Sunteți pe pagina 1din 144

AUSTENITE GRAIN GROWTH KINETICS

AND THE GRAIN SIZE DISTRIBUTION

BY

A L A N GIUMELLI

B.Eng.,Hons., The University of Wollongong, Australia, 1991.

A THESIS SUBMITTED IN P A R T I A L F U L F I L L M E N T OF

T H E REQUIREMENTS FOR T H E D E G R E E OF

M A S T E R S OF APPLIED SCIENCE

in

T H E 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

T H E UNIVERSITY OF BRITISH C O L U M B I A

APRIL, 1995.

Alan Giumelli, 1995.


In presenting this thesis in partial fulfilment 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 financial gain shall not be allowed without my written
permission.

Department of ^(/fPr(_s 4. r^fKSXi^ Acs c r y - f C r ^ C i T n-<r~*(r

The University of British Columbia


Vancouver, Canada

Date h/^ / .

DE-6 (2/88)
ABSTRACT

The aims of this work are to assess the kinetics of austenite grain growth, and to

estimate the three dimensional grain size distribution. Measurements are to be used for the

validation of a statistical model for the estimation of the kinetics of austenite grain growth

during hot strip rolling. A series of tests are performed with two industrial plain carbon

steels in the temperature range from 950C to 1150C. Grain size is measured by standard

manual procedures and by the use of image analysis equipment. Systematic errors in the

measurements are identified. A correction is proposed to account for the systematic error

in the measurement of the two dimensional grain size distribution by the use of image

analysis. The kinetics of austenite grain growth are discussed and the influence of second

phase particles is observed. Results are described well by a statistical model for grain

growth. A number of methods for the estimation of the three dimensional grain size

distribution are compared. It is concluded that the method proposed by Matsuura and Itoh

is the most suitable.

ii
T A B L E OF CONTENTS

Page

ABSTRACT ii

LIST OF T A B L E S vi

LIST OF ILLUSTRATIONS viii

NOMENCLATURE x

ACKNOWLEDGMENTS xiii

C H A P T E R 1. INTRODUCTION 1

C H A P T E R 2. L I T E R A T U R E R E V I E W 4

2.1 Modelling of Hot Strip Rolling of Steel 4

2.2 Modelling of Grain Growth 5

2.2.1 The Power Law for Grain Growth 5

2.2.2 Statistical Modelling of Grain Growth 8

2.3 Normal and Abnormal Grain Growth 12

2.4 Measurement of Grain Size 14

2.5 The Grain Size Distribution 17

2.6 Estimation of the Three Dimensional Grain Size Distribution 20

2.6.1 Methods Based on Spherical Grain Shape 20

2.6.2 The Method of Takayama et al 24

2.6.3 The Method of Matsuura and Itoh 26

C H A P T E R 3. SCOPE A N D OBJECTIVES 31

C H A P T E R 4. E X P E R I M E N T A L 32

4.1 Sample Composition and Specimen Design 32

4.2 Equipment for the Thermal Treatment : 35

4.2.1 The Gleeble 1500 Thermomechanical Simulator 35

4.2.2 The Vertical Tube Furnace 38


iii
4.3 The Thermal Cycles 42

4.4 Revealing the Prior Austenite Microstructure 46

4.5 Measurement of the Austenite Grain Size 46

C H A P T E R 5. RESULTS A N D DISCUSSION I.

T H E KINETICS OF AUSTENITE G R A I N G R O W T H 50

5.1 The Structure of the Quenched and Etched Specimens 50

5.1.1 Helium Quenched A36 Steel 50

5.1.2 Water Quenched A36 Steel 52

5.1.3 Water Quenched D Q S K Steel 53

5.1.4 Water Quenched 1080 Steel 55

5.2 Measurement of the Grain Size and Estimation of the Error 56

5.2.1 Interpretation of the Structure 56

5.2.2 Measurement of the Grain Size 57

5.2.3 The 95% Confidence Intervals 59

5.2.4 Systematic Errors in Measurement of Equivalent Area Diameter.... 60

5.2.5 Systematic Errors in Measurement of Linear Intercept Length 64

5.3 The Kinetics of Austenite Grain Growth 64

5.3.1 A36 Steel Heated to Temperature at 5C/second 64

5.3.2 D Q S K Steel Heated to Temperature in the Tube Furnace 67

5.4 Heating Rate Effects 69

5.5 Models Describing the Kinetics of Grain Growth 70

5.5.1 Application of the Power Law 70

5.5.2 Application of the Statistical Model 72

5.6 The Influence of Second Phase Particles 77

5.7 The Evolution of the Grain Size Distribution 80

5.9 Solution Treatment 86

iv
5.10 Grain Growth During Hot Strip Rolling 89

C H A P T E R 6. RESULTS A N D DISCUSSION II.

E S T I M A T I O N OF T H E T H R E E D I M E N S I O N A L G R A I N SIZE DISTRIBUTION.... 91

6.0 Introduction 91

6.1 Methods Based on the Assumption of Spherical Grain Shape 91

6.2 The Method of Takayama et al 94

6.3 The Method of Matsuura and Itoh 100

6.3.1 Application of the Method 100

6.3.2 The Shape of the Estimated 3-D Distribution 105

6.3.3 Convergence of the Distribution With Number of Size Classes 107

6.4 Comparison of the Methods of Takayama et al. and Matsuura and Itoh 110

6.5 Validation of the Method of Matsuura and Itoh With Measured 3-D

Results Ill

C H A P T E R 7. S U M M A R Y A N D CONCLUSIONS 114

BIBLIOGRAPHY 118

APPENDIX 124

v
LIST OF TABLES

Table Page

Table 2.1 - Grain growth exponents for the power law, for zone refined metals 7

Table 2.2 - Power law parameters reported for austenite grain growth 8

Table 2.3 - Description of geometric and arithmetic size class scales 19

Table 4.1 - Composition of steels to be used in this investigation 32

Table 4.2 - Standard thermal cycles performed on the Gleeble 42

Table 4.3 - Standard furnace thermal cycles 44

Table 4.4 - Stepped Gleeble thermal cycles 44

Table 4.5 - Continuous cooling thermal cycles 46

Table 4.6 - Summary of etching procedures 48

Table 5.1 - Correction of systematic errors for the mean equivalent area diameter. 63

Table 5.2 - Power law fitting parameters for the A36 and D Q S K steels 72

Table 5.3 - Evolution of the grain size distribution 81

Table 6.1 - The 3-D grain size distribution from Takayama et al 95

Table 6.2 - The 3-D grain size distribution using the image analyser results

from Matsuura and Itoh 102

Table 6.3 - The 3-D grain size distribution from the corrected 2-D distribution

from Matsuura and Itoh 103

Table 6.4 - The 3-D grain size distribution predicted using the method of

Takayama et al. and Matsuura and Itoh 110

Table 6.5 - Comparison of methods for Al-Sn alloy and stainless steel 112

Table A . l - A36 results, standard Gleeble thermal treatment 124

Table A.2 - A36 results, Gleeble thermal treatment 127

Table A.3 - A36 results, continuous cooling on the Gleeble 127


vi
Table A.4 - A36 results, stepped Gleeble thermal treatment 128

Table A. 5 - D Q S K results, tube furnace thermal treatment 129

Table A.6 - D Q S K results, standard Gleeble thermal treatment 130

Table A.7 - 1080 steel results, furnace thermal treatment 130

vii
LIST OF ILLUSTRATIONS

Figure Page

Figure 1.1- Schematic diagram of a hot strip mill 2

Figure 2.1 - The geometry of two dimensional array of grains 10

Figure 2.2 - The intersection of a sphere and a random plane 21

Figure 2.3 - Schematic illustration of the method of Matsuura and Itoh 28

Figure 4.1 - Design of the Gleeble specimens 34

Figure 4.2 - Specimens used in the tube furnace 35

Figure 4.3 - Schematic diagram of the Gleeble 36

Figure 4.4 - Schematic diagram of vertical tube furnace and specimen holder 39

Figure 4.5 - Typical furnace thermal cycle 41

Figure 4.6 - Standard Gleeble thermal cycles 43

Figure 4.7 - Stepped Gleeble thermal cycles 45

Figure 4.8 - Continuous cooling thermal cycles 47

Figure 5.1 - A36 Steel, Gleeble heated at 5C/sec to 1150 C, 60 seconds 51

Figure 5.2 - A36 Steel, Gleeble heated at 5C/sec to 1000C, lOseconds 52

Figure 5.3 - DQSK Steel, furnace heated to 1050C, 420seconds 54

Figure 5.4 - DQSK Steel, furnace heated to 1150C, 227seconds 55

Figure 5.5 - 1080 Steel, furnace heated to 1100C, immediately quenched 56

Figure 5.6 - A36 Steel, Gleeble heated to 1000C, 120seconds 58

Figure 5.7 - Kinetics of austenite grain growth in A36 steel, heating rate=5C/sec. 66

Figure 5.8 - Kinetics of austenite grain growth in DQSK steel 68

Figure 5.9 - Effect of heating rate on grain growth in A36 steel 71

Figure 5.10 - The statistical model, A36 steel, heating rate=5 C/sec 74

Figure 5.11 - The statistical model, D Q S K steel 75


viii
Figure 5.12 - Calculated and experimental Rn m 76

Figure 5.13 - Application of the statistical model to continuous cooling conditions.79

Figure 5.14 - Evolution of the A36 grain size distribution, normal grain growth.... 82

Figure 5.15 - Evolution of the grain structure, in the A36 steel, abnormal growth. 84

Figure 5.16 - Evolution of the A36 grain size distribution, abnormal grain growth. 85

Figure 5.17 - Austenite grain size in the as-received and solution treated A36 88

Figure 6.1 - Saltikov's and Huang and Form's 3-D grain size distribution 92

Figure 6.2 - The relationship between .syand the ratio of l to V a


m o t 97

Figure 6.3 - Flow chart for applying the method of Matsuura and Itoh 101

Figure 6.4 - Relationship between and dy ld^


m m 104

Figure 6.5 - Evolution of the 3-D grain size distribution 108

Figure 6.6 - Effect of the number of size classes on the method of

Matsuura and Itoh 109

Figure 6.7 - Stainless steel, the estimated 3-D distribution

compared to the measured distribution, from Matsuura et al 113

ix
NOMENCLATURE

A grain section area

A m mean grain area

b Burger's vector

c lower limit of the smallest size class for the geometric scale

C n coefficient for the nth term in Saltikov's equation

d grain diameter

dg initial grain diameter

d A equivalent area diameter

d Ab average equivalent area diameter

d Ag peak equivalent area diameter of the log normal distribution

d Am mean equivalent area diameter

d b average diameter

dg peak diameter of the log normal distribution

D gb grain boundary diffusivity

d m mean diameter

d R relative diameter

d v equivalent volume diameter

d Vb average equivalent volume diameter

d Vg peak equivalent volume diameter the of log normal distribution

d Vm mean equivalent volume diameter

f(d) distribution of grain diameters

fAi fraction of grains measured with equivalent area diameter in size

class i

f ACi corrected fraction of grains measured with area diameter in size

class i
x
fa calculated fraction of grains with equivalent area diameter in size

class i

fpi fraction of grains in size class i, in contact with the edge of the

measurement field

fi fraction of grains of size i

fvi fraction of grains with equivalent volume diameter in size class i

F driving force for grain growth

Fij driving force for growth of grains of size / relative to size j

F P pinning force

/ number of grain faces

Jl number of faces for grains from size class i

k total number of grain size classes

K grain growth constant (theoretical)

K a grain growth constant (power law)

l m mean linear intercept

m grain growth exponent

Mgfy grain boundary mobility

N^i measured number of grain sections of size class i

Nj^a corrected number of grain sections of size class i

Npi number of grains of size class i which touch the edge of the field of

measurement

Ni number of grains in size class i

N tot total number of grains per unit volume

N Vti number of grains per unit volume after truncation correction

P pinning parameter

xi
Pp(d) probability that a grain is in contact with the edge of the field of

measurement

PM^R) probability that the relative diameter is equal to d R

Psid) probability that the section diameter is less than d for a sphere

Q apparent activation energy for grain growth (power law)

r average particle radius

R gas constant

R 2
correlation coefficient

Rg average grain radius

R( radius of a grain of size class i

Rli m limiting grain radius

R max maximum grain radius

s standard deviation of the log normal grain size distribution

s A standard deviation of the log normal equivalent area grain size

distribution

s v standard deviation of log normal equivalent volume grain size

distribution

S 2
sum of the differences squared

t time

T temperature

V grain volume

W(j probability that grains i and j are neighbors

A width of size class for arithmetically scaled distribution

a multiplying factor for geometrically scaled size distribution

7g grain boundary energy

v volume fraction of precipitates

xii
ACKNOWLEDGMENTS

This work is a part of a group project to model the industrial process of hot strip

rolling. In completing this thesis, I have worked closely with other members of the

research group, in order to make progress toward our common goal. The cooperative

atmosphere has been the catalyst for many group discussions regarding experimental

methods and the interpretation of results. I would like to thank all members of the hot

strip mill modelling group for their contributions to this work. In particular, Dr. Matthias

Militzer deserves a special mention for his guidance and support throughout the course of

this research. I am also grateful for the technical assistance provided by Mr. Rudy Cardeno

and Mr. Binh Chau and for the support and guidance of Dr. Bruce Hawbolt. I would like

to acknowledge the financial support provided by the American Iron and Steel Institute

and the United States Department of Energy.

I owe thanks also to my wife, Katrina, for the patience and support she has so

generously shown me, especially during the later stages of this work.

xiii
CHAPTER 1

INTRODUCTION

Steel is a versatile material which has found widespread use because of its

exceptional mechanical and magnetic properties, its relatively low cost, and the ease with

which it can be used in manufacturing processes such as forming, welding and machining.

The steel industry has experienced great successes as a result of these characteristics.

However, in recent times, economic developments have generated a highly competitive

atmosphere in many industries, and steel producers have been forced to do battle with

each other, and with other materials producers for a share of a market which is demanding

higher quality products at lower cost. This has driven steel producers to improve their

operations by the application of improved and advanced technology. An example of this is

process modelling, which links measurable processing parameters to the final properties of

the product. Process models are an excellent knowledge base for off line process

simulation, and ultimately, it is likely that models will be used for on line process control.

The main steps in the production of steel are refining of the material in the molten

state, followed by casting to form a solid product, then deformation of the solid to give

the specified mechanical properties, and a more useful product shape. If the desire is to

produce steel in form of sheet, the first stages of deformation will be performed by hot

rolling.

The process of hot strip rolling is shown schematically in Figure 1.1. A steel slab is

heated to temperature in a reheat furnace. Deformation begins in the roughing mill where

the slab is reduced in thickness by a series of rolling passes. The slab leaves the roughing

null with Utile change in width, but with significantly increased length and reduced

1
2

thickness; at this stage the steel is referred to as transfer bar. The next stage of

deformation is the finishing mill, at the end of which the material is referred to as strip.

Frequently, the aim is to complete all deformation while the steel is austenitic. The

subsequent decomposition of the austenite can be controlled by cooling at an appropriate

rate, using water sprays on the runout table. At the end of the runout table the material is

coiled on the down coiler.

Roughing mill

Reheat furnace

Figure 1.1- Schematic diagram of a hot strip mill.

Most rolling passes are carried out at an elevated temperature, where

recrystallization occurs, so that large amounts of deformation are possible. Other

metallurgical phenomena which accompany hot rolling are grain growth, phase

transformation, and in some steels, precipitation. The microstructure of the finished

product dictates its properties; the control of the metallurgical phenomena can ensure that

the product meets the property requirements specified by the customer.

Process modelling of hot strip rolling is a sensible approach to achieving the goals of

improved product quality and reduced cost by the application of improved technology.

The modelhng task requires the luiking of the process parameters with the kinetics of the

metallurgical phenomena. The kinetics of grain growth assume significant importance in

the evolution of the structure during hot strip rolling.

This thesis examines the kinetics of austenite grain growth, measured

experimentally, under conditions of time and temperature which are relevant to the
3

process of hot strip rolling. Measurements are to be used to develop and validate a model

to describe the kinetics of austenite grain growth during hot strip rolling. The model to be

used will have a fundamental basis, and the phenomenon of grain growth will be treated as

the evolution of a distribution of three dimensional grains. Grain size measurements are

generally performed on planar sections through the structure, and direct measurement of

the three dimensional grain size is a difficult task. However, procedures have been

reported to estimate the three dimensional grain size from two dimensional measurements.

These methods will also be examined in this work.

The grain growth model is one component of the process model being developed by

an engineering research team at the University of British Columbia (UBC); the author of

this work is a member of that team. The U B C team is working in conjunction with United

States Steel and the National Institute of Standards and Technology. The aim of the U B C

group is to develop a versatile and fundamentally based model for the prediction and

control of the microstructure and mechanical properties of steel during hot rolling. The

model will be developed for a range of mill processing conditions and steel compositions,

including plain carbon, interstitial free and microalloyed grades. In this stage of

development of the process model, investigations have been directed toward the plain

carbon steels. The deliverable from this research will be a PC (personal computer) based

model, which will link the processing variables to the final microstructure and properties

of the product. Application of this model will allow the steel producer to deliver a higher

quality product to the customer, at lower cost.


CHAPTER 2

LITERATURE REVIEW

2.1 Modelling of Hot Strip Rolling of Steel

Several workers have developed process models for the prediction and control of

microstructure and mechanical properties during hot strip rolling of steel [1-5]. Process

models which have been developed are generally based on empirical relationships which

relate processing parameters to material structure and properties. A more desirable model

would have a fundamental basis. Such fundamental models will be capable of accurately

describing the thermal history of the strip, the rolling forces at each rolling pass, and the

kinetics of the microstructural phenomena of recrystallisation, grain growth, precipitation

and phase transformation, which accompany hot strip rolling. The kinetics of the

microstructural events are modelled by investigating each event separately, using

laboratory equipment which is able to approximate the thermal and mechanical conditions

in the mill. The individual events are combined using a model which describes the thermal

and mechanical history of the strip. Since the microstructural evolution can be quantified

and related to the stresses of deformation, the mechanical properties at each stage can be

anticipated; consequently, the mill forces can be predicted based on the properties. The

resulting model must then be validated by performing measurements on the mill under

industrial conditions to verify temperature and force predictions. The structure must also

be assessed in the hot rolled product, to verify microstructural predictions. This approach

is being taken by the U B C team. Here, one of the metallurgical events to be modelled is

examined; that is, the kinetics of austenite grain growth.

