applied. Good

agreement of measured and predicted transformation kinetics has been observed for both

steels.

7. A semi-empirical model has been applied to predict the polygonal ferrite grain size

as a function of transformation start temperature. The effect of retained strain can again

be rationalized with using the effective austenite grain size. Comparison between the

experimental data and the model shows reasonable agreement.

7.2 Future Work

1. The kinetics of polygonal ferrite formation has been determined. However, the

second products need more consideration to be specified and quantified in terms of

morphology, kinetics of formation, and structure.

2. Additional work is required to establish the conditions for transition from polygonal

to non-polygonal transformation products.

103
3. A modified model of transformation start temperature for the larger austenite grain

size for both steels needs to be applied due to the complexity of the resulting

microstructure.

104
REFERENCES

1. A.Guimelli, M.A.Sc. Thesis, The University of British Columbia, Vancouver, Canada,


1995

2. J.W.Christian, The Theory of Transformations in Metals and Alloys, 2 nd


ed., Pergamon
Press, Oxford, U K , 1975

3. G.R.Speich, T.M.Scoonover, Microstructure and Properties of HSLA Steels, ed.


A.J.DeArdo, Jr, TMS, Pittsburgh, PA, 1988, p. 263

4. E.V.Pereloma, J.D.Boyd, Materials Science and Technology, Vol. 12, 1996, p. 808

5. R.Honeycombe, H.K.D.H.Bhadeshia, Steels-Microstructure and properties, Edward


Arnold- a Division of Ffodder Headline PLC, London, U K , 1995

6. M.Atkins, Atlas of Continuous Cooling Transformation Diagram for Engineering


Steels, A S M , Metal Park, O H , 1980, p. 20

7. D.A.Porter, K.E.Easterling, Phase Transformations in Metals and Alloys, 2 nd


ed.,
Chapman & Hall, London, U K , 1992

8. A.K.Sinha, Ferrous Physical Metallurgy, Butterworths, Stoneham, M A , 1988

9. A.Dube, H.I.Aaronson, R.F.Mehl, Revue de Metallurgie, Vol. 3, 1958, p. 201

10. R.Pandi, Ph.D. Thesis, The University of British Columbia, Vancouver, Canada, 1998

11. H.I. Aaronson, Decomposition of Austenite by Diffusional Processes, ed. V.F.Zackay


and H.I.Aaronson, Interscience Publishers, N Y , 1960, p. 387

12. W.T.Reynolds, Jr., M.Enomoto, H.I.Aaronson, Phase Transformations in Ferrous


Alloys, ed. A.R.Marder and J.I.Goldstein, T M S - A I M E , Philadelphia, P A , 1984, p. 155

13. G.R.Speich, L.J.Cudy, C.R.Gordon, A.J.DeArdo, Phase Transformations in Ferrous


Alloys, ed. A.R.Marder and J.I.Goldstein, TMS-AIME, Philadelphia, PA, 1984, p. 341

14. M.Enomoto, W.F.Lange III, H.I.Aaronson, Metallurgical Transactions, Vol. 17A,


1986, p. 1399

15. M.Enomoto, H.I.Aaronson, Metallurgical Transactions, Vol. 17A, 1986, p. 1381

16. W.F.Lange, III, M.Enomoto, H.I.Aaronson, Metallurgical Transactions A, Vol. 19A,


1988, p. 427

105
17. M.Militzer, R.Pandi, E.B.Hawbolt, Metallurgical and Materials Transactions A,
Vol. 27A, 1996, p. 1547

18. H.I.Aaronson, C.Laird, K.R.Kinsman, Phase Transformations, A S M , Metals Park,


OH, 1968, p.313

19. E.P.Simonen, H.I.Aaronson, Metallurgical Transactions, Vol. 4, 1973,p. 1239

20. J.R.Bradley, J.M.Rigsbee, H.I.Aaronson, Metallurgical Transactions A, V o l 8A, 1977,


p. 323

21. C.Atkinson, H.B.Aaron, K.R.Kinsman, H.I.Aaronson, Metallurgical Transactions,


Vol.4, 1973, p. 783

22. J.R.Bradley, H.I.Aaronson, Metallurgical Transactions, Vol. 12A, 1981, p. 1729

23. C.Zener, Journal of Applied Physics, Vol. 20, 1949, p. 950

24. R.G. Kamat, Ph.D. Thesis, The University of British Columbia, Vancouver, Canada,
1990

25. K.R.Kinsman, H.I.Aaronson, Transformation and Hardenability in Steels, Climax


Molybdenum Corporation, Ann Arbor, MI, 1967, p. 39

26. M.Militzer, E.B.Hawbolt, T.R.Meadowcroft, Metallurgical and Materials


Transactions, Vol. 31 A , 2000, p. 1247

27. M.Militzer, E.B.Hawbolt, T.R.Meadowcroft, Phase Transformation During the


Thermal/Mechanical Processing of Steel, ed. E.B.Hawbolt and S.Yue, The Metallurgical
Society of CIM, Montreal, PQ, 1995, p. 445