4
2.2 Modelling of Grain Growth

2.2.1 The Power Law for Grain Growth

Burke and Turnbull were among the first to model the kinetics of grain growth [6].

They assumed that the driving force for growth is the reduction in grain boundary area and

the subsequent release of boundary surface energy, 7^. The model was proposed for the

case of normal grain growth in a pure homogeneous material and the assumption was

made that the only forces acting on boundaries are those due to surface curvature. It was

also assumed that the average radius of curvature of a boundary is proportional to the

grain diameter, d, and that the rate of change of the diameter is proportional to the driving

force for growth, F, as described by

^ = k.F
dt 2.1

where k is a constant of proportionality.

Since the driving force for growth is assumed to be directly related to boundary area

per unit volume, it can be shown that F is proportional to the inverse of grain diameter, d.

By integration of equation 2.1, Burke and Turnbull were able to derive the parabolic

equation for the kinetics of isothermal grain growth,

d -dQ=Ktexp(-Q/RT)
2
2 .2

where dg is the grain size at time equal to zero, t is time, AT is the growth constant, Q is the

activation energy for grain growth, R is the gas constant and T is absolute temperature.

The grain boundary energy, y , is included in the constant, K, and the exponential term
gb

accounts for the temperature dependence of the rate of boundary migration.

Feltham in 1957 [7] and Hillert in 1965 [8] derived equations of the form of

equation 2.2 to describe the kinetics of grain growth. Feltham focused on the relationship
6

between driving force and boundary mobility, whereas Hillert concentrated on the effect of

a distribution of grain sizes.

The parabolic growth law has been found to apply to the increase in mean cell size

of a froth of soap bubbles [9]. This implies that the analysis is geometrically correct, since

in a network of soap bubbles, as in a network of grains, the driving force for growth is the

decrease in boundary surface area [6]. Monte-Carlo simulations of growth have also

yielded growth kinetics which approach those predicted by the parabolic law [10].

Unfortunately, experimental results often do not obey the parabolic equation. For

this reason, the original equation [6] is often modified and used in the form,

d -d(? = K t.exp(-Q/RT)
m
0 2.3

where m is the grain growth exponent and K is the fitting constant. The equation is
0

referred to as the power law for grain growth.

A short summary of experimental results from very pure zone refined materials,

which have been described using the power law, was compiled by Anderson [10] and is

shown in Table 2.1. The table demonstrates that m is rarely equal to two, even for

extremely pure metals. It is likely that this is due to the influence of imperfections such as

solute atoms and voids which limit boundary mobility, or have a boundary pinning effect

[6,7,10,16].
7

Metal, Exponent, m Reference

zone refined

Al 4.0 [11]

Fe 2.5 [12]

Pb 2.5 [13]

Pb 2.4 [14]

Sn 2.3 [14]

Sn 2.0 [15]

Table 2.1 - Grain growth exponents for the power law, for zone refined metals [10].

To explain the deviation from parabolic growth kinetics, it is necessary to examine

the assumption that the only forces which influence boundary motion are due to surface

curvature. In Burke and Turnbull's derivation [6], this assumption leads to the

proportional relationship between driving force and the inverse of grain size, the

integration of which leads to the exponent of two. Atkinson [16] suggests that the

mechanisms which influence the value of m include the effect of void size and distribution,

the presence of solute atoms, solute segregation, second phase particle size and

distribution, and texture effects. A study of these effects would be difficult to perform, due

to the small size of the voids and particles, and the low concentrations of solute which are

thought to influence grain growth kinetics.

The power law has often been used to describe the kinetics of austenite grain

growth in process models for the hot rolling of steel [1,3-5]. The kinetics of austenite

grain growth are usually investigated by reheating and holding specimens of steel at a

constant temperature for varying times. A more elaborate cycle involving deformation as
8

well as heat treatment may be used to replicate grain growth during hot rolling. The

specimens are quenched and etched after the discrete holding time to reveal the grain

structure and allow the austenite grain size to be measured. A table of data relating grain

size to time at temperature is produced and models are developed by fitting the results to

the power law. The parameters which can be used for fitting are m, K and Q. Table 2.2 is
0

an overview of literature results reported for plain carbon steels.

C,wt% Mn,wt% m K ,|im /sec


0
m
Q ,kJ7mol Reference

0.22 0.9 2 1.44xl0 12


266 [17]

0.77 7.5 4.2X10 27


400 [3]

0.69 0.76 6 8.2xl0 24


400 [18]

0.08 to 0.30 to 10 3.87X10 32


7M100C 400 [5]

0.89 1.68 1.31x10 s2


7/<1100C 914

Table 2.2 - Power law parameters reported for austenite grain growth.

The use of the power law for ml is empirical. Therefore, the resulting equations

should only be used over the range of experimental conditions for which the model has

been validated. Since it is impossible to carry out experiments for all steels and all

conditions encountered in a hot strip mill, it is desirable to develop a more versatile model

which is based on fundamental principles, which can be validated with limited data.

2.2.2 Statistical Modelling of Grain Growth

Several workers have reported the development of fundamentally based statistical

grain growth models [19-23]. The factors which are known to influence grain growth

kinetics are the driving force for grain growth, the mobility of the grain boundary, and the
9

distribution of grain sizes in the structure. A statistical model is able to account for each of

these.

To understand the influence of the width and shape of the grain size distribution, it is

important to understand the geometry of a network of grains. A number of investigations

have been directed towards this task [24-27]. A grain boundary has a surface energy and

the arrangement of the boundaries is such that the boundary area tends to be minimised

[6]. It is helpful to visualize this effect as being equivalent to surface tension. Atkinson

[16] has summarized the geometry of the structure and the trends which are a

consequence of the driving force for growth, as illustrated for a two dimensional structure

in Figure 2.1. Generally, in such an array, only three grain boundaries meet at a corner,

with the angle between them being 120. In a three dimensional array, three grain

boundaries meet at a grain edge at angles of 120. At a corner, four edges meet at

10928'. A polycrystalline metal is made up of grains which vary in size and shape. To

satisfy the geometrical considerations, grain boundaries must be curved surfaces. The

smallest grains will have boundaries with the lowest radius of curvature, with their centre

of curvature inside the grain. Larger grains will have boundaries with a higher radius of

curvature, with their centre of curvature outside the grain. The motion of a grain boundary

will reduce grain boundary area if the boundary is curved; the boundary moves toward the

centre of curvature. Since the driving force for growth is the liberation of energy by the

reduction of boundary area, grains which are large relative to the grains around them tend

to grow and grains which are small tend to shrink. Therefore, an important factor to

consider when modelling the kinetics of grain growth is the width and shape of the grain

size distribution.
10

Figure 2.1 - The geometry of two dimensional array of grains.

Abbruzzese and Liicke [20,21] have demonstrated the use of a statistical model

which has been developed from the work of Hillert [8]. Abbruzzese and Liicke were able

to use the statistical model to describe grain growth under the influence of texture and in

the presence of second phase particles. For austenite grain growth in low carbon steels on

the hot strip mill, the pinning effect due to second phase particles is of primary

importance.

The statistical model for grain growth which accounts for the pinning effect of

second phase particles is based on the following assumptions:

The driving force for growth, F, is the liberation of grain boundary energy, y .
gb

Grains of size / can be described using which is the radius of a sphere of the same

volume as the grain. The relationship between the grain boundary surface area and the

radius can also be approximated by assuming the grain shape to be spherical.


11

The grain size distribution can be approximated by subdividing the distribution into

discrete grain size classes.

If a grain from size class i is larger than its neighbor from size class j, a driving force for

growth will exist, which can be calculated using

F
ij = Vgb
yj
R R
t )

Fjj = 0 if Fy < 0 for R > Rjt

F
ij F
ji 2.4

where P is related to the boundary pinning force.

The growth rate for grains in size class i can be determined by assuming grains from

class i are surrounded statistically by grains from all other classes. Growth rate is

calculated using

dR
i
^Mg^WijFij
2.5

where is the probability that a grain from class i is in contact with a grain from class j

and M gb is the grain boundary mobility. Grain boundary mobility is a function of the grain

boundary diffusivity, the magnitude of the burgers vector, and the temperature.
12

The probability that a grain from class i is in contact with a grain from class j can be

calculated assuming that spherical grains are randomly distributed, using

fvj j
R

wa =

2.6

where/yy is the fraction of grains of size class j.

The fitting parameter for the model is P. The model is used by performing a

calculation for each time step to determine the number of grains entering or leaving each

size class. Smaller grains tend to srmnk and disappear while larger grains tend to grow, so

that the mean grain size increases with time.

The model treats grains as three dimensional objects. The unique growth kinetics for

each grain size class are determined by considering the influence of grains from all

surrounding size classes. Therefore, the width and shape of the three dimensional grain

size distribution is important. To validate the statistical model, it is necessary to estimate

the three dimensional grain size distribution. This will be described in a later section.

2.3 Normal and Abnormal Grain Growth

Grain growth models are generally used to describe normal grain growth, which is

the continuous increase in mean grain size with time. Atkinson [16] describes normal grain

growth as scaling, since the grain structure can be made to appear identical at any point in

time by changing the magnification of the image. During normal grain growth, the width

and shape of the grain size distribution do not change significantly with time.

Abnormal grain growth can also occur, when discontinuous growth kinetics are

observed. During abnormal growth, the largest grains in the structure tend to grow much

faster than the grains around them, and the grain size distribution broadens. The growth
13

behaviour is unstable until the larger grains consume almost all of the smaller grains and a

structure consisting of uniformly large grains has been established.

It is generally accepted that abnormal grain growth is due to the pinning effects of

second phase particles, as discussed in recent review articles by Gladman [28], and

Worner and Hazzledine [29]. It is generally accepted that much of the progress in this field

has resulted from the work of Zener, as quoted in a classic paper on microstructure by

Smith [30]. As a result, the effect of particle pinning has become known as Zener pinning.

A moving boundary can be pinned by a second phase particle or pore in its path.

When a distribution of particles are present, a net pmning force will exist, the magnitude

of which, according to Zener, can be calculated from

j-,
3 v
v
Ar 2.7

where r is the particle radius, v is the second phase particle volume fraction and ygb is the

grain boundary energy [30]. Grain growth in the presence of particles will be affected if

the magnitude of the pinning force is significant compared to that of the driving force for

growth. The growth rate can approach zero as the grain radius tends toward a linnting

value, Rn . Zener proposed that Rn can be calculated from the pinning force. Since the
m m

pinning force is related to the pinning parameter, P, used in the statistical model for grain

growth [20], it is possible to write

_J_
Rlim ~ p 2.8

A grain size distribution which has a mean grain size approaching the lairing value

can become stagnant. However over time, second phase particles tend to either coarsen or
14

dissolve, depending on the equilibrium solubility of the second phase. As this happens, the

size and distribution of the second phase particles will change. Since the pinning force is

dependent on the size of the particles, the j^iting grain radius, Rn , will also change with
m

time. The grains which have the highest driving force for growth are those which are the

largest relative to the grains around them. They are the first to break free from the piruiing

particles and they consume the matrix of smaller grains which remain pinned. Gladman

[31,32] proposed the expression

6RQV (3 r RQ 1
-2
K
l 2 -Rmax -)

for the relationship between the size of the pinning particles, r, the volume fraction of the

second phase, v, the average grain radius, R and the radius of the largest grains present,
0

By predicting the coarsening behavior of second phase particles, and choosing an

appropriate ratio of RQ to R ,
MAX Gladman was able to predict the time and temperature for

the onset of abnormal grain growth for a number of steels. In plain carbon steels, Gladman

showed that the likely identity of the second phase particles which cause abnormal grain

growth is aluminum nitride. This result has been reported also by other workers [33,34].

2.4 Measurement of Grain Size

A grain is a three dimensional object with a size which can be characterized by its

volume. However, for the sake of simplicity, grain size tends to be described using units of

length. Most measurements of grain size are obtained from a planar section through the

three dimensional structure. The measured value, which is a two dimensional result, is not

equal to the true three dimensional grain size. In this work, grain size measurements are to
15

be used to validate a statistical model for grain growth. It is the three dimensional grain

size distribution that is of interest since the statistical model is fundamentally based.

The three dimensional grain size can be described by the equivalent volume

diameter, d , which is the diameter of a sphere with the same volume as the grain [35], as
v

defined by

2.10

where Vis the grain volume.

The three dimensional grain size distribution, f(d ), can be obtained by measuring
v

the volume of a significant number of grains. However, grain volume in an opaque

material is not easy to measure. Individual grains would have to be observed by complete

disintegration of the specimen, or by using an x-ray technique, or by serial sectioning. For

most investigations of grain growth kinetics, these procedures are impractical. As a result,

the three dimensional grain size distribution, f(d ), is often estimated from the two
v

dimensional measurement [36-39].

Measurement of the two dimensional grain size can be made by determining the area

of each grain visible in the planar section through the structure. The two dimensional grain

size can be expressed as the equivalent area diameter, d , which is defined by


A

2.11

where A is grain area [35].

By determining d A for a large number of grains, it is possible to construct a two

dimensional grain size distribution,/^), which can be used directly to estimate the three
16

dimensional distribution [36,37,39]. The two dimensional arithmetic mean equivalent area

diameter, d , is defined as the sum of all grain diameters divided by the number of grains.
Ab

Similarly, a three dimensional mean diameter, d , can be defined.


Vb

The mean grain area, A , is defined as the sum of the area of all grains, divided by
m

the total number of grains. By the substitution of A m for A in equation 2.10, the geometric

mean equivalent area diameter, d , Am can be calculated. The mean area can be determined

by using quantitative image analysis equipment to measure the area of each grain, as

described by A S T M E1382 [40].

Mean grain area can also be determined using Jeffries' method, as described by

A S T M E l 12 [41]. A simple count of the number of grains is all that is required, and the

mean area is calculated by dividing the image area by the number of grains. The major

difference between the measurement of mean area by image analysis and by Jeffries'

method is the treatment of grains which touch the edge of the field of view. Using the

image analyser, only whole grains are measured. Therefore grains which touch the edge of

the frame are ignored. Using Jeffries' method, each grain which touches the edge of the

frame is counted as a half grain.

A n alternative two dimensional grain size measurement is Heyn's linear intercept

method. Again, measurements can be made using image analysis equipment, as described

in A S T M E1382 [40] or by manual techniques, as described by A S T M E l 12 [41]. Both

methods require that lines of known length are superimposed on the planar section of

grain boundaries, and each intersection of a line and a boundary is counted as one

intercept. The mean linear intercept length, l , is the total line length divided by the
m

number of intercepts. A number of methods have been proposed which use the linear

intercept diameter for the estimation of the three dimensional grain size distribution

[35,38].
17

Two dimensional area or intercept measurements of grain size are performed on a

random planar section through a three dimensional structure. The number of grains which

are measured must be significant so that errors due to the randomness of the section are

avoided. A S T M E1382 [40] and E l 12 [41] suggest that a minimum of five separate fields

are measured, and that at least fifty grains are visible in each field [40,41].

The methods previously described are satisfactory for cases where the grain

structure is relatively uniform. For structures with a grain size distribution which is

bimodal (composed of fine and coarse grains), A S T M E l 181 contains some additional

suggestions [42]. For structures where the fine and coarse grains are distinctly separate, a

mean grain size can be determined in each region by measuring the grain size in each

region using standard techniques already described. The area fraction of each region can

be determined by image analysis measurement, or estimation using comparison techniques.

The overall mean grain size can be calculated by combining the mean and area fraction of

each region.

2.5 T h e G r a i n Size Distribution

It is generally thought that during normal grain growth, the three dimensional grain

size distribution is approximately log normal [7,22,23,38,43]. A log normal distribution

can be described by the peak grain size, d, and the standard deviation, s, [23,38] using

2.12
18

The average grain size, d , is given by


b

dfr = 2Zfid t = d exp{{lns) /2J


g
2

i=l 2.1.3

For a three dimensional distribution, Equation 2.12 and 2.13 are modified by the

addition of the subscript, V. For example, d becomes d . The mean equivalent volume
v

diameter, d , is given by
Vm

2.14

The two dimensional grain size distribution which exists during normal grain growth

can also be approximated by a log normal distribution, as demonstrated in the hterature

[7,22,23,38]. For a two dimensional distribution, the subscript, A, would be used in

equations 2.12 and 2.13. Other distributions have been proposed [16,35] which describe

experimental data with a limited improvement in accuracy, but the log normal distribution

is characteristically simple and convenient.


19

Parameter Arithmetic Geometric

Number of grains Nt Nj,


Fraction of grains fi fi
Lower limit of smallest size 0 c

class, i=l

Upper class limit LA c.a 1

Lower class limit (i-l).A cat' 1

Class midpoint (i-l/2).A c.a -


1 1/2

Class width A q.(l-cW)

Upper limit of largest size k.A c.a k

class, i=k

Table 2.3 - Description of geometric and arithmetic size class scales

A grain size distribution can be approximated by subdivision of the distribution into

discrete size classes. The class limits are usually arranged using an arithmetic or a

geometric scale, as described by Table 2.3 [35]. The arithmetic scale is described by the

class width, A, and the number of classes, k. The geometric scale is described by the lower

limit of the smallest size class, c, the geometric multiplying factor, a, and the number of

classes, k.

2.6 Estimation of the Three Dimensional Grain Size Distribution

A number of mathematical conversions have been proposed for estimating the three

dimensional grain size distribution from two dimensional measurements. Some methods

are based solely on the measurement of the two dimensional linear intercept length
20

[44,45], while others are based on the measurements of area [36,37,39], or on the

combined measurement of intercept and area [38].

2.6.1 Methods Based on Spherical Grain Shape

Saltikov [36] and Huang and Form [37,46] assumed the grain shape to be spherical.

Saltikov proposed a method which requires the measurement of the two dimensional grain

size distribution, f{dA). The method is used to apply a correction to the number of grains

observed in each size class. Huang and Form proposed a method which is similar to the

method of Saltikov, but with an additional correction applied.

Use of Saltikov's method requires that the following assumptions are made. Firstly,

grains are assumed to be spherical in shape. Secondly, the measured distribution can be

represented by dividing the data into a number of discrete size classes. And finally, the

three dimensional grain size for each size class is not a range of grain sizes, but it is the

upper limit of that size class. That is, the number of grains in the size class in the two

dimensional distribution is equal to the number of grains which have a diameter which is

between the upper and lower hmit of the size class. But, in the calculated three

dimensional distribution, all grains in a size class are assumed to have a diameter equal to

the upper limit of that class.

The method is based on the probable diameters which result from intersection of a

plane and a sphere, as illustrated by Figure 2.2. The shape of the intersection between a

plane and a sphere of diameter, d , is a circle. The diameter of the circle, d , will be less
v A

than or equal to dy. By definition, d is the equivalent volume diameter, while d is the
v A

equivalent area diameter.


Figure 2 . 2 - The intersection of a sphere and a random plane
22

It can be easily shown that the probability, Ps(d ), that the measured diameter is
A

between zero and d , is given by [35]


A

2.15

Hence the probability that the measured diameter is between d Ai and d ^ will be the
A

difference between P^{d \) and P${d ^).


A A

For a distribution of spherical grains of different sizes, the problem becomes more

complex. A grain section with measured diameter, d , must be from a grain which has a
A

true diameter, d , that is greater than or equal to the measured one. Considering the
v

largest size class, the true number of grains in this class will be greater than the observed

number, since some of the grain section diameters measured in lower size classes are due

to truncation of grains from the largest size class. The number of grains per unit volume

which are estimated to be in the largest class, N , is determined by applying a probability


Vtk

correction to the number of grains observed, N^, as described by

1-Hdvk-l) 2.16

where dy^ is the upper limit of the size class for the measured data [36].

Only one correction is applied to the largest size class, since it is assumed that there

is no chance that grain sections observed in that size class actually belong to spherical

grains from a larger size class. For all size classes smaller that the largest one, other

corrections must be applied. The number of grains observed in classes smaller than the

class of interest, which actually belong in the class of interest, is accounted for. This

increases the number of grains in the size class. Then, the number of grains observed in the
23

class, which belong in size classes larger than the class of interest, is determined using

equation 2.15. This number is subtracted from the result. A general equation can be

written,

2.17

where C is the coefficient for the nth term and k is the number of size classes.
n

The work of Saltikov is not unique in its approach to estimating the three

dimensional distribution, as others have used a similar theory [47,48]. Methods were

developed in the earlier part of this century at a time when advanced means of calculation

were not available, and tables of coefficients were generally reported. Each method can

only be applied to a distribution described by a specific arrangement of size classes. For

example, Saltikov provided coefficients for twelve size classes which are geometrically

scaled with a multiplication factor, a, equal to 1 0 0 1


[36]. Scheil provided coefficients for

fifteen arithmetically scaled classes [47].

Saltikov's method applies a correction only for the probability that the planar section

reveals a diameter which is less than the diameter of the grain; this is called truncation

[35]. A second correction should also be applied to account for the fact that it is more

likely that a large grain will be intersected by a planar section than a small grain; this a

correction for sampling [35]. Huang and Form [37,46] proposed a method which is

identical to Saltikov's, except that it accounts for sampling as well as truncation. If a cube

of unit volume is considered, the correction for the sampling is estimated by dividing the

number of grains after truncation correction by the size class diameter.

2.18
24

The main problem reported for methods based on the assumption of spherical grain

shape is the prediction of a negative number of grains for the smaller size classes. Huang

and Form reported such a result [46]. Aaron et al. suggested that negative counts can

occur if some of the smallest grains are not detected when the measurements are being

performed [49]. The smallest classes are most affected, since errors are compounded by

the nature of the corrections applied. The biggest source of error is likely to be the

assumption that the grain shape is spherical, since an array of spherical grains cannot fill

space [49]. A number of methods have been proposed which make use of more realistic

grain shapes [38,39,45,50].

2.6.2 The Method of Takayama et aL

A method has been proposed by Takayama et al. to estimate the three dimensional

grain size distribution from the measured mean area , A , and the mean linear intercept
m

length, l , [38,51]. This method assumes that all grains are tetrakaidecahedral in shape,
m

and therefore the grains are space filling. In addition, the grain size distribution is assumed

to be log normal and grains of different size are distributed randomly in space.

The first step taken by Takayama et al. [38,51] in the development of the method

was the determination of the probability distribution of the linear intercept length. The

intersection of a random line with a solid object of volume, V, will produce an intercept of

length, /. From V, the equivalent volume diameter, d , can be easily calculated. If a


v

number of random lines are constructed, a probability distribution which relates / to dy can

be determined. Takayama et al. performed a computer simulation using this approach to

determine the probability distribution for the tetrakaidecahedral grain shape and for a log

normal distribution of grain sizes [52]. The mean linear intercept length was determined to

be a function of the width of the grain size distribution. A n equation was fit to the
25

simulation results relating the standard deviation of the log normal distribution, s , and the v

peak volumetric grain size, d , to the mean linear intercept, l . The mean area, A , and
Vg m m

mean linear intercept, l , were related to the arithmetic mean volume diameter, d , using
m vt)

the relationships described by DeHoff [53].

By applying the characteristic equation of a log normal distribution (Equation 2.12),

it was possible to relate the mean area, A , and the mean linear intercept, l , directly to
m m

the peak grain size, d , and the standard deviation, s , as described by


Vg v

= 0.6066W ^exp(5(lnsv) / 2 ) 2

"m y
2.19

A = 0.48610 i ^ exp(4(ln^)
m L y
2

2.20

These equations can be rearranged and combined with equation 2.14 to give

(
s =exp 21rJ 1.14935
v

2.21

2^
0.82575A2.5 3(hu ) v

-exp
li
2 . 2 2

which can be applied to the measured results, to estimate the three dimensional grain size

distribution.

Jeffries' method for measuring mean area, A , and Heyn's method for measuring
m

mean linear intercept, l , are used [41]. These parameters are easily determined with or
m

without image analysis equipment, making it possible to obtain a quick estimate of the

three dimensional grain size distribution. However, it should be emphasised that this
26

method is limited by the assumption that the grain size distribution is log normal, and that

all grains are tetrakaidecahedral in shape. In fact, a range of grain shapes have been

observed [24-27,43,54].

2.6.3 The Method of Matsuura and Itoh

An alternative method for estimating the three dimensional grain size distribution,

based on a range of grain shapes, has been proposed by Matsuura and Itoh [25,39]. They

used twelve different types of regular, equiaxed polyhedra to represent the range of grain

shapes observed in a polycrystaUine solid [39]. By examining the work of Rhines and

Patterson [43], they were able to determine a simple relationship between the number of

grain faces, / , the equivalent volume diameter for the grain, dy, and the mean equivalent

volume diameter for the entire distribution, dy , as described by


b

2.23

Matsuura and Itoh [39] used a similar approach to that taken by Takayama et al.

[52] to determine a probability distribution for the equivalent area diameter. The

intersection of a random plane with a solid object having volume, V, will result in a

polygon with area, A, and appropriate values of dy and d can be easily calculated. If a
A

large number of random planes are generated, a probability distribution for the equivalent

area diameter can be constructed. Matsuura and Itoh chose to describe the probability

distribution in terms of a relative diameter, d /d , which can also be denoted as d . They


A v R

performed computer simulations using the twelve grain shapes to determine the probability

distribution, P {d ), as a function of the number of grain faces. The relative diameter


M R

distribution curves were characterised by the position and magnitude of the peak
27

probability and the maxirnum value of the relative diameter, d . These parameters were
RM

used to develop general equations to describe the shape of the curves.

The method is applied by measuring the two dimensional grain size distribution,

f{d ), using quantitative image analysis. The measured distribution is then divided into a
A

number of discrete size classes. Each size class, /, is made up of grains having a measured

diameter between d Ai and d ^.iy A l l grains in size class i for the three dimensional
A

distribution are represented by the upper limit of the size class, d , which is equal to d .
Vi Ai

It is assumed that a three dimensional size class contains grains of only one shape, and the

number of faces, J , can be calculated using equation 2.23, where d


t Ab is used as a first

approximation for d . There is a finite probability that the relative diameter is equal to
vb

any value between zero and d RMi . The measured fraction of grains in each size class, f i, is
A

assumed to be equal to the sum of the probability distributions for each size class, / Q . The

method is illustrated for three size classes in Figure 2.3 [39]. In the figure, the hatched

region under each curve is representative of the integration of the probability function

from each size class, over the range of diameters in the second size class. The results from

each integration can be multiplied by the true three dimensional fraction of each size class

(fvbfv2 and/yj), then summed, to determine the fraction of grains observed in the two

dimensional distribution, in the second size class, f 2- In reality, it is the two dimensional
A

distribution that is most easily measured, and the true three dimensional fractions are

initially not known.


28

Figure 2.3 - Schematic illustration of the method of Matsuura and Itoh


[39] for the case of three grain size classes.
29

The summation of the integrations is described again by the equation,

2.24

where k is the number of size classes, and fyj is the true fraction of grains in each size

class. The fraction of grains which should be observed in the two dimensional distribution

in size class i will be equal to f . Ci Note that the last term in the equation accounts for the

probability that a grain belonging to size class (/-l) has produced a section area with an

equivalent area diameter that belongs in size class i. This is illustrated in Figure 2.3, where

the tail of the probability function from size class 1 is integrated from 1.0 to d j, to RM

account for the section diameters which will be observed in the two dimensional

distribution, in size class 2.

The three dimensional grain size distribution is known when values of f Vi are

determined for size classes from i=l to i=k. Equation 2.24 is used to create a matrix of k

equations with k unknowns. A solution is determined for/y; to fy^ by nrinmising the sum

of the squares, S , where


2

2.25

Values of f Vi which are determined by solving the matrix of equations are used to

calculate a revised mean volumetric grain size, dy . New values of Jj are determined using
b

equation 2.23 and probability functions are also redefined. The new functions are used in
30

equation 2.24 to solve again for f . The calculations are repeated until the values of f
Vi Vi

are seen to converge.

The method of Matsuura and Itoh does not rely on unrealistic assumptions regarding

the type of grain size distribution or the grain shape. To apply the method, the user needs

to create a routine using computer code, making the method less convenient than the

method of Takayama et al. [38,51].


CHAPTER 3

SCOPE AND OBJECTIVES

The goal of this research is the measurement of austenite grain growth kinetics in

plain, low carbon steels. Relevant time and temperature conditions are selected to reflect

the industrial process of hot strip rolling. Area and linear intercept grain size

measurements are made and used for the validation of a statistical model for grain growth.

This study has two objectives. Firstly, austenite grain size will be measured in two

commercial plain carbon steels. Experiments will be performed over a range of

temperature and time conditions which are relevant to hot strip rolling. The measured

results will be used to quantify the kinetics of austenite grain growth.

Secondly, an appropriate method for the estimation of the three dimensional grain

size distribution will be determined. The methods proposed by Saltikov [36], Huang and

Form [37], Takayama et al. [38] and Matsuura and Itoh [39] will be compared. The most

suitable methods will be used to estimate the three dimensional grain size distribution, to

be applied to the statistical model for grain growth described by Abbruzzese and Liicke

[20,21].

31
CHAPTER 4

EXPERIMENTAL

4.1 Sample Composition and Specimen Design

Samples of two plain carbon steels, denoted as A36 and DQSK, were obtained from

United States Steel (US Steel) Gary Works. Specimens were cut from the samples, to be

thermally cycled by heating in a tube furnace or by resistance heating in a

thermomechanical simulator. Specimens from a third steel, denoted as 1080, were

thermally cycled in a previous study [55]. Measurements from the 1080 steel were used

for comparing the methods for estimating the three dimensional grain size distribution.

The chemical compositions of the three steels used in this study, in weight percent,

are reported in Table 4.1. The A36 and DQSK steels differ primarily in carbon content and

manganese content, with the A36 being more highly alloyed. A low alloy content makes

the austenitic microstructure difficult to reveal, affecting the specimen shape and the

design of experiments. The 1080 steel was chosen for its higher alloy content since the

austenitic microstructure in this material can be more easily revealed. A l l steels are

aluminum killed. The chemical analysis of the 1080 steel is included in Table 4.1.

Steel C Mn P S Si Cu Ni Cr Al N

A36 0.17 0.74 .009 .008 .012 .016 .010 .019 .040 .0047

DQSK 0.038 0.30 .010 .008 .009 .015 .025 .033 .040 .0052

1080 0.78 0.68 .024 .014 0.25 .035

Table 4.1 - Composition of steels to be used in this investigation.

32
33

Inspection of the Fe-Fe3C phase diagram shows that for these carbon levels, the

equilibrium microstructure obtained below 727C is a combination of ferrite and iron

carbide [56], these products resulting from the decomposition of high temperature

austenite. In order to reveal the location of the prior austenite boundaries, it is necessary

to quench at a high rate so that the decomposition of austenite does not produce a

structure which is dominated by polygonal ferrite. During the transformation, ferrite

nucleates at the austenite grain boundaries. If the amount of ferrite can be limited to

between 5% and 15%, it will outline the prior austenite boundaries. Alternatively, if a

higher quench rate is achieved, the diffusional transformation of austenite can be averted

and a martensitic structure will result. Certain etchants have been used to reveal the prior

austenite boundaries in the martensitic transformation product.

Samples of the A36 and DQSK steels were received in the form of transfer bar,

taken at the end of rough rolling on a hot strip mill. While many specimens were thermally

cycled in the as-received condition, some of the A36 specimens were solution treated in an

air furnace for 3 hours at 1200C, prior to being tested. These specimens were sealed in

quartz tubes under a partial vacuum to avoid excessive oxidation and decarburization. The

specimen temperature was not monitored directly, but furnace temperature was

maintained within 10C of the set point. After holding for 3 hours, specimens were

quenched by breaking the quartz tubes in a bath of room temperature water.

The size and shape of specimens machined from the transfer bar were determined by

the equipment to be used to apply the thermal cycle. The A36 tests and a few D Q S K tests

were performed on a Gleeble 1500 thermomechanical simulator using tubular specimens

so that a quenching fluid could be passed down the axis after the thermal cycle. The

Gleeble designation is a registered trade name for the thermomechanical simulator,

marketed by Dynamic Systems Inc., Troy, N Y . The minimum tubular wall thickness was

limited to the diameter of ten austenite grains, in order to rninimize the effects of the free
34

surface on the grain growth kinetics [6]. The maximum wall thickness was timited by the

quench rate required to reveal the structure. Two tubular geometries were chosen, each

20mm in length with an inner diameter of 6mm. The wall thickness was either 1mm or

2mm. The design of a Gleeble specimen is shown in Figure 4.1.

Figure 4.1 - Design of the Gleeble specimens.

A vertical tube furnace was also used, to perform many of the tests for the D Q S K

steel. The specimen thickness was again limited to a rmnimum of ten grain diameters and a

maximum determined by the required quench rate. Specimens were rectangular in shape,

15mm wide, 25mm high and either 1.5mm or 3mm thick. A 2mm diameter hole was

drilled through the thickness at the top of each specimen so that it could be hung in the

furnace. Figure 4.2 illustrates the specimen design for the vertical tube furnace.
35

15 mm

E

LO

Figure 4.2 - Specimens used in the tube furnace

4.2 Equipment for the Thermal Treatment

4.2.1 The Gleeble 1500 Thermomechanical Simulator

Tubular specimens were thermally cycled on the Gleeble by resistance heating to

temperatures in the range of 900C to 1200C, using soak times extending to 1200

seconds. It was possible to impose carefully controlled thermal cycles using computer

controlled resistance heating capabilities of the Gleeble. During heating, the specimens

were held in a vacuum of less than 10~ Torr to minimize decarburization and oxidation.
4

Figure 4.3 is a schematic diagram of the specimen support in the Gleeble test

chamber. Each thermal cycle, including the quench, was programmed into the Gleeble

control system. Specimens were prepared by cleaning and spot welding the thermocouple

wires. Wires were welded at the specimen mid-length with each wire aligned with the tube

axis and in contact electrically, only on the specimen surface.


36

Figure 4.3 - Schematic diagram of the specimen support in the Gleeble test chamber.
37

During the thermal cycle, the temperature was controlled by a feedback system

operating at a sampling rate of 100Hz. Temperature was monitored using the output of

the Pt/Pt-10%Rh, Type S, thermocouple intrinsically spot welded to the specimen surface.

The thermocouple wires were 0.25mm in diameter (0.010").

Heating rates of 5C/second, 50C/second, 100C/second and 300C/second were

employed to heat specimens to a holding temperature. Control of temperature during

heating was excellent for the lower heating rates of 5C/second and 50C/second.

However, reproducible temperature overshoot was observed for the higher heating rates,

since a time delay exists between the application of power and the resultant temperature

change. The magnitude of the temperature overshoot was dependent on the heating rate,

holding temperature and the specimen design.

Temperature control during the isothermal holding period was excellent, always

being within 2C of the desired temperature.

Most specimens were quenched by passing hehum down the tube axis. Higher

quench rates of up to 250C/second were achieved by also directing hehum onto the outer

specimen surface. Water was used as an alternative to hehum when the maximum quench

rate was required; rates as high as 600C/second were achieved. However water

quenching was not a favorable option since water vapor contaminated the Gleeble vacuum

system.

4.2.2 The Vertical Tube Furnace

The DQSK steel required an extremely high quench rate to retain the outline of the

prior austenite microstructure. The vertical tube furnace was used for most of the D Q S K

tests to avoid the problems associated with water quenching in the Gleeble, and to achieve

the highest possible quench rates (approaching 1500C/second). The heating rate and

thermal history of specimens heated in the tube furnace was dictated by the laws of heat
38

transfer. During the thermal cycle, oxidation was minimized by directing argon past the

specimen at a constant flow rate of 3 litres/minute. Figure 4.4 is a schematic diagram of

the furnace arrangement.

The furnace consists of a vertical alumina tube, 900mm in length, with an inner

diameter of 38 mm and an outer diameter of 44 mm. The tube is electrically heated using

a Super Kanthal element. Temperature is controlled by a proportional digital system with

thermocouple feed back, the position of the control thermocouple being outside of the

alumina tube. The furnace was set to 22C above the desired specimen temperature, this

offset being experimentally determined. The furnace temperature was allowed to stabilize

for a period of at least 8 hours at each temperature before the tests could be performed.

The furnace rheostat was set so that the controller was on a heating cycle and on a cooling

cycle for equal periods of time, the control limits being within 5C.