28. R.Pandi, M.Militzer, E.B.Hawbolt, T.R.Meadowcroft, Phase Transformation During


the Thermal/Mechanical Processing of Steel, ed. E.B.Hawbolt and S.Yue, The
Metallurgical Society of C I M , Montreal, PQ, 1995, p.459

29. A.P.Coldren, R.L.Cryderman, M.Semchyshen, Steel Strengthening Mechanisms,


A M A X , 1969, p. 17

30. P.Heedman, A.Sjostrom, M.Jarl, HSLA Steels Technology and Applications, organized
by M.Korchinsky, Philadelphia, PA, 1983, A S M , p. 121

31. J.D.Grozier, Microalloying'75, Union Carbide Corp., New York, N Y , 1977, p. 241

32. A.P.Coldren, G.Tither, Journal of Metals, Vol. 30, #4, 1978, p.6

106
33. J.Neutjens, Ph.Harlet, Th.Bakolas, P.Cantinieaux, Mechanical Working and Steel
Processing Conference Proceedings, Vol. X X X V I , ISS, Warrendale, PA, 1998, p. 311

34. G.Thewils, Materials Science and Technology, Vol. 10, 1994, p. 110

35. N.Nakata, M.Militzer, Mechanical Working and Steel Processing Conference


Proceedings, Vol. X X X V I I I , ISS, Warrendale, PA, 2000, p. 813

3 6. M . Umemoto, N.Komatsubara, I. Tamura, Journal of Heat Treating, Vol. 1, 1980,


p. 57

37. R.K.Amin, F.B.Pickering, Thermomechanical Processing of Microalloyed Austenite,


ed. A.J.DeArdo, G.A.Ratz, P.J.Wray, Pittsburgh, PA, 1981, p. 377

38. E.Essadigi, M.G.Akben, J.J.Jonas, Phase Transformations in Ferrous Alloys, ed.


A.R.Marder and J.I.Goldstein, The Metallurgical Society of A I M E , Philadelphia, PA,
1983, p. 391

39. L.E.Collins, J.R.Barry, J.D.Boyd, Phase Transformations in Ferrous Alloys, ed.


A.R.Marder and J.I.Goldstein, The Metallurgical Society of A I M E , Philadelphia, PA,
1983,p. 397

40. D.M.Fegredo, J.D.Boyd, M.J.Stewart, HSLA Steels Technology and Applications, ed.
M.Korchinsky, A S M , Philadelphia, PA, 1983, p. 95

41. F.Zhang, J.D.Boyd, Phase Transformation During the Thermal/Mechanical


Processing of Steel, ed. E.B.Hawbolt and S.Yue, The Metallurgical Society of CIM,
Montreal, PQ, 1995, p. 147

42. V.M.Khlestov, E.V.Konopleva, H.J.McQueen, Canadian Metallurgical Quarterly,


Vol. 35, No 2, 1996, p. 169

43. A.Sandberg, W.Roberts, Thermomechanical Processing of Microalloyed Austenite,


ed. A.J.DeArdo, G.A.Ratz, P.J.Wray, 1981, The Metallurgical Society of A I M E ,
Pittsburgh, PA, p. 405

44. C.Ouchi, T.Sampei, I.Kozasu, Transactions ISIJ, Vol.22, 1982, p. 214

45. M.Umemoto, H.Ohtsuka, I.Tamura, Transactions ISIJ, Vol. 23, 1983, p. 775

46.I.Kozasu, C.Ouchi, T. Sampei, T.Okita, Microalloying'75, Union Carbide Corp., New


York, N Y , 1977, p. 100

47.S. Lacroix, Y . Brechet, M . Veron, D. Quidort, M . Kandel, T. lung, Materials Science


& Technology 2003 Conference Proceedings, Section Austenite Formation and
Decomposition, ISS and TMS, Warrendale, PA, 2003, p. 367

107
48. J.B.Gilmour, G.R.Purdy, J.S.Kirkaldy, Metallurgical Transactions, Vol. 3A, 1972,
p.3213