The specimen was hung in the furnace from a chromel wire 1.2mm thick. The wire

was also used to support a Type S thermocouple, electrically insulated by a twin bore

alumina sheath. The chromel wire and thermocouple arrangement were housed in an open

ended quartz tube with an inner diameter of 12mm and an outer diameter of 14mm. The

quartz tube allowed for the motion of the apparatus within the alumina furnace tube

during quenching. Above the furnace, a support arrangement was used to locate the

specimen at the same height in the furnace for every test.


39

Thermocouple, type S

Figure 4.4 - Schematic diagram of the vertical tube furnace and specimen holder.
40

The specimen temperature was measured using a Pt/Pt-10%Rh, Type S,

thermocouple mtrinsically spot welded on the centre of the broad face, 3mm up from the

specimen's base. The thermocouple wire diameter was 0.25mm. Temperature was

monitored using a digital display with cold junction compensation. The thermocouple

output was recorded by plotting the millivolt signal directly on a Kipp and Zonan, having a

pen speed of 2mm/second and operated on a 20mV full scale. The specimen thermal

history was monitored during heating, soaking and quenching, the chart recorder plots

being used to quantify the thermal cycle.

Heating rates were a function of the furnace temperature and specimen thickness.

The vertical position of the specimen was kept constant for all tests, at a location 25mm

from the furnace temperature control thermocouple. Soak temperatures were 1000C,

1050C, 1100C and 1150C. Due to the expected grain size, specimens 3mm thick were

used at the highest temperature. At lower temperatures, 1.5mm thick specimens were

used. The surface of the specimens generally reached the desired temperature within 3

minutes of entering the furnace.

After reaching temperature, specimens were held for periods of up to 420 seconds.

The reproducibility of the thermal history for each specimen was verified by comparing the

chart recorder plots in the temperature range above 900C. Each specimen was assumed

to have reached temperature when the temperature interpreted from the plot was within

5C of the desired soak temperature. Figure 4.5 illustrates a typical furnace thermal cycle.

Following the isothermal soak, specimens were quenched by emersion in an agitated

solution of iced salt water, maintained at a temperature of approximately -10C. The salt

water bath was placed at the bottom of the vertical tube, as illustrated in Figure 4.4, so

that the transport time from the furnace to the quenchant was minimized.
41

Figure 4.5 - Typical furnace thermal cycle


42
4.3 The Thermal Cycles

Most thermal cycles involved simply heating specimens to the desired temperature,

holding, then quenching. However, some Gleeble tests used more elaborate treatments,

such as, multiple heating rates and hold times, and periods of controlled cooling prior to

quenching. The heat treatments can be divided into four groups: 1. Standard Gleeble

thermal cycles, 2. Furnace tests, 3. Stepped Gleeble thermal cycles, and 4. Continuous

cooling tests.

Figure 4.6 shows a schematic time versus temperature plot for a standard Gleeble

thermal cycle. Table 4.2 gives an overview of all standard Gleeble tests which were

performed.

Specimen Wall Heat. Hold T, C Hold t, seconds.

thick., Rate

mm C/sec.
A36 as-received 2 5 900 120
A36 as-received 2 5 950 10, 120, 600
A3 6 as-received 2 5 1000 10, 60,120, 600, 900
A36 as-received 2 5 1050 1,10, 60,120, 300, 750
A36 as-received 2 5 1100 1,10, 30, 60, 120, 300, 600
A3 6 as-received 2 5 1150 1,10, 30, 60,120, 300, 450
A36 as-received 2 5 1200 300, 600
A3 6 as-received 2 50 950,1050,1150 120
A36 solution treated 2 50 950,1050,1150 120
DQSK as-received 1 5 1050 120
DQSK as-received 1 5 1100 60, 120
DQSK as-received 1 5 1150 60

Table 4.2 - Standard thermal cycles performed on the Gleeble.


time

Figure 4.6 - Standard Gleeble thermal cycle


44

A typical furnace thermal cycle is illustrated by the schematic time versus

temperature plot shown in Figure 4.5. Table 4.3 summarizes the standard furnace cycles

which were used.

Specimen Aim hold T, C Aim hold t,

seconds.

DQSK, 3mm thick 1150 0, 60, 120, 240, 420

DQSK, 3mm thick 1100 0, 60, 120, 240, 420

DQSK, 1.5mm thick 1100 0, 420

DQSK, 1.5mm thick 1050 0, 60, 120, 240, 420

DQSK, 1.5mm thick 1000 0, 60, 120, 240, 420

Table 4.3 - Standard furnace thermal cycles

Stepped thermal cycles applied to A36 specimens on the Gleeble are described in

Figure 4.7 and Table 4.4.

Specimen HeatRatei, Hold Holdfi. Heat.Rate2 PeakT Hold Hold t 2

C/sec. T\ sec. C/sec. C. T C


2 sec.

A3 6 as-received, 5 900 120 100 1100 1100 1, 10, 30,

2mm tube 120,600

A36 as-received, 5 900 120 100 1100 1050 1, 10, 30,

2mm tube 120, 750

Table 4.4 - Stepped Gleeble thermal cycles


45

time

Figure 4.7 - Stepped Gleeble thermal cycle


46

Tests performed to examine grain growth kinetics under conditions of continuous

cooling are described in Figure 4.8 and Table 4.5.

Specimen Hold time, sec. Quench Temp.,C

A36, as-received, 1, 5, 10, 20, 45, 60 1120, 1110, 1100,

2mm wall 1080, 1030, 1000.

Table 4.5 - Continuous cooling thermal cycles

4.4 Revealing the Prior Austenite Microstructure

Specimens to be used for grain size measurement were cut at the thermocouple

location, to ensure a known thermal history. Specimens were mounted in a cold setting

aery he resin and subsequently ground and polished to a l[im diamond finish. A number of

different etchants were used, depending on the condition of the specimen. These etching

procedures are described in Table 4.6.

4.5 Measurement of the Austenite Grain Size

Polished and etched specimens were photographed on a Leitz Vario-Orthomat

metallograph system using black and white Polaroid Type 55 film at magnifications from

x80 to x900. The revealed austenite grain boundaries were traced on transparent plastic

film using felt tipped pen.

From the traced austenite structure, the mean grain area, A , was determined by
m

Jeffries' method [41]. Each whole grain was counted once and each partial grain, cut by

the edge of the field of measurement, was counted as a half grain, as described by A S T M

standards [41]. From the mean area, the mean equivalent area diameter, d , Am was

determined using equation 2.11.


47

cooling, 2C/sec.

time

Figure 4.8 - Continuous cooling thermal cycle


48

Specimen Condition Etchant Reference

He quenched A36 4% picric acid in ethanol. [56]

Immerse at room temperature for 1 to 5 minutes.

H 0 quenched A3 6
2 Saturated aqueous picric acid or [56]

lOOmL of saturated aqueous picric acid with 4g of sodium

dodecylbenzene in lOOmL H,0 and 10 drops of Triton X- [57]

100 surface active agent.

Immerse and swab at 60C to 80C for up to 10 minutes.

DQSK Immerse in 2% Nital, room temperature, 20 to 30 seconds, [58]

(coarse grains) followed by lOg sodium metabisulfite in lOOmL H,0, room

temperature, 10 to 30 seconds. Do not clean after latter step.

DQSK 4% picric acid in alcohol. [56]

(fine grains) Immerse at room temperature for 1 to 5 minutes.

1080 4g picric acid and 50g NaOH in 250mL H,0. [58]

Immerse, boil and swab sample, 10 minutes then light lum

polish, followed by light 1% Nital etch.

Table 4.6 - Summary of etching procedures.

A number of the traced microstructures were examined using image analysis

equipment with C I M A G I N G systems SIMPLE software. The system uses a screen of

480 by 640 colour pixel points. The area of each whole grain.was measured individually

from the traced microstructure using standard A S T M methods [40]. An image processing

routine was developed so that the width of grain boundaries detected by the system was
49

on average, 2 pixels. The area of a grain was defined as its net internal area plus half the

area of its boundary, so that the pixels taken up by the boundary could be accounted for. It

was verified that the sum of the area of each grain was equal the total area of the image.

The image analyzer did not measure the area of partial grains. That is any grains which

touched or were cut by the edge of the field of measurement were ignored. From the mean

area, A , the mean equivalent area diameter, d ,


m Am was determined. From each measured

grain area, A, an equivalent area diameter, d , was also determined, and the two
A

dimensional grain size distribution,/^), was assessed.

The mean linear intercept length, l ,


m was measured on some of the traced

microstructures using Heyn's method, as described by A S T M standards [41]. A

rectangular grid of 12 lines was placed randomly over the tracing. Each intersection of a

grain boundary with a line from the grid was counted as 1. A line which touched a

boundary but did not cross it was counted as a half, and a line which exactly intercepted a

grain corner, where three boundaries were seen to meet, was counted as 1.5. The mean

linear intercept length was determine by dividing the total grid length by the number of

counted intercepts.

The mean linear intercept was also determined for some specimens from the traced

structure using the image analyzer and standard A S T M methods [40]. The analyzer used a

grid made up of all of the pixels on the screen. An image processing routine was used to

reduce the width of each boundary to 1 pixel. The mean intercept length was calculated as

the total number of pixels in the screen divided by the number of pixels which made up the

boundaries.
CHAPTER 5

RESULTS AND DISCUSSION I.

T H E KINETICS OF AUSTENITE GRAIN GROWTH

5.1 The Structure of the Quenched and Etched Specimens

5.1.1 Helium Quenched A3 6 Steel

Specimens of the A36 steel were heat treated on the Gleeble then quenched using

helium, and etched in 4% picral. Appropriate quench rates were determined to be between

100C/second and 200C/second. Figure 5.1 is a photomicrograph of an A36 specimen

heated to 1150C at a rate of 5C/second and held for 60 seconds before being quenched

at a rate of 120C/second. The mean equivalent area diameter, d , is 144.0fJ.rn.


Am

The structure is approximately 10% ferrite, with the remainder being a combination

of pearlite, bainite and martensite. During the transformation, ferrite, which is white in the

print, has nucleated on the prior austenite grain boundaries in the form of grain boundary

allotriomorphs; almost all boundaries appear to be decorated in this way. Due to the high

cooling rate, secondary Widmanstatten side plates have grown out from the grain

boundary ferrite, creating a feathery appearance. At the boundary between the growing

ferrite and the austenite, pearlite has nucleated, and it is visible as the dark structure inside

the austenite grains. Within some grains, acicular regions are also visible and are thought

to be upper bainite. However, detail regarding these non-equihbrium phases is

unimportant for the purposes of grain size measurement, as long as the prior austenite

boundaries are decorated with ferrite. Near the top left corner of the print, a lightly etched

region within a larger grain is visible, and laths of martensite are easily distinguished.

50
51

500pjn

Figure 5.1 - A36 Steel, Gleeble heated at a rate of 5C/sec to 1150C and held for
60 seconds. He quenched. Magnification is approximately x85.

Figure 5.1 shows features typical of all of the hehum quenched A36 specimens.

Larger grained specimens generally contained more regions of martensite within the

austenite grains, while smaller grained specimens had higher proportions of ferrite and

pearlite. The rate of formation of ferrite increased with decreasing austenite grain size,

requiring higher quench rates to obtain the necessary austenite grain boundary

identification. Specimens with an austenite grain size (mean equivalent area diameter),

d ,
Am of less than 20(im transformed at such a high rate that the formation of excessive

ferrite could not be avoided without the use of a water quench.


52

5.1.2 Water Quenched A36 Steel

Fine grained specimens of A36 were water quenched in the Gleeble attaining cooling

rates between 300C/second and 600C/second. Figure 5.2 is a photomicrograph of an

A36 specimen heated to 1000C at a rate of 5C/second and held for 10 seconds. The

specimen was water quenched and etched in aqueous picric acid. The mean equivalent

area diameter of the austenite grains, d , is 14.5p:m.


Am

50um

Figure 5.2 - A36 Steel, Gleeble heated at a rate of 5C/sec to 1000C and held for
lOseconds. H 0 quenched. Magnification is approximately x820.
2

The water quenched structure is almost entirely martensitic. The prior austenite

boundaries appear to be dark since they have been preferentially attacked by the etchant.

Some boundaries are decorated by nucleating allotriomorphic ferrite, but significant


53

growth has been prevented by the high quench rate. The action of the etchant is likely to

be due to the segregation of solute to the austenite boundaries, and the lack of solute

redistribution during the martensitic transformation [57]. Residual strains from the

transformation may also promote the etchant attack.

5.1.3 Water Quenched DQSK Steel

D Q S K specimens were water quenched after heat treatment in the Gleeble or in the

vertical tube furnace. Specimens with mean austenite grain size, d , Am of less than 40(i.m

were etched using a 4% picral solution, while those with larger grains were etched in nital,

then sodium metabisulfite. Figure 5.3 is a photomicrograph of a fine grained D Q S K

specimen heated to 1050C in the tube furnace, held for 420 seconds before water

quenching. The specimen is etched in 4% picral to give mean equivalent area diameter,

d ,
Am of 36.2|im.

The prior austenite boundaries are decorated by ferrite, but the ferrite distribution is

non-uniform along the boundaries. Within the austenite grains, the darker structures of

pearlite and bainite are visible. The rate of the transformation from austenite to ferrite is

limited by the rate of redistribution of alloying elements (carbon and manganese). Since

the alloy content of D Q S K is relatively low, the transformation occurs very quickly. The

low alloy content also affects the quality of the etch. The quench rate was achieved by

immersing the specimen in an agitated, iced solution of saturated salt water at -10C.

Although the D Q S K steel did not etch as well as the A36 steel, some signs of the location

of the prior austenite boundaries are present. However, a more ideal structure would be

preferred.
54

l(X)|im

Figure 5.3 - D Q S K Steel, furnace heated to 1050C and held for 420seconds. H 0
2

quenched. Magnification is approximately x420.

The rate of the transformation for more coarsely grained austenite was lower and a

different microstructure was observed. Figure 5.4 is a photomicrograph of a DQSK

specimen heated to 1150C in the vertical tube furnace and held for 227 seconds before

being quenched in the iced salt water solution. The specimen was etched using nital then

sodium metabisulfite, showing a mean equivalent area diameter, d , of 168.0|im.


Am

The structure is similar to that of the helium quenched A36, in that ferrite has

nucleated first as grain boundary allotriomorphs from which secondary Widmanstatten

side plates have grown. Adjacent to the decorated austenite boundaries, regions of darker

etching pearlite can be seen. Also, grains of polygonal ferrite have formed within some of

the smaller grains, nucleating at the interface between the grain boundary ferrite and the
55

receding austenite. As observed in the A36 steel, some of the intermediate regions may be

upper bainite.

50f)Lim

Figure 5.4 - D Q S K Steel, furnace heated to 1150C and held for 227seconds. H 0
2

quenched. Magnification is approximately x85.

5.1.4 Water Quenched 1080 Steel

The 1080 steel was used in a previous grain growth study [55] because it could be

quenched to martensite and etched to reveal prior austenite boundaries. The structure is

shown in Figure 5.5, for a specimen heated to 1100C and immediately water quenched.

This heat treatment produced a mean equivalent area diameter, d , of 67.2|im.


Am
56

500|jm

Figure 5.5 - 1080 Steel, furnace heated to 1100C and immediately H 0 quenched.
2

Magnification is approximately X85.

A number of dark, parallel streaks can be seen. These are thought to be MnS

inclusions, as they were visible prior to etching.

5.2 Measurement of the Grain Size and Estimation of the Error

5.2.1 Interpretation of the Structure

A transparent tracing of the prior austenite grain structure was produced from each

photomicrograph, to be used for the measurement of grain size. Depending on the

microstructure, it was sometimes difficult to determine the location of the boundaries. As

a guide, the following assumptions were made to assist in the interpretation. Firstly, grains

do not exist within the boundaries of another grain. Secondly, grain corners are at the
57

intersection of three boundaries; the angle between the boundaries is assumed to be

approximately 120. Finally, the grain shape is equiaxed.

The water quenched A36 and 1080 specimens were relatively easy to interpret since

the structure was predominantly martensitic and the appropriate etchant produced an

image with good contrast.

The D Q S K and the hehum quenched A36 steels were examined on the assumption

that the grain boundaries were highhghted by the nucleation and growth of ferrite

allotriomorphs. In some cases, extensive growth of Widmanstatten side plates

complicated the structure. In regions where little ferrite was visible, the orientation of the

pearlite provided some clues as to the prior austenite structure. In all cases, when the

visible structure was discontinuous, the typical geometry of a network of grains was

considered.

5.2.2 Measurement of the Grain Size

The results of the grain size measurements are summarized in the Appendix. Most

specimens were measured using A S T M standard methods, which recommend that at least

fifty grains be visible in each field [40,41]. For some specimens, high magnification was

required in order to accurately interpret the structure and as a result, fewer grains were

visible in the field of view. The magnification was chosen so that a compromise was

reached between the number of grains visible and the practical interpretation of the

structure; both factors contribute to errors in the measurement. In situations where the

total number of grains from all fields was less than 100, the results were considered to be

only an estimate of the grain size rather than a statistically valid measurement.

Some specimens had a structure of distinctly separated fine and coarse grains, as

illustrated in Figure 5.6. The area fraction of each region was measured using the image

analyser, by A S T M standard methods [42].


58

Mean equivalent area diameter, d ,Am and mean linear intercept length, l , were
m

determined by the use of manual methods and by the use of image analysis [40,41]. A

more complete description of the measurement methods used is found in Section 4.5.

250|jm

Figure 5.6 - A36 Steel, Gleeble heated to 1000C and held for 120seconds. He
quenched. Magnification is approximately x410.
59

5.2.3 The 95% Confidence Intervals

For each of the measured parameters, a 95% confidence interval was calculated. The

measured value in each field was assumed to be a variable distributed normally about a

true mean value. The student's ^-distribution was used for the calculation of the confidence

interval, as described by any general statistics text [65]. In most cases, five measurement

fields were used.

The 95% confidence intervals for the manually measured values of linear intercept

length, l had a maximum of 14.0%, a minimum of 2.2% and a mean value of 8.2%.
m

Similarly, the 95% confidence intervals for the manual measurement of the mean

grain area, A , had a maximum of 23.8% and a nidnimum of 1.4%, with the mean value
m

being 10.6%. From the mean area, an equivalent area diameter, d , Am was calculated; it is

this value that is actually of interest. Since the equivalent area diameter is a function of the

square root of the area, the confidence intervals become smaller and are a maximum of

12.3% and a minimum of 0.7%, with the mean value being 5.3%.

The mean equivalent area diameter measured on the image analyser had a 95%

confidence interval with a maximum value of 12.3%, a minimum of 1.0% and a mean

value of 5.5%.

A complete list of the 95% confidence intervals are reported with the results in

Tables A.1 to A.7 in the Appendix. The confidence intervals are a reflection of the

variability of the structure since they are calculated from measurements made on a number

of different fields from the same specimen. A large interval indicates that the grain size in a

specimen has some degree of non-uniformity while a small value indicates that the

structure is relatively uniform. Generally, the larger confidence intervals correspond to

specimens where the austenite grain structure was more difficult to reveal.
60

5.2.4 Systematic Errors in Measurement of Equivalent Area Diameter

Comparison of the equivalent area diameter measured manually and that obtained on

the image analyser reveals that the manual measurement was consistently higher, by an

average of 7%. Calibration checks showed that the area measured for a single object by

each method was the same. The differences between the mean equivalent area diameters

were found to be a result of the treatment of grains which touch the edge of a field. Using

Jeffries' method [41], grains which touched the edge of the frame were counted as half

grains. Using the image analyser, grains which touched the edge were ignored. This

implies that in order to get the higher result, the structure intersected by the edge of the

field consistently had larger grains than the structure in the centre of the image.

In order to explain the larger grains at the edge of the image, the following

argument must be considered. A single large grain on a plane is more likely to be

intersected by a random line than a single small grain. In fact, the probability of

intersection is directly proportional to the grain diameter. If it is assumed that the spatial

distribution of grains in a structure is random, it follows that the probability that a grain

will be intersected by the edge of a field is proportional to the diameter of the grain, since

the location of the edge of the field is essentially a random line. Obviously, the probability

of intersection is also proportional to the number of grains of that diameter which are

present in the distribution. Therefore, large grains are intersected by the edge of the field

more often than small grains. The grain size distribution measured by image analysis

ignores the edge grains and therefore skews the result towards a distribution with a lower

mean.

This argument applies to all equiaxed structures and it can account for almost all of

the discrepancy between the two mean equivalent area diameter results. Less grains are

measured by the analyser than by the manual method. The size distribution of the grains

which are ignored by the analyser can be estimated by assuming that the probability that
61

they are in contact with the edge of the field, Pp{d), is proportional to the equivalent area

grain diameter, d , and to the fraction of grains in the two dimensional distribution, f , as
Ai Ai

described by

k
^fAj Aj
d

j =1
5.1

As an example, the specimen of 1080 steel heated to 1050C and quenched had a

mean equivalent area diameter, d , Am of 50.3|im, measured on 1142 grains using the image

analyser. The mean equivalent area diameter measured by Jeffries' method [41] was

53.6(im from 1319 grains. Note that the grains cut by the edge of the field are counted as

half grains by Jeffries' method. Therefore, 354 grains were ignored in the image analyser

measurement. The grain size distribution of the missing edge grains can be calculated by

determining the number of edge grains in each size class, N , using Fi

354
^Fi ~ k fAAAI
^fAjd jA

j=1
5.2

To remain consistent with Jeffries' method, half of this number can be added to the

number of grains measured in each size class, N , by the image analyser, to obtain the
Ai

corrected number of grains,

N Fi

N ci=-r+N
A 2 Ai 5.3

The correction which is proposed here is consistent with Jeffries' method, since only

half of the edge grains are added. Rejection of all edge grains, as is done during the image
62

analysis, skews the distribution to a lower mean since more large grains are rejected than

small ones. On the other hand, consideration of all of the edge grains would skew the

distribution toward a higher mean. The intermediate approach of considering half of the

edge grains is the most appropriate because it will give the best approximation of the true

two dimensional grain size distribution.

The corrected fraction of grains in each size class can then be determined by

r _ NAa
JACi ~ k

1 ACJ
N

i=1
5.4

The true fraction of grains in each size class, f , Ai is initially not known, but the

fraction measured by the image analyser can be used as a first approximation. The

corrected fraction, f a,
A
w u
^ b e c
l s e r
to the true fraction, and so equations 5.2, 5.3 and

5.4 can be applied iteratively, by substituting f ACi for f Ai in equation 5.2. The corrected

grain size distribution converges to a final solution in approximately ten iterations. The

final distribution will be closer to the true two dimensional grain size distribution than the

measured result obtained from the image analyser, since the skewing effect of the edge of

the field of measurement is accounted for.

Returning to the example, a corrected mean equivalent area diameter of 52.8|im is

calculated which compares well with the Jeffries value of 53.6|im. Without the applied

correction, the discrepancy between the two results is 6.0%. With the correction applied,

the discrepancy is reduced to 1.5%, which in view of the required accuracy is negligible.

Results from other examples are reported in Table 5.1. The difference between the results

are reported as a percentage of the Jeffries result.


63

Steel Temp time Jeffries Jeff. Image LA. diff. corrected diff.

C sec. no. analyzer no. % result, %

d , urn grains
Am d ,\im
Am grains d , |lm
Am

1080 1050 0 53.6 1313 50.3 1142 -6.1 52.8 -1.5

1080 1050 120 59.8 1055 55.3 885 -7.6 58.8 -1.6

1080 1050 420 70.7 756 64.7 629 -8.5 67.8 -4.2

A36 1000 600 92.3 229 81.5 185 -11.7 87.4 -5.3

A36 1150 1 99.8 237 89.7 193 -10.2 103.4 +3.6

A36 1150 60 144.0 281 134.6 206 -6.5 149.8 +4.0

DQSK 1000 0 29.0 352 27.3 299 -6.1 27.7 -4.5

DQSK 1000 120 32.3 356 31.3 271 -2.9 32.4 +0.3

DQSK 1150 420 178.1 298 162.1 231 -9.0 174.6 -2.0

Table 5.1 - Correction of systematic errors for the mean equivalent area diameter

The corrected distribution can be assumed to be closest to the true two dimensional

grain size distribution. The mean equivalent area diameter, d , from Jeffries' method [41]
Am

is always closer to the true mean equivalent area diameter than the result obtained directly

from the image analyser. The corrected mean result is generally within the 95% confidence

hmits of the Jeffries result. A l l results are affected by the statistics of the measurement and

Table 5.1 demonstrates that the agreement between Jeffries' results and the corrected

mean equivalent area diameter improves as the measured number of grains increases. The

scatter in the relative magnitude of the difference between Jeffries' results and the true

mean result can be attributed to the influence of measurement statistics.


64

5.2.5 Systematic Errors in Measurement of Linear Intercept Length

Comparison of linear intercept results determined on the image analyser and by

manual methods also show that the manual result is larger, by an average amount of

approximately 15%. The difference can be attributed to the way that each method

determines the number of intercepts. Manually, the intersection between the grid and a

grain boundary is counted as one intersection. Using the image analyser, the grid length is

considered to be the total number of pixels on the screen, and that is 640 by 480. The

number of intercepts is counted as the number of pixels which make up the area of the

grain boundaries. The matrix of pixels in the screen can be assumed to be a grid of 480

horizontal lines that are 640 pixels in length. Using Heyn's method, a grain boundary or a

portion thereof, that is horizontal on the screen will only constitute a single intercept.

Using the image analyser, a horizontal boundary will be counted by the analyser as several

intercepts, with the number of intercepts being equal to the number of pixels in the

horizontal portion of the boundary. Therefore the image analyser counts more intercepts

than Heyn's method. The mean linear intercept length is equal to the total line length

divided by the number of intercepts. Therefore, the analyser result is lower than the Heyn's

measurement. The true mean linear intercept length, l , is closest to the value obtained
m

manually, using Heyn's method. Linear intercept measurements performed with the image

analyser should thus be viewed with caution. The argument above could also be applied by

assuming that the image analyser screen is a grid of 640 vertical lines, 480 pixels in length.

5.3 The Kinetics of Austenite Grain Growth

5.3.1 A36 Steel Heated to Temperature at 5C/second

Mean equivalent area diameter measured by Jeffries' method [41] is plotted against

time in Figure 5.6, for as-received A36 steel specimens heated at a rate of 5C/sec. The

symbols are the data points and the lines show the fit obtained using the power law, to be
65

described in Section 5.5. The scatter of results is reasonable when the magnitude of the

errors is considered.

At all temperatures, the grain size is seen to increase with time and the growth rate

increases with rising temperature. At 1000C for 60 seconds and 120 seconds and at

1050C for 10 seconds and 60 seconds, two grain sizes are reported to exist

simultaneously, for each specimen. Under these conditions, abnormal grain growth was

observed and separate, measurable groups of grains were present. The grain size

distribution appeared to be bimodal. As an example, Figure 5.6 is a photomicrograph of

the specimen held at 1000C for 120 seconds.

The data appears to be divided into two sets, with one group of results being less

than 50|im and the other being greater than 80(im. The finer structures were observed

after heat treatment at lower temperatures and shorter times. To explain the discontinuous

growth between 50|im and 80|nm, the presence of a second phase precipitate, having a

boundary pinning effect is assumed. At 950C the pinned structure is stable for times up to

600 seconds. A t 1000C and 1050C the fine structure is unstable and with time, abnormal

grain growth begins as the pinning particles dissolve or coarsen leading to a reduction of

the pinning force [31,32].

In the coarse structure, the data shows, that the rate of grain growth slows down

substantially with increasing time; at times greater than 450 seconds at all temperatures, a

hmiting grain size is being approached. At 1000C and 1050C there is little growth of the

coarse structure, once abnormal grain growth has ceased. In the specimens held at 1100C

and 1150C, a substantial amount of grain growth has occurred during heating, since the

initial grain sizes are greater than 80(J.m. It is likely that abnormal grain growth has

occurred during the heating period.


66

Figure 5.7 - Kinetics of austenite grain gowth in A36 steel, heating rate=5C/second.
Lines are the power law, for ^ > 8 0 L i m , / -d ' =1.51E47xexp(-840/i?r)
22
0
8 22

for d <80Lim,
Am d - -d =5.46E54.*.exp(-1291/i?r)
3 37
0
337
67

5.3.2 D Q S K Steel Heated to Temperature in the Tube Furnace

Results from the DQSK specimens heated in the vertical tube furnace are plotted in

Figure 5.8. The mean equivalent area diameters were measured manually using Jeffries'

method and curves were fit using the power law, to be discussed in section 5.5.

At 1000C and 1050C significant grain growth was not observed. At 1100C

abnormal growth was observed and a single curve does not fit well to the stepped nature

of the data. Unlike the A36 steel, measurements for the DQSK could not be made on the

fine and coarse structure separately. At 1150C, significant growth has occurred during

heating since the initial grain size is greater than 120(im. At all temperatures, the grain size

is approaching a limiting value, although at 1100C and 1150C, after 450 seconds, grain

growth is continuing, but at a decreasing rate.

Investigations of grain growth kinetics in plain carbon steels reported by other

workers show similar results to those reported here [31,55,60]. Austenite grain size is

generally reported in the range of 10|im to 250|Lim and abnormal growth is often observed

when samples are heated in the temperature range of 900C to 1150C. Results differ

primarily in the time dependence of growth. For both steels in this investigation, the

growth rate is seen to decrease significantly with time and a limiting grain size is quickly

approached. Some other investigations have reported similar growth behaviour [31,55],

while others have not observed a limiting grain size [60].

Abnormal growth and a limiting grain size are phenomena which are typically

observed when second phase particles are present. In a steel similar in composition to the

A36, Gladman investigated abnormal growth [31] and was able to quantitatively describe

the phenomena by considering the pinning effects of aluminium nitride. His calculations

cannot be applied to this study because the particle size and volume fraction of the A1N in

the as-received condition in the A36 and DQSK steels is not known.
68

200

20 H

o -h 1 1 1 1 1 1
0 100 200 300 400 500 600
Time, seconds

Figure 5.8 - Kinetics of austenite grain growth in D Q S K steel, heated in the tube furnace.
Lines are the power law, d -d - =5.02E56.texp(-1271/fl7)
5M 5 66
69

Variability in the growth kinetics of austenite in the DQSK and the A36 steel can be

partially accounted for by differences in the puuiing effects. The influence of pinning

particles is determined by the amount of aluminium and nitrogen in the steel, the

precipitation kinetics of A1N and the size distribution of the austenite grains, as described

by equations 2.4, 2.7, and 2.8. Each of these factors is affected by the thermal and

mechanical history of the material. Further, the content of alloying elements such as

carbon and manganese has an influence on the kinetics of growth since the boundary

energy and mobility are affected.

5.4 Heating Rate Effects

Differences in the heating cycle can have a significant influence on the growth

kinetics. Figure 5.9 is a comparison of austenite grain growth in the as-received A36 steel,

heated to temperature by different heating cycles. The slowly heated specimens were

heated at a constant rate of 5C/second to 1100C and held. Rapidly heated specimens

underwent a stepped heating cycle, and were heated initially at 5C/second to 900C and

held for 120 seconds, then heated again at 100C/second to 1100C. The curves in the

plot are the fit of the power law to the data.

The initial grain size in specimens heated at the 100C/second was 47|i.m while in

those heated at 5C/second the initial grain size was 88p:m. The apparent j ^ i t i n g grain

size was lOOiim in the quickly heated specimen and 140|Lim in the slowly heated one.

These different sizes can be explained by grain growth during heating, and the influence of

A1N particles. Nucleation effects during the transformation to austenite are likely to be

^significant due to the identical thermal cycles up to 900C. The difference in initial grain

size can be explained by grain growth occurring during heating since the rapidly heated

specimen spent less time at high temperature, and therefore less growth occurred. The

effect of holding time at 900C can be ignored with respect to grain growth since at this

temperature, limited grain coarsening has been observed. The different limiting grain size
70

is evidence that the dispersion of particles is affected by the heating cycle, since Zener has

described the limiting grain size as function of particle volume fraction and radius [30]. If

equilibrium conditions exist, particle volume fraction is unchanging, and the smaller

j ^ i t i n g grain size must be due to a finer dispersion of closely spaced particles. The

influence of second phase particles will be discussed in greater detail in Section 5.6.

5.5 Models Describing the Kinetics of Grain Growth

The power law and the statistical model for grain growth were fit to data from the

A36 and the D Q S K steels. Measured values of mean equivalent area diameter, d Am were

used to determine the constants for the fit of the power law. Estimated values of the

average equivalent volume diameter, d , vb were used to fit the statistical model. The

statistical model of Abbruzzese and Liicke [20,21] was applied by Militzer et al. [61,62].

5.5.1 Application of the Power Law

The power law for grain growth was fit to the results of mean equivalent area

diameter measured for the D Q S K steel heated in the tube furnace and the A36 steel heated

in the Gleeble at a rate of 5C/second and 100C/second. In each case the parameters of

m, K and Q were used to fit the power law equation by a least squares procedure. The
0

sum of the squares was minimised using a solving routine on a Microsoft Excel

spreadsheet.

Separate curves were needed to fit to fine and coarse austenite grain growth A3 6

steel heated at 5C/second. Table 5.2 is a summary of the parameters determined to

describe the results.


71

180

160 H

I 140

120

100

80

60
5C/sec(1100C)
cf -d
22
0
822
=1.51 E47.f.exp(-84CW?7)
40
100C/sec (1100C)
20 du
9
-L1 14
=1.94E68.f.exp(-1089/flT)
9

0 ~I 1 1 1 1 1 1
0 100 200 300 400 500 600 700 800
time, seconds

Figure 5.9 - Effect of heating rate on austenite grain growth in A36 at 1100C.
72

Condition m K , |im /sec.


n
m
Q, kJ/mol.

Fine A36, 5C/sec 3.37 5.46E+54 1291

Coarse A36, 5C/sec 8.22 1.51E+47 840

A36, 100C/sec 14.85 1.94E+68 1089

DQSK 5.66 5.02E+56 1271

Table 5.2 - Power law fitting parameters for the A36 and D Q S K steels, for the
equation, d -d =K(pxTp(-Q/RT).
m
0
m

Use of the power law with m not equal to 2 is empirical. When m is used as a fitting

parameter, the equation becomes a curve fitting empirical expression for describing the

austenite growth kinetics, as shown in Figure 5.7, 5.8 and 5.9. However, the reported

values obtained for m, K and Q are widely scattered. The values obtained for the A36
0

steel heated to temperature using two different heating rates are entirely unrelated, despite

the fact that results were obtained using the same steel. The equations cannot be used

outside of the experimental conditions for which they were determined. For these reasons,

the development of a more fundamentally based relationship is of considerable importance.

5.5.2 Application of the Statistical Model

The statistical model was applied by Militzer et al. [61] to the results for the A36

steel heated at 5C/second and 100C/second, and the DQSK steel. The calculations are

based on the average equivalent volume diameter, dy^, estimated from the measured mean

equivalent area diameter, d , by the method of Takayama et al. [38].


Am

The grain boundary energy, y ^ , is calculated as a function of the carbon content, as

described by Gjostein [63]. Consequently, a value of 7^=0.7Jhr is used for the A36 and
2
73

7g=0.8Jrrr is used for the DQSK steel. The grain boundary mobihty, M ,
2
gb is

approximated by that of pure austenitic iron and is calculated using the grain boundary

diffusivity, D , and the Burgers vector, b, from Frost and Ashby [64],
gb

D b ah
2

kT 5.5

Mobihty decreases quickly as the temperature decreases. The pimiing parameter, P, was

estimated, in order to fit the model to the data of each isothermal test series, for each

steel.

The fit of the model is shown in Figure 5.10 for the A36 steel heated at 5C/second

in the Gleeble, and in Figure 5.11 for the D Q S K steel heated in the tube furnace.

The model fits well to the growth kinetics of the coarse A36. A t this stage, no

attempt to fit the model to the abnormal growth condition has been reported for the A36

steel. For the DQSK, all data, including the abnormal growth results were desrcibed using

the model. Since during abnormal growth the pinning force changes with time, at 1100C,

two values of the pmning parameter were used, resulting in a good fit for the abnormal

grain growth data. The discontinuity in the curve is a result of the sudden change in P,

related to the changing pinning behaviour of the A1N particles in the material. For the A36

and the D Q S K steel results, the statistical model is more effective than the power law in

describing the experimental data.