49. M.P.Puls, J.S.Kirkaldy, Metallurgical Transactions, Vol. 3A, 1972, p. 2777

50.1.Tamura, C.Ouchi, T.Tanaka, H.Sekine, Thermomechanical Processing of High


Strength Low Alloy Steels, Butterworths, London, U K , 1988

51. K.Lucke, K.Detert, Acta Metallurgica, Vol. 5, 1957, p . l 18

52. G.R.Purdy, Acta Metallurgica, Vol. 26, 1978, p.487

53. A.P.Coldren, G.T.Eldis, Journal of Metals, Vol. 32, #3, 1980, p. 41

54. L.E.Collins, W.J.Liu, Phase Transformation During the Thermal/Mechanical


Processing of Steel, ed. E.B.Hawbolt and S.Yue, The Metallurgical Society of CIM,
1995, p.419

55. M.H.Thomas, G.M.Michal, Solid-Solid Phase Transformation; eds. H.I.Aaronson,


D.E.Laughlin, R.F.Sekerka, C.M.Wayman, TMS-AIME, Warrendale, PA, 1981, p.469

56. C.Fossaert, G.Rees, T.Maurickx, H.K.D.H.Bhadeshia, Metallurgical and Materials


Transactions A, Vol. 26A, 1995, p. 21

57. G.L.Fisher, R.H.Geils, Transactions TMS-AIME, Vol. 245, 1969, p. 2405

58. D.Webster, J.H.Woodhead, Journal of Iron and Steel Inst., Vol. 202, 1964, p. 987

59. Y.Ohmori, Transactions ISIJ, Vol.11, 1971, p. 333

60. A.N.Kolmogorov, Izv. Akademii Nauk USSR Ser. Matemat, Vol. 1, 1937, p. 355

61. W.A.Johnson, R.F.Mehl, Transactions AIME, Vol. 135, 1939, p.416

62. M.Avrami, Journal of Chemical Physics, Vol. 7, 1939, p.l 103

63. M.Avrami, Journal of Chemical Physics, Vol. 8, 1940, p. 212

64. M . Avrami, Journal of Chemical Physics, Vol. 9, 1941, p. 177

65. M.Umemoto, A.Hiramatsu, A.Moriya, T.Watanabe, S.Nanba, N.Nakajima, G.Anan,


Y.Higo, ISIJ International, Vol. 32, 1992, p. 306

66. E.Scheil, Arch. Eisenhiittenwesen, Vol. 8, 1935, p. 565

67. J.W.Cahn, Acta Metallurgica, Vol. 4, 1956, p. 572

108
68. M.Lusk, H . Jou, Metallurgical and Materials Transactions A, Vol. 28 A , 1997, p.287

69. M.Umemoto, K.Horiuchi, I.Tamura, Transactions ISIJ, Vol. 23, 1983, p. 690

70.I.Tamura, Thermec-88, The Iron and Steel Institute of Japan, Tokyo, Japan, 1988, .
p.l

71. E.B.Hawbolt, B.Chau, J.K.Brimacombe, Metallurgical Transactions, Vol. 14A, 1983,


p. 1803

72. E.B.Hawbolt, B.Chau, J.K.Brimacombe, Metallurgical Transactions, Vol. 16A, 1985,


p. 565

73. J.H.Hollomon, L.D.Jaffe, M.R.Norton, Transactions AIME, Vol. 167, 1946, p. 419

74. G.K.Manning, C.H.Lorig, Transactions AIME, Vol. 167, 1946, p. 442

75. J.B.Leblond, J.Devaux, Acta Metallurgica, Vol. 32, 1984, p.137

76. M.Militzer, E.B.Hawbolt, T.R.Meadowcroft, HSLA Steel'95, Ed. By L.Guoxun,


H.Stuart, Z.Hongtao, L.Chenggji, China Science & Technology Press, Beijing, China,
1995, p.271

77. M . Suehiro, K.Sato, Y.Tsukano, H.Yada, Y.Matsumura, Transactions of ISIJ, Vol. 27,
1987, p. 439

78. H.L.Andrade, M.G.Akben, J J Jonas, Metallurgical Transactions A, Vol. 14A, 1983,


p.1967

79. R.A.Djaic, J.JJonas, Journal of the Iron and Steel Institute, Vol. 210, 1972, p. 256

80.S.J.Lechuk, M.A.Sc. Thesis, The University of British Columbia, 2000

81. Annual Book of A S T M Standards, vol. 3.01, ASTM Standard Designation, E 112-88,
Philadelphia Pa, 1994, p.227

82. M.Militzer, A.Giumelli, E.B.Hawbolt, T.R.Meadowcroft, Metallurgical and Materials


Transactions A, Vol. 27 A , 1996, p. 3399

83. J.Agren, Scripta Metallurgica, Vol. 20, 1986, p.1507

109

You might also like