The j ^ i t i n g grain radius, Rn , was estimated from the experimental results.


m

According to equation 2.7, Rn is the reciprocal of the pinning parameter, P , so that the
m

limiting grain radius can be predicted from the pinning force estimations made with the

statistical model. Figure 5.12 is a comparison of the experimental and predicted values of

Rli , showing good agreement between the two.


m
74

240

0 100 200 300 400 500 600 700 800


time, seconds

Figure 5.10 - The statistical model and the results for the A36 steel
for a heating rate of 5C/second. Lines are generated by the model [61,62].
75

Figure 5.11 - The statistical model and the results for D Q S K steel.
Lines are generated by the model [61].
0 20 40 60 80 100 120 140
R lim (experimental), (im

Figure 5.12 - Predicted and experimentalR u [61,62].


77

The results indicate that the statistical model can be applied to describe the grain

growth kinetics using a single fitting parameter, P. Furthermore, the value of P can be

used to predict the jjiniting grain radius with reasonable accuracy, demonstrating a

consistency between the experimental evidence and the model's theoretical basis. Since the

statistical model is developed from fundamental principles, it can be extrapolated to

describe grain growth kinetics under a wide range of conditions, as long as reasonable

assumptions are made regarding the pinning forces which exist.

The aim in applying the statistical model is to describe the kinetics of grain growth

during hot strip rolling, under conditions which are not isothermal. A test series was

performed with the A36 steel to measure the kinetics of grain growth during continuous

cooling conditions. During hot rolling, between the roughing mill and the finishing mill, a

typical temperature change from 1120C to 1000C can occur over a period of

approximately 60 seconds. A36 steel specimens were heated to temperature using a

stepped thermal cycle to 1120C and a continuous cooling cycle of 2C/second was

imposed. The statistical model was applied by Militzer et al. [61], employing the

temperature dependence of the pinning parameter, P, determined during stepped

isothermal tests. Figure 5.13 illustrates the results. Good agreement is obtained,

demonstrating that the statistical model can be used to model the kinetics of grain growth

under non-isothermal conditions.

5.6 The Influence of Second Phase Particles

Most of the grain growth kinetics observed in this investigation were influenced by

the presence of second phase particles, since at all temperatures, a tendency toward a

limiting grain size was observed. The A36 and DQSK steels contained similar proportions

of aluminium and nitrogen.


78

75

980 1000 1020 1040 1060 1080 1100 1120 1140


Temperature, C

Figure 5.13 - Application of the statistical model to describe continuous cooling conditions [61].
79

The temperature at which both constituents were completely soluble in both steels is

approximately 1170C, according to the solubility product,

log [Al][N] = - - + 1.48


10

- 5.6

reported by Gladman [32], where [Al] is weight percent aluminum, [N] is weight percent

nitrogen, and T is temperature in Kelvin.

In the as-received condition, the dispersion of A1N in each of the samples, water

quenched off the hot strip mill, is not known. Gladman has reported that A1N precipitates

slowly during cooling, but more quickly during heating [32]. Even if no precipitation

occurred in the steels during the initial quench on the mill, it is certain that a dispersion of

particles exists during the isothermal holding period in the furnace heated D Q S K and in

the slowly heated A36, due to the low heating rates employed. Even specimens of A36

heated at 100C/second were first taken to 900C at a rate of 5C/second and held for 120

seconds. It is likely that these also had a dispersion of A1N particles present during

isothermal grain growth.

The observed abnormal growth kinetics and the tendency to approach a jj^ting

grain size are evidence of the presence of A1N particles. Since the initial particle size

distribution is not known, accurate prediction of the kinetics of precipitation is not

possible. However Gladman's expression [32] can be used to estimate the size of the

particles which are present when abnormal growth is observed. In the A36 steel heated at

5C/second and held at 1000C for 60 seconds, if it is assumed that the initial ratio of the

largest to the average grain radius is 2.0, and A1N is in equilibrium, then the critical

particle radius is approximately 30nm. Transmission electron microscopy would be

required to observe a particle of this size; such an investigation is difficult to perform and

is outside the scope of this work.


80

It seems likely that both of the as-received steels are in a supersaturated condition

with respect to soluble aluminium and nitrogen, consistent with water quenching of the

transfer bar on the hot strip mill; but the degree of supersaturation is not known. The rate

at which solutes can precipitate and coarsen, and the rate at which precipitates can

dissolve, is diffusion limited. In either case, the distribution of particles is changing with

time, as is the pinning force. The observed phenomena of abnormal grain growth is due to

the dynamic nature of the pinning force. The limiting grain sizes are also likely to change

with time, as the distribution of pmning particles tends to dissolve or coarsen. Further, the

effect of heating rate on the Ijmiting grain size in the A36 steel can be attributed to a

change in the A1N particle distribution during heating. Additional studies are required to

disclose detail of the nature of this process.

5.7 The Evolution of the Grain Size Distribution

The two dimensional grain size distribution was measured using the image analyser.

Changes in the distribution during grain growth will be discussed for the case of normal

grain growth, as occurs in the A36 steel heated to 1150C at 5C/sec, and for abnormal

grain growth as observed in the A36 steel heated to 1000C at 5C/second. As a measure

of the log normality of each distribution, a correlation coefficient, R , will be reported


2

[65]. Calculation of R is described by


2

i(z,-X f c )(i--Fj
i=i
R2
=

where X ; is the measured fraction of size class i and 7; is the calculated fraction of size

class /, determined from the log normal distribution described by equation 2.11, the peak,

d , and the standard deviation, s . X is the average of all X and Y is the average of all
Ag A b t b
81

Y . The log normal distribution is determined by solving for the best fit of the log normal
t

curve to the measured results. A least squares solution, similar to that used to fit the

power law, is applied, where the sum of the difference between calculated and measured

fractions is minimised. Values of the standard deviation, s , can be calculated directly


A

from the measured distribution using the general equation,

s = exp 2/i(l^)-lnG/))'
i=l
5.8

A l l diameters are mean equivalent area diameters, d .


Am A l l standard deviations and

diameters in this section are measured using the image analyser.

Figure 5.14 shows the evolution of the measured two dimensional grain size

distribution during normal grain growth at 1150C, at times of 1 second, 60 second and

450seconds. The curves shown are the best fit log normal distributions. Table 5.3

summarizes the relevant parameters.

Steel Temp,C time, sec. d ,


Am (im i?2

A36 1150 1 89.7 1.71 0.96

A36 1150 60 134.6 1.58 0.99

A36 1150 450 187.0 1.62 0.97

A36 1000 10 14.4 1.59 0.97

A36 1000 120 47.8 2.40 0.85

A36 1000 600 81.5 1.63 0.97

Table 5.3 - Evolution of the grain size distribution


82

0.20 450 sec.

o 0.15 H
+

o
CO

Lt 0.10

0.05 H

0.00 i r- ' i' ' ' ' i' ' ' ' i' ' 1
' i' r
20.0 31.3 48.8 76.3 119.2 186.3 291.0 454.8
Grain size, EQAD, |nm
Figure 5.14 - Evolution of the A36 2-D grain size distribution at 1150C.
83

During normal grain growth at 1150C, the relative width of the distribution does

not change greatly with time, but the peak steadily increases. The shape of the distribution

is closely approximated by a log normal distribution at all times. The size distribution

evolves by the shrinkage and disappearance of the smaller grains and the coarsening of

large grains. The distribution observed at 1 second is the widest, possibly due to the after

effects of abnormal grain growth during heating. However, the distribution width is

observed to be stable and constant at 60 seconds and 450 seconds. Therefore, after a short

time at 1150C, normal grain growth is observed to occur by scaling, as described by

Atkinson [16].

Photomicrographs showing the structure before, during and after abnormal grain

growth in the A36 steel held at 1000C, are shown in Figure 5.15. At 10 seconds, the

specimen was water quenched so that the fine grains could be revealed. At 120 seconds

and 600 seconds, a helium quench was employed. The development of the structure over

time is visibly dramatic, being uniform and fine at 10 seconds, bimodal at 120 seconds and

finally uniform and coarse at 600 seconds.

The evolution of the two dimensional grain size distribution during abnormal growth

at 1000C is shown in Figure 5.16. Initially, at 10 seconds, the structure is uniform and

fine. A t the onset of abnormal growth, the largest grains in the structure break away from

the changing dispersion of pinning particles, since they have the greatest driving force for

growth. This is evident in the widening of the distribution and the appearance of the

bimodal structure at 120 seconds. The plot at 120 seconds suggests the existence of two

peaks, consistent with the observed bimodal structure.


(a) Uniform fine structure at 10 seconds

(b) Bimodal structure at 120 seconds

(c) Uniform coarse structure at 600 seconds


250um

Figure 5.15 - Evolution of the grain structure, in the A36 steel held at 1000C,
showing abnormal grain growth. Magnification is approximately x410.
85
0.25
10 sec.
0.20 H

c
0.15
o
cd
Lt 0.10 H

0.05 H

0.00

i 1 1 i i i i i r
4.0 6.3 9.8 15.3 23.8 37.3 58.2 91.0 142.1 222.0
Grain size, EQAD, (xm
Figure 5.16 - Evolution of the A36 2-D grain size distribution at 1000C.
86

The reduced correlation coefficient in Table 5.3 demonstrates that the distribution

deviates from log normality. When the pinning particles are sufficiently dissolved or

coarsened, all grains are able to grow, and the fine grains quickly disappear. The structure

regains its uniform appearance as the distribution narrows and the distribution shape

becomes log normal again at 600 seconds.

5.9 Solution Treatment

Thus, having found a significant effect of A1N precipitates on the grain growth

behaviour, attempts were made to eliminate or minimise the effect by performing a

solution treatment. A number of A36 specimens were soaked in a furnace at 1200C for 3

hours to dissolve the A1N precipitates, then water quenched to prevent their precipitation

during cooling. A test series was performed to compare the kinetics of austenite grain

growth in the heat treated and the as-received specimens. The results of these tests are

shown in Figure 5.17. The specimens were heated at 50C/second to temperatures of

950C, 1050C and 1150C, and held at each temperature for 120 seconds. A n
J

intermediate heating rate was chosen to reduce the precipitation of A1N during heating and

to minimise temperature over shoot.

The solution treated specimens displayed a grain size which was larger than that

obtained in the as-received specimens for all thermal cycles employed. Since the solution

treatment was designed with the intent of minimising the pinning effect of the second

phase particles, it is not surprising that the solution treated specimens have a larger grain

size; a higher rate of grain growth would be expected with the reduced influence of second

phase particles.

Differences in the structure of the specimens prior to the thermal cycle may also

have some influence on the observed differences in grain size. Prior to thermal cycling, the

structure of the as-received A36 steel was mostly acicular ferrite and pearlite, while the
87

solution treated specimens were almost all martensite. The prior austenite grain size of the

as-received structure also appeared to be much finer than that of the solution treated

material. During the transformation back to austenite on heating, the initial structure of the

as-received specimen would be expected to have a refining effect on the nucleation of the

new phase, since the as-received structure would be expected to offer more nucleation

sites for transforming to austenite. A larger number of nuclei could explain the finer grain

sizes. Further tests would be required in order to conclusively state whether nucleation

effects, or the effect of second phase particle dominate the grain growth kinetics. It is

likely that both have a some influence.

The structure of the as-received specimen after heating to 950C was bimodal,

showing signs of abnormal grain growth, whereas, the structure in the solution treated

specimen heated to 950C was uniform. This result can only be explained by an initial

difference in the dispersion of second phase particles. Therefore, from these results, it can

be concluded that the solution treatment reduced the effect of second phase particles.

Both sets of results show an inflection in the plot at 1050C. The inflection indicates

that, despite the anticipated difference in the initial particle distribution and structure, a

similar trend with respect to the kinetics of grain growth is occurring in each type of

specimen. Few specific statements can be made regarding the kinetics of isothermal

growth at each temperature, since only one holding time has been used.
88

Figure 5.17 - Comparison of austenite grain size in as received and solution treated
A36 steel, heated to 950C, 1050C and 1150C and held for 120 seconds.
89

At 1150C, the grain size is much larger than in the specimens held at lower

temperatures. This result can be explained by a the combined effect of increased A1N

solubility and increased atomic mobility. The solution temperature of A1N in the A36 steel

is approximately 1170C, according to the solubility product reported by Gladman [32].

Therefore at 1150C, much of the A1N will be already in solution, or in the process of

dissolving. With fewer second phase particles present, the rate of grain growth is higher

since the pinning force is reduced. Also with increasing temperature, grain boundary

mobility is higher, thereby increasing the rate of grain growth.

5.10 G r a i n G r o w t h D u r i n g H o t Strip Rolling

In conventional process models the power law is used to describe grain growth

kinetics. It is usually fit first to results similar to those reported here, and then extrapolated

to describe grain growth occurring under the conditions of hot strip rolling.

In this work however, it has been observed that the kinetics of grain growth differ

when the heating cycle is changed or when the as-received specimens are solution treated

prior to being used for grain growth tests. The power law has been fit with different

parameters, to the results obtained even for the same material, heated to temperature by

different thermal cycles. This scatter in the fitting parameters cannot be described from any

fundamental perspective.

During conventional grain growth tests, rapid growth occurs during heating and the

initial grain size is large. A smaller initial grain size can be achieved by deformation and

recrystalhzation of the specimen. However, quenching the structure after the test, to

reveal the prior austenite grain size, then becomes difficult. Therefore, the early stages of

grain growth, which are relevant to the industrial process of hot strip rolling, are rarely

measured. In laboratory tests, the specimen must be held at temperature for extended

periods so that significant growth is observed and a reasonable time dependence for the

change in grain size can be formulated.


90

Under hot strip rolling conditions, deformation occurs in each of the rolling passes,

and grain growth occurs in the time between stages of deformation, when recrystallization

is complete. In a typical rolling mill, interpass times range from 1 second to 20 seconds

and the delay between the roughing mill and finishing mill is typically 60 seconds. Hold

times for laboratory studies of grain growth are usually in the range of 500 seconds to

1000 seconds [3,4,18], or longer [5]. Thus an extrapolation of the empirical power law to

mill conditions appears to have little validity.

Differences in the grain growth kinetics have been described with respect to the

influence of second phase particles. Abnormal grain growth and the attainment of a

limiting grain size are rarely observed during hot rolling of plain carbon steels, providing

evidence that second phase particles have little influence. Before rolling, the steel is

soaked at between 1200C and 1300C in a reheat furnace. At this temperature, ALN is

usually completely soluble. During rolling, the temperature of the material, and the

solubility of A1N drop, but Gladman states that precipitation of A1N during cooling is slow

[31]. Therefore it would be expected that A1N particles have limited effect until the later

stages of rolling, at lower temperatures where austenite grain growth is not significant.

Militzer et al. have described the application of a statistical model to the prediction of

austenite grain growth kinetics in A36 and DQSK steels during hot strip rolling [61,62]. It

is reported that a pjjming parameter of P=0 is required to model grain growth occurring

between the roughing and the finishing mills, which corresponds to unpinned grain

growth. Such a situation is realistic, since in the hot strip rnill, it is likely that little or no

ALN particles will precipitate between the roughing and fMshing mills, in the

corresponding temperature range from 1120C to 10Q0C. The kinetics of austenite grain

growth can then be described by Burke and Turnbull's parabolic growth kinetics [6], as

discussed by Militzer et al. [61].


CHAPTER 6

RESULTS AND DISCUSSION II.

ESTIMATION OF T H E THREE DIMENSIONAL GRAIN SIZE DISTRIBUTION

6.0 Introduction

Methods for the estimation of the three dimensional grain size distribution were

applied to two dimensional results obtained from the 1080, the A36 and the D Q S K steel.

The 1080 steel was used for much of the work since the microstructure was easily

interpreted and a large number of grains could be measured. The three dimensional grain

size distribution was described using the mean equivalent volume diameter, d Vm and the

log normal standard deviation, s .


v

6.1 Methods Based on the Assumption of Spherical Grain Shape

The methods of Saltikov [36] and Huang and Form [37,46] were applied to the two

dimensional grain size distributions measured on the image analyzer for the 1080 steel,

heated to 1050C and held for 0 seconds, 120 seconds and 420 seconds. Figure 6.1 shows

the measured distribution and the distribution as predicted by each method, for the

specimen held for a time of 0 seconds. The correlation coefficient, R , for the fit of the
2

measured distribution to a log normal distribution is 0.99. The results from each method

are quite different, with the mean equivalent volume diameter, dy , from Saltikov's
m

method [36] being equal to 70.2|im and that obtained from Huang and Form's method

[37,46], equal to 59.0|im.

91
92

0.24
0.22 -
0.20 -
0.18 -
0.16
0.14
0.12
0.10
0.08
0.06 -
0.04 -
0.02 -
0.00
-0.02 Measured 2-D data
-0.04 Saltikov's method
Huang and Form's method
-0.06 H
~l 1 1 1 1 1 1 1 1
6.3 9.8 15.3 23.8 37.3 58.2 91.0 142.1 222.0

Grain size, EQVD, Lim

Figure 6.1 - Saltikov's [36] and Huang and Form's [37] 3-D grain size distribution
compared to the measured 2-D distribution for 1080 steel heated to 1050C and quenched.
93

In Saltikov's original publication [36], his method required that the grain size

distribution be represented by size classes geometrically spaced with a multiplying factor,

a, of 1 0 0 1
(1.2589) since application of the method required the use of reported tables of

coefficients. In this case, a equal to 1.25 has been used. The mathematics of the procedure

were set up using a Microsoft Excel spreadsheet so that any multiplying factor could be

chosen, since the principles of the calculation are independent of a. The major difference

between the two methods is that Saltikov [36] corrects for sampling only, while Huang

and Form [37,46] correct for both sampling and truncation. Up to the point where the

truncation correction is applied, the two methods are identical. The truncation correction

increases the number of smaller grains and decreases the number of large grains, since it

accounts for the probability that a grain is sectioned, based on the grain diameter. It is

therefore expected that the mean predicted by Saltikov is greater than that predicted by

Huang and Form; this is seen to be the case. The width of the estimated distribution

cannot be assessed by calculation of the log normal standard deviation because the

negative number of grains in the smallest size classes makes such a calculation impossible.

The prediction of a negative number of grains, as shown in Figure 6.1, illustrates

one of the major problems with both of these methods. Even when an artificially

generated, perfect log normal distribution was used as input for the methods, a negative

number of grains was predicted in the smaller size classes. This is obviously an unrealistic

result. Huang and Form [46] and Aaron et al. [49] have reported similar negative counts

when using these techniques.

Both methods operate by correcting the number of grains in each size class, by

application of an equation of summation, as described by equation 2.17. The number of

terms in the equation is equivalent to the number of corrections applied, and is a function

of the relative magnitude of the size class, with the largest class having a single correction

and the nth largest class having n corrections. Therefore, any errors which result from
94

assumptions inherent in each method are compounded for the smallest size classes. The

assumption that grains are spherical in shape is incorrect, as stated by Aaron et al. [49],

because an array of spherical grains cannot fill space. The methods are based on the

probability that an object is sectioned to give a plane of lesser diameter. It has been shown

that such probabilities are a function of object shape [36,39]. It is the probability

calculation resulting from the assumption of spherical grain shape that directly introduces

the errors observed.

6.2 The Method of Takayama et al.

The three dimensional grain size distribution was estimated from the measured

results of mean linear intercept length, l , and mean area, A , using the method of
m m

Takayama et al. [38]. This method was applied to all of the A36 specimens heated at

5C/second and 100C/second, and to the 1080 steel specimens held at 1050C.

Some of the results for the A36 steel heated at 5C/second are shown in Table 6.1.

Mean linear intercept, l , was determined by manual measurement using Heyn's linear
m

intercept method [41]. The measurement involved the placement of a grid over a

photomicrograph of the structure so that number of intercepts per unit length could be

determined. Mean area was measured by a manual procedure using Jeffries' method [41].

Photomicrographs of the structure were used to determine the number of grains per unit

area. The correlation coefficient, R , for the fit of the measured distribution to a log
2

normal one is reported for specimens where the two dimensional distribution was also

determined.

Values of the standard deviation of the log normal distribution, s , were calculated
v

directly from the measured results by using equations 2.21 and 2.22. Calculated values of

s are reported in Table 6.1, as are overall mean values. The calculated and mean values
v

of Sy were used to determine dy . m


95

Steel 7/,C t R2 From cale. s v From i lean s v

sec. / dVm sv dvm sv

|im (im 2
|_im (im

A36 1000 10 0.97 11.6 165 17.7 1.31 17.7 1.33

A36 1000 600 0.97 78.8 6687 105.6 1.57 112.7 1.33

A36 1100 1 73.4 6036 102.6 1.50 107.0 1.33

A36 1100 10 67.2 5836 108.5 1.15 105.3 1.33

A36 1100 30 88.7 10387 146.2 1.02 140.4 1.33

A36 1100 60 103.6 13108 157.9 1.32 157.7 1.33

A36 1100 120 106.4 13593 159.4 1.36 160.6 1.33

A36 1100 300 112.6 14363 159.1 1.48 165.1 1.33

A36 1100 600 113.5 15315 168.4 1.38 170.5 1.33

A36 1150 1 0.96 77.7 7819 125.6 1.15 121.8 1.33

A36 1150 60 0.99 116 16287 175.2 1.34 175.8 1.33

A36 1150 450 0.97 169 33506 247.1 1.42 252.2 1.33

Table 6.1 - The three dimensional grain size distribution estimated using the method
of Takayama et al. [38].

There is significant scatter in the s results, as reported in Table 6.1. The scatter
v

predicted by the method is not consistent with experimental observations of the two

dimensional grain size distribution. It is reasonable to assume that the width of the two

dimensional distribution is a strong function of the width of the true three dimensional

distribution. During normal grain growth in the A36 at 1150C, the two dimensional

distribution width was seen to be fairly constant, s being 1.71 at 1 second, 1.58 at 60
A

seconds and 1.62 at 450 seconds (refer to the discussion in Section 5.7). From Table 6.1,

the variation in sy ranges from 1.15 to 1.42.


96

The log normal standard deviation, s , is dependent on the ratio of l to the square
v m

root of A ,m as shown in Figure 6.2. Scatter in the measured results of l m and Am

introduced unacceptable variability in the estimated value of s . The 95% confidence


v

intervals for l were calculated to be approximately 10%. From Figure 6.2, it is apparent
m

that a 10% variation in the measurements would introduce a 30% difference in s . v

Accuracy in the determination of l and A would improve with an increase in the number
m m

of grains measured.

The mean equivalent volume diameter, d , is a function of s . In order to remove


Vm v

the unrealistic variability of s and its subsequent influence on d ,


v Vm an average value of

the standard deviation was used. Table 6.1 reports d Vm results obtained for the A36

specimens, determined by using an average standard deviation. The average standard

deviation of the log normal distribution is estimated to be equal to 1.33. Equation 2.22

leads to

^Vm = 1
-22^Am 6.1

The complete list of results from the A36 steel heated at 5C/second shows that the

average of the standard deviation is approximately 1.33 for each of the five different

temperatures: At 950C the mean is 1.31; at 1000C it is 1.38; at 1050C it is 1.34; at

1100C it is 1.32 and at 1150C it is 1.26. This indicates that the scatter in s v can be

explained by the statistical variability of the measurement of l m and A . When enough


m

measurements are used together, a statistically stable result for s can be determined; this
v

value approaches 1.33. It is unfortunate that the required number of grains seems to be of

the order of 1000, since it is not practical to propose that every measurement of grain size

be made on such a large number of grains.


97

2.4

I 2.2

2.0

1.8 H

1.6 H

1.4 H

1.2

1.0
0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 1.25
/ divided by the square root of A
m

1/2
Figure 6.2 - The relationship between s and the ratio, / /(A )
y m m , using the method of
Takayama et al. [38].
98

The variability of the l m and A m measurement was smaller for the 1080 steel

specimens, since the structure was more clearly revealed, and a much larger number of

grains were measured for each of these specimens. Linear intercept length and mean area

were determined manually [41] and by using image analysis equipment [40]. The method

of Takayama et al. [38] was applied to both sets of results.

The systematic error in the measurement of mean linear intercept, l , and mean m

area, A ,
m by the image analyzer was discussed in Section 5.2. The linear intercept length

measured on the image analyzer is much lower than that measured manually, leading to

smaller values of s , as explained by Figure 6.2. Mean area is also lower but the difference
v

is not as great. If the image analyzer results are used then a value of s equal to 1.15 is
v

estimated. Using manual methods, s is estimated to be equal to 1.43. Comparison of the


v

calculated values of d Vm and reveals that there is only a minor difference in the mean

volume diameters. Since the estimated three dimensional distribution is very sensitive to

the ratio of lm to the square root of A , the systematic error in l


m m manifests itself in an

under estimation of s . Therefore, measured values of l and A


v m m from the image analyzer

cannot be used with the method of Takayama et al. [38] to estimate the three dimensional

grain size distribution.

The method of Takayama et al. [38] applied to the 1080 steel using the manual

measurements produced reasonable results: In a specimen heated to 1050C and

immediately quenched, the equivalent volume diameter, d , was estimated to be 64.3|im


Vm

with a standard deviation of 1.41; in a specimen held at 1050C for 120 seconds, d Vm was

71.2(j,m and s v was 1.44 and in a specimen held at 1050C for 420 seconds, d Vm was

83.6|im and s v was 1.46. As would be expected during normal grain growth, the

distribution width is estimated to change little, with the mean grain size steadily increasing.
99

If the mean value of s , equal to 1.43, is used with the 1080 steel manually
v

measured results, then equation 2.22 becomes

d Vm = U9d Am 6 2

From equation 6.1 and 6.2, it can be said that the mean equivalent volume diameter,

d ,
Vm predicted by the method of Takayama et al. [38] is approximately 20% greater than

the mean equivalent area diameter, d , Am determined by Jeffries' method [41], for the cases

examined. In addition, comparison of d Vm to lm reveals that dVm is consistently 50%

greater than l . The approximate general equation,


m

dVm - l.2d Am = \.5l m 6 3

can be written.

In summary, there are two main problems with the method of Takayama et al. [38].

Firstly, the method requires that the grain size distribution is log normal. Examples

discussed in this section were observed to have log normal grain size distributions in two

dimensions; for these cases, the method is therefore apphcable. However, several

specimens were observed to have grain size distributions which were not log normal

(abnormal grain growth was observed) and in these cases, use of the method of Takayama

et al. [38] is not valid.

Secondly, the method is very sensitive to the ratio of l m to the square root of A . m

During measurement of grain size in the A36 and DQSK steels, errors in the order of 10%

were encountered because only a hmited number of grains could be measured. Errors of

this size introduce unreasonable scatter in the calculated values of s v and d .


Vm The

method works reasonably well with microstructures which are easily measured, and with
100

other structures when a large number of grains are analyzed. Thus, successful application

of the method is dependent on good measurement statistics.

6.3 The Method of Matsuura and Itoh

6.3.1 Application of the Method

The three dimensional grain size distribution was also estimated using the method of

Matsuura and Itoh [39]. The method was applied to results obtained from the 1080 steel,

the A36 steel heated at 5C/second, and the DQSK steel heated in the furnace, as

summarized in Table 6.2. The mean equivalent volume diameter, d , Vm and the standard

deviation of the log normal distribution, s , v were calculated from the estimated

distribution. A flow chart shown in Figure 6.3 demonstrates how the method is applied,

using a programming routine written in F O R T R A N .

As a first assessment of the method of Matsuura and Itoh [39], the two dimensional

grain size distribution measured directly by the image analyzer was used as input. The

method consistently predicted that the three dimensional grain size distribution has a

higher mean than the two dimensional grain size distribution. The width is predicted to

narrow slightly. A n average conversion factor of 1.23 can be calculated by assessment of

the ratio of the mean equivalent area diameter (obtained from the image analyzer), d ,Am to

the mean equivalent volume diameter, d . Vm From Table 6.2, it is obvious that there is

some variability in the results, and for the wider distributions, a higher ratio has been

calculated. A simple conversion factor can be used for cases where the width of the grain

size distribution does not vary substantially; this is the case, where grain growth is normal

and the grain size distribution is increasing by scaling.


(^""start

Determine 2-D distribution and


represent in k discrete size classes
as single columned matrix, A .

Calculate dAb, and assume equal to dvp

Calculate the number of grain faces for


each size class, Ji, using equation 2.19 and dvb.
t
Determine the relative diameter
distribution curve for each grain shape from values of Ji.

Integrate over each distribution curve


for each grain size class and
place results in a /cx/c matrix, B.

Solve matrix problem C . B = A


by determining least squares
solution for matrix C.
i
Use single columned solution
matrix, C, to calculate new result for dVb.
Compare new dvb to previous dvb.

<^STOP^)

Figure 6.3 - Flow chart of the procedure for applying the method of Matsuura and Itoh [39].
102

Measured 2-D Estimated 3-D

Steel T aim t No. sv


C sec. grains Lim |im

1080 1050 0 1142 50.33 1.73 62.85 1.59 1.25

1080 1050 120 885 55.26 1.70 68.37 1.57 1.24

1080 1050 420 629 64.69 1.63 78.21 1.49 1.21

A36 1000 10 250 14.39 1.59 17.74 1.51 1.23

A36 1000 120 300 47.79 2.40 71.06 2.25 1.49

A36 1000 600 185 81.50 1.63 104.0 1.54 1.28

A36 1150 1 193 89.65 1.71 109.8 1.61 1.22

A36 1150 60 206 134.6 1.58 166.4 1.48 1.24

A36 1150 450 115 187.0 1.62 227.0 1.52 1.21

DQSK 1000 0 299 27.62 1.36 31.84 1.28 1.17

DQSK 1000 120 271 31.33 1.38 36.39 1.29 1.16

DQSK 1000 420 199 33.51 1.35 38.62 1.27 1.15

DQSK 1150 0 184 113.8 1.51 136.9 1.43 1.20

DQSK 1150 120 161 145.7 1.50 177.3 1.42 1.22

DQSK 1150 420 231 162.1 1.58 197.9 1.49 1.22

Table 6.2 - 3-D distribution from the image analyzer results, for 1080 steel, A36
steel and DQSK steel, after Matsuura and Itoh [39].

Figure 6.4 shows the relationship between the distribution width, s , and the ratio of
A

dAm to d .Vm From the plot, it appears that a linear relationship exits between s A and

d /d . In order to confirm this relationship, it would be desirable to estimate the three


Vm Am

dimensional grain size distribution for specimens having a two dimensional log normal
103

standard deviation, s , in the range from 1.8 to 2.4. However, none of the distributions
A

assessed in this work had those values of s . A n alternative route to assessment of the
A

conversion ratio in this range would be the artificial generation of distributions by

computer simulation.

The method of Matsuura and Itoh [39] should actually be applied to the true two

dimensional grain size distribution, rather than to the distribution measured on the image

analyzer. As discussed in Section 5.2, a systematic error is made when the two

dimensional grain size distribution is measured. A correction can be applied to account for

the error, thus obtaining the true two dimensional grain size distribution. The method of

Matsuura and Itoh [39] was applied to a number of cases where this correction was

applied, as reported in Table 6.3.

Corrected 2-D Estimated 3-D

Steel T . aim t No. S


A sv

C sec. grains |im fim

1080 1050 0 1142 52.83 1.14 66.14 1.60 1.25

1080 1050 120 885 58.84 1.71 74.25 1.58 1.26

1080 1050 420 629 67.67 1.63 81.85 1.50 1.21

A36 1000 600 185 87.37 1.65 111.8 1.57 1.28

A36 1150 1 193 103.4 1.75 128.1 1.63 1.24

A36 1150 60 206 149.8 1.62 189.7 1.51 1.27

Table 6.3 - 3-D distribution from corrected 2-D distribution, after Matsuura and
Itoh.
104

1.55

Standard deviation of the 2-D log normal distribution, s.

Figure 6.4 - The relationship between s and d /d


A Vm Am obtained from the application of
the method of Matsuura and Itoh [39] to measurements made on the 1080 steel,
the A36 steel and the D Q S K steel.
105

Comparison of Table 6.2 with Table 6.3 reveals that the ratio of d Vm to dAm for

corresponding specimens is almost the same, with the average ratio from Table 6.3 being

slightly higher at 1.25. For the same specimens from Table 6.3, the ratio is 1.24. This

difference can be explained by the relationship between the distribution width and the ratio

of dVm to d ,
Am as described by Figure 6.4. The corrected two dimensional distributions

are slightly wider than the measured distributions; this leads to the wider three dimensional

distribution estimated by the method of Matsuura and Itoh [39].

The difference between the d /d


Vm Am ratios from each Table is not substantial and

the conversion factor of 1.23 from all of the results in Table 6.2 is still approximately true.

This result compares well with the result of approximately 1.2, found using the method of

Takayama et al. [38]. A more detailed comparison of the method of Takayama et al. [38]

and the method of Matsuura and Itoh [39] will be made in Section 6.4.

6.3.2 The Shape of the Estimated 3-D Distribution

The distribution estimated by the method of Matsuura and Itoh [39] was generally

similar in shape to the measured distribution. Distributions which were measured as log

normal in two dimensions, remained so after the method was applied. Figure 6.5 shows

the three dimensional grain size distribution in the A36 steel held at 1000C for 10

seconds, 120 seconds and 600 seconds, as representative examples. The log normal

distribution (shown) was fit by the method of least squares, by rninimizing the value of S 2

calculated using equation 2.21. The measured two dimensional grain size distribution and

its corresponding log normal curve are also shown; the measured results are plotted as

symbols and the log normal curve plotted as dotted lines.

The two dimensional and three dimensional distributions at 10 seconds and 600

seconds are reasonably well approximated by the log normal curve. At 10 seconds, R for 2
106

the two dimensional distribution is 0.97 and 0.94 for the three dimensional distribution. At

600 seconds, R for the two dimensional distribution is 0.97 and 0.88 for the three
2

dimensional distribution. Both examples show that the estimated distribution tends to be

more scattered than the measured one. The method seems to exaggerate deviations from

log normality. For the case of abnormal grain growth at 120 seconds, the initial

distribution is not well approximated by a log normal distribution, as R is 0.85. The


2

estimated three dimensional distribution deviates even further from log normality, with R 2

equal to 0.62. This result is not unreasonable since the grain size distribution during

abnormal grain growth is not expected to be log normal; a bimodal grain structure was

actually observed.

The scatter in the estimated distribution is likely to be due to approximations

inherent in the calculation method, and inaccuracies due to the measurement statistics.

One error in the calculation method is the determination of the number of faces of a

grain. In reality, the number of faces of a grain, / , must be an integer whereas the

calculation permits / to be equal to any decimal value greater than 4. A lower limit of 4 is

chosen since it is impossible for a polyhedron to have less than 4 sides. Furthermore, in

order to smplify the calculation of the probability distribution function for each grain

shape, Matsuura and Itoh assumed that the grain shape could be approximated by planar

faced, regular polyhedra. In reality, grains have curved surfaces and are unlikely to be

entirely regular. The assumption that each three dimensional size class can be represented

by a single grain shape is also an over simplification, which could be the source of some

error.

The scatter in the estimated three dimensional distribution is a magnification of the

initial measured scatter. The two dimensional grain size distributions were determined

from a limited number of grains and the initial scatter in the measured distributions is a

result of this. It follows that if the two dimensional distribution could be more accurately
107

determined, the scatter in the three dimensional distribution could be minimized. In order

to obtain a good approximation of the two dimensional distribution, it is likely that 30 or

more size classes should be measured, with, on average, 50 to 100 grains for each size

class. Therefore, an impractical number of grains, between 1500 and 3000, would need to

be measured. Other investigators, in studies directed specifically towards investigation of

the grain size distribution, have performed measurements of the distribution with the

number of grains in this range [25,43,46,66]. In this study, approximations of the two

dimensional grain size distribution were usually obtained from 200 to 300 grains.

6.3.3 Convergence of the Distribution With Number of Size Classes

The stability of the method of Matsuura and Itoh [39] was assessed by varying the

width and number of size classes used to predict the mean equivalent volume diameter in

the 1080 steel specimen, held at 1050C for 0 seconds. Both arithmetically scaled and

geometrically scaled size classes were assessed, with a total of 1142 grains being measured

in the two dimensional distribution. Figure 6.6 shows the convergence of the mean

diameter. The arithmetically scaled and geometrically scaled size classes produced similar

results. For the number of size classes greater than 20, the method predicted a mean

equivalent volume diameter of approximately 62|im. The results predicted by the

geometrically scaled distribution for less than 20 size classes were less stable than those

obtained from the arithmetically scaled distribution; but the differences were not

significant. Both types of distributions provided mean values which were within 3% of

each other, when the number of size classes were greater than 10. The log normal

standard deviation, s , of the predicted distribution showed similar trends.


v
0.25

4.0 6.3 9.8 15.3 23.8 37.3 58.2 91.0 142.1 222.0
Grain size, EQAD, Lim
Figure 6.5 - Evolution of the 3-D grain size distribution for A36 steel held at 1000C
from the method of Matsuura and Itoh [39].
109

Figure 6.6 - Effect of the number of size classes on the method of Matsuura and Itoh [39].
110

6.4 Comparison of the Methods of Takayama et al. and Matsuura and Itoh

The mean equivalent volume diameter, d , Vm and the standard deviation, s,


v

predicted by each of the methods for the estimation of the three dimensional grain size

distribution are compared in Table 6.4 using the results from the 1080 steel held at

1050C. The results measured manually by Heyn's method and by Jeffries' method [41]

were used with the method of Takayama et al. [38]. The true, corrected two dimensional

grain size distribution was used with the method of Matsuura and Itoh [39]. The mean

values estimated by the method of Takayama et al. [38] and the method of Matsuura and

Itoh [39] are in very close agreement, with the difference being on average, 3%.

The estimation of the log normal standard deviation is a different matter. The

standard deviations predicted by each method differ by approximately 15%. By

considering the sensitivity of the method of Takayama et al. [38], and by comparison of

the estimated value of s v to the measured value of s A for each method, the standard

deviation estimated using the method of Matsuura and Itoh [39] seems to be the most

reasonable.

Hold No. True 2-D Takayama Matsuura and

time grains et al.


[38] Itoh 39]

sec. dvm s v dvm s v

|im (im |im

0 1142 52.8 1.74 64.3 1.41 66.1 1.60

120 885 58.8 1.71 71.2 1.44 74.3 1.58

420 629 67.6 1.63 83.6 1.46 81.9 1.50

Table 6.4 - Three dimensional grain size distribution predicted using the method of
Takayama et al. [38] and Matsuura and Itoh [39] for 1080 steel held at 1050C.
Ill

6.5 Validation of the Method of Matsuura and Itoh With Measured 3-D Results

As seen from the above discussion, the method of Matsuura and Itoh [39] is the

most reasonable so far reported in the literature. Matsuura et al. [25] confirmed this by

validation of the method with three dimensional measurements. The three dimensional

grain size distribution was measured in a specimen of stainless steel by complete

disintegration, and individual measurement of each grain volume. The two dimensional

grain size distribution was also measured on a section through the specimen, by the use of

image analysis equipment. Approximately 1600 grains were measured. The results were

reported in the form of eleven arithmetically scaled size classes. The measured distribution

compares well to that predicted by the method of Matsuura et al. [39], as shown in Figure

6.7.

Very little literature is available which reports the measurement of the three

dimensional grain size distribution, probably because such a task is difficult to perform.

However, three dimensional measurements have also been made by Williams and Smith

[26] using stereo radiography on an aluminum-tin alloy. Two dimensional data was

obtained on the same specimen by Aaron et al. [49], who compared a number of

estimation methods based on the assumption of spherical grain shape. The actual measured

two dimensional data is not reported, but it can be estimated by back calculation using the

results from Saltikov's method [36]. Therefore, it is possible to obtain two dimensional

and three dimensional measurements for the alummum-tin specimen; that data can be used

to further validate the method of Matsuura and Itoh [39].

The data from the aluminum tin alloy is far from ideal, since the two dimensional

results are in the form of back calculated values, and only nine size classes are given. With

only nine size classes, it is likely from Figure 6.6 that the method of Matsuura and Itoh

[39] has not converged to a stable solution. The three dimensional results are also

somewhat dubious, since the distribution was measured in three dimensions by x-ray
112

stereo radiography, with the size of each grain being determined by comparison with a

sphere of known diameter. The accuracy of such a method is not likely to be high and only

seven measured size classes were reported. Even so, an estimation of the three

dimensional grain size distribution obtained using the method of Matsuura and Itoh [39]

compares favorably to the measured results, as shown in Table 6.5.

Measured 3-D Matsuura and

[25,49] Itoh [39]

Material d ,\im
vh *v d ,\im
vh s v

Al-Sn 414 1.43 424 1.34

St. Steel 460 1.53 464 1.50

Table 6.5 - Comparison of methods for Al-Sn alloy and stainless steel.
113

0.30

0.25 H

0 200 400 600 800 1000 1200 1400


Grain size, EQVD, urn

Figure 6.7 - Comparison of the measured 3-D grain size distribution with
the estimated distribution, for stainless steel, using the method of
Matsuura and Itoh [39], from Matsuura et al. [25].
CHAPTER 7

SUMMARY AND CONCLUSIONS

The objectives of this work were firstly, to quantify the kinetics of austenite grain

growth in two plain carbon steels, and secondly, to determine an appropriate method for

the estimation of the three dimensional grain size distribution.

For the purpose of investigating the kinetics of grain growth, the grain size was

measured in specimens of A36 and DQSK steel, heated to temperatures between 950C

and 1150C, and held for periods up to 900 seconds. Linear intercept, l , and mean area,
m

A , were measured using the Heyn and Jeffries manual methods and by the use of image
m

analysis equipment, as describe by A S T M standards [40-42]. Mean equivalent area

diameter, d , was calculated from mean area, A . Standard statistical methods were used
Am m

to determine 95% confidence intervals of approximately 10% for the mean linear

intercept, l , and the mean equivalent area diameter, d ,


m Am determined using both types of

measurement methods.

Systematic errors in the determination of the linear intercept length and the mean

equivalent area diameter were identified and discussed. The true linear intercept is closest

to that determined manually by Heyn's method [41]. The true mean area can be

determined by the application of an iterative correction to measurements determined by

image analysis [40]. The number of grains which touch the edge of the image being

measured must be known. Mean area determined by Jeffries' method [41] is a good first

approximation of the true mean area.

The best type of measurement to perform is the measurement of the two

dimensional grain size distribution by the use of the image analyzer, since the true two

dimensional grain size distribution can then be determined from this result. Such a

114
115

measurement is not difficult to perform and the necessary image analysis equipment is

becoming commonly available, as a standard metallographic laboratory tool.

The kinetics of austenite grain growth were observed in the A36 steel and in the

D Q S K steel. In both materials abnormal growth and a limiting grain size were observed.

Growth kinetics in the A36 steel were influenced by heating rate and by solution treatment

of the specimens. The kinetics of grain growth were adequately interpreted with respect to

the influence of second phase particles, which were concluded to be aluminum nitride,

from the work of Gladman [28]. The phenomenon of abnormal grain growth and a hndting

grain size were also explained by the presence of aluminum nitride particles.

The kinetics of grain growth observed under laboratory conditions are expected to

be quite different from those observed in a hot strip mill, because the thermal and

mechanical history obtained during hot rolling is difficult to reproduce. Both the

dispersion of second phase particles and the width of the grain size distribution influence

the kinetics of grain growth and both are influenced by the thermal and mechanical

processing history.

The power law was fit to results of mean equivalent area diameter obtained for the

A36 and the DQSK steels. The fitting constants used were found to be dependent on the

thermal history. Because the exponent, m was greater than two, no fundamental basis for

the power law equation was found. It must be concluded that the power law is an

empirical relationship, which should not be extrapolated to model grain growth for

conditions outside of those for which the results have been obtained.

The statistical model described by Abbruzzese and Lticke [20,21] was applied to the

results from the A36 and the DQSK steels by Militzer et al. [61,62]. The pmiing

parameter, P, was used to fit the model and the parameter could be interpreted from a

theoretical perspective [61,62]. The pirating parameter, P, was also estimated for hot strip
116

rolling conditions [61,62]. The results obtained in this work were therefore used to

successfully validate the statistical model.

Thus, as anticipated in the first objective, the kinetics of austenite grain growth have

been quantified. The results have been used to validate a statistical model, which can be

used to model the kinetics of austenite grain growth during hot strip rolling.

The evolution of the measured two dimensional grain size distribution was described

for a case of normal growth and abnormal growth in the A36 steel specimens heated at

5C/second. Under conditions of normal grain growth at 1150C, the distribution width

and shape were seen to change little over time, with the distribution being closely

approximated by a log normal curve. The mean of the distribution was observed to

increase by scaling.

At 1000C and after 10 seconds, the distribution shape was also apparently log

normal. Under conditions of abnormal growth at 120 seconds, the distribution was seen to

widen dramatically, and then at 600 seconds, to return to a width and shape similar to that

initially observed.

Methods for estimation of the three dimensional grain size distribution were

compared by applying each to results obtained from the 1080, the A36 and the D Q S K

steels. It was determined that the methods of Saltikov [36] and Huang and Form [37],

based on the assumption that the grain shape is spherical, were unacceptable; both

methods predicted a negative number of grains for the smallest size classes observed.

The method of Takayama et al. [38] was found to be sensitive to the ratio of linear

intercept to the square root of the mean grain area. When the method's sensitivity was

dampened by the use of a mean value for the log normal standard deviation, s , reasonable
v

estimations were obtained for the equivalent volume diameter, d . Vm The method was

found to be unsuitable for determining the width of the distribution and could not be

applied to distributions which were not log normal.


117

The method of Matsuura and Itoh [39] was found to be appropriate for calculating

the mean equivalent volume diameter and the width of the distribution. The true, corrected

two dimensional grain size distribution must be used as input. Some scatter was observed

in the histograms predicted by the method, but the distribution shape generally appeared to

be reasonable. The scatter was a results of the statistics of the measurements, which were

performed using a limited number of grains, due practical considerations. Some scatter

could also attributed to inaccuracies in the simplifying assumptions, inherent in the

procedure. The method was applied to measured distributions where the shape of the

distribution was log normal, and also to distributions where the shape was not log normal.

It was found that a stable solution was predicted when greater than ten size classes were

employed. The method was originally validated using measured three dimensional results

by Matsuura et al. [25]. A second example of a measured three dimensional distribution

was reported in the literature [49] and this was also used to validate the method.

The methods of Takayama et al. [38] and Matsuura and Itoh [39] produced values

for the mean equivalent volume diameter, d , Vm which were almost equal. Differences

between the two results could be accounted for by the statistics of the measurements. For

cases where normal grain growth was observed, both methods confirm that the

approximate relationship,

dvm =
L2dAm = l.5lm

can be used to estimate the mean equivalent volume diameter, d .


Vm

It must be concluded that the method of Matsuura and Itoh [39] was the most

suitable of those compared, for the estimation of the three dimensional grain size

distribution. Therefore, the second objective, to determine an appropriate method for the

estimation of the three dimensional grain size distribution, was also met.
BIBLIOGRAPHY

[1] H . Yada, Prediction of Microstructural Changes and Mechanical Properties in Hot

Strip Rolling, Proceedings of the International Symposium on Accelerated Cooling

of Steel, Winnipeg, Canada, Edited by G.E. Ruddle and A.F. Crawley, Aug. 1987,

Permagon Press, ppl05-l 19.

[2] N . Komatsubara, K . Kunishige, S. Okaguchi, T. Hashimoto, K . Ohshima and I.

Tamura, Computer Modelling for the Prediction and Control of Mechanical

Properties in Plate and Sheet Steel Production, The Sumitomo Search, No.44, Dec.

1990, ppl59-168.

[3] C. Devadas, I.V. Samarasekera and E.B. Hawbolt, The Thermal and Metallurgical

State of Steel Strip during Hot Rolling: Part III. Microstructural Evolution, Met.

Trans. A, Vol. 22A, Feb. 1991, pp335-349.

[4] P.D. Hodgson and R . K . Gibbs, A Mathematical Model to Predict the Mechanical

Properties of Hot Rolled C-Mn and Microalloyed Steels, ISIJ International, Vol. 32,

1992, No. 12, ppl329-1338.

[5] C M . Sellars and J.A. Whiteman, Recrystalfization and Grain Growth in Hot Rolling,

Met. Sci., Vol. 13, 1979, pp 187-194.

[6] J.E. Burke and D. Turnbull, Recrystallization and Grain Growth, Progress in Metal

Physics 3, Permagon, 1952, pp220-292.

[7] P. Feltham, Grain Growth in Metals, Acta.Metall., Vol.5, Feb. 1957, pp97-105.

[8] M . Hillert, On the Theory of Normal and Abnormal Grain Growth, Acta MetalL, V o l .

13,Mar.l965,pp227-238.

[9] J.C. Fisher and R. Fullman, Private communcation to Burke and Turnbull, [6.].

[10] M . P . Anderson, G.J. Srolovitz, G.S. Grest and P.S. Sahni, Computer Simulation of

Grain Growth, Acta MetalL, Vol. 32, No. 5, 1984, pp783-791.


119

[11] P. Gordon and T.A. El-Bassyouni, The Effect of Purity on Grain Growth in

Aluminum, Trans.A.I.M.E.,233,1965, pp391-396.

[12] H . Hu, Grain Growth in Zone Refined Iron, Can.Metall.Q., Vol.13, 1974, pp275-

286.

[13] G.F. Boiling and W.C. Winegard, Grain Growth in Zone Refined Lead, Acta

Metall.,Vol.6, 1958, pp283-287.

[14] J.P. Drolet and A . Galibois, The Impurity Drag Effect on Grain Growth, Acta

Metall., Vol.16, 1968,ppl387-1399.

[15] E.L. Holmes, and W.C. Winegard, Grain Growth in Zone Refined Tin, Acta Metall.,

Vol.7, 1959, pp411-414.

[16] H . V . Atkinson, Theories of Normal Grain Growth in Pure Single Phase Systems,

Acta Metall., Vol. 36, No. 3, 1988, pp469-491.

[17] T. Senuma and H . Yada, Microstructural Evolution of Plain Carbon Steels in

Multiple Hot Working, Proc. 7th Risty Int.Symp.on Metallurgy and Materials

Science, Denmark, 1986, pp547-552.

[18] P.C. Campbell, E.B. Hawbolt and J.K Brimacombe, Microstructural Engineering

Applied to the Controlled Cooling of Steel Wire Rod: Part II. Microstructural

Evolution and Mechanical Properties Correlations, Met.TransA, V o l . 22A, 1991,

pp2779-2790.

[19] V . Y . Novikov, Computer Simulation of Normal Grain Growth, Acta Metall, V o l .

26, 1978, ppl739-1744.

[20] G . Abbruzzese and K . Lticke, A Theory of Texture Controlled Grain Growth - I.

Derivation and General Discussion of the Model, Acta Metall, V o l . 34, 1986,

pp905-914.

[21] G . Abbruzzese and K . Lticke, Theory of Grain Growth in the Presence of Second

Phase Particles, Mat.Sci.Forum, Vol. 94-96, 1992, pp597-604.


[22] C.S. Pande, On a Stochastic Theory of Grain Growth, Acta MetalL, Vol. 35, 1987,

pp2671-2678.

[23] S.K. Kurtz and F . M . A . Carpay, Microstructure and Normal Grain Growth in Metals

and Ceramics. Part 1. Theory, JAppl.Phys., Vol. 51(11), 1980, pp5725-5744.

[24] F.N. Rhines and K.R. Craig, Mechanism of Steady State Grain Growth in Aluminum,

Met.Trans., Vol. 5, Feb.1974, pp413-425.

[25] K . Matsuura, Y . Itoh, M . Kudoh, T. Ohmi and K. Ishii, Three-dimensional Grain Size

Distribution in SUS304 Stainless Steel, ISIJInt., Vol. 34, No. 2, 1994, ppl86-190.

[26] W . M . Williams and C.S. Smith, A Study of Grain Shape in an Aluminum Alloy And

Other Applications of Stereoscopic Microradiography, JOM, July, 1952, pp755-765.

[27] C . H . Desch, The Solidification of Metals from the Liquid State, Journ.Inst.Metals,

Vol.22, 1919, pp241-263.

[28] T. Gladman, The Theory and Inhibition of Abnormal Grain Growth in Steels, JOM,

Sep. 1992, pp21-24.

[29] C.H. Worner and P . M . Hazzledine, Grain Growth Stagnation by Inclusions or Pores,

JOM, Sep. 1992, ppl6-20.

[30] C. Zener as quoted by C.S. Smith, Grains, Phases and Interfaces: A n Interpretation of

Microstructure, TransAIME, Vol.175, 1948, ppl5.

[31] T. Gladman, On the Theory of the Effect of Precipitate Particles on Grain Growth in

Metals, Proc.Roy.Soc.London,A, Vol.294, 1966, pp298-309.

[32] T. Gladman and F.B. Pickering, Grain-Coarsening of Austenite, Journ.Iron.Steel.

Inst., June 1967, pp653-664.

[33] M D . M . A h Bepari, Effects of Second Phase Particles on Coarsening of Austenite in

0.15Pct Carbon Steels, Met.TransA, Vol.20A, 1989, ppl3-16.

[34] P.T. Mazarre, S.W. Thompson and G. Krauss, Microalloying and Austenite Grain

Size Control: Implications for Direct-Cooled Forgings, Proceedings.Int.Conf. on


Processing, Microstructure, and Properties, I.S.S. of the A.I.M.E., Pittsburg, June,

1991,pp497-509.

[35] H.E. Exner, Analysis of Grain and Particle size Distributions in Metallic Materials,

Int.Met.Rev., Vol.17, 1972, pp25-41.

[36] S.A. Saltikov, The Determination of the Size Distribution of Particles in an Opaque

Material from a Measurement of the Size Distribution of Their Sections, Stereology,

Proc.2nd.Int.Cong. for Stereology, Chicago, Springer-Verlag, Edit. H . Elias, 1967,

ppl63-169.

[37] K.P. Huang and W. Form, A Simplified Method to Reconstruct a Spatial Grain Size

Distribution from Planar Observations Part I: Basic Considerations, Prakt.Met., 27,

1990, pp332-340.

[38] Y . Takayama, N . Furushiro, T. Tozawa, H . Kato and S. Hori, A Significant Method

for the Estimation of the Grain Size of Polyciystalline Materials, Mat.Trans., JIM,

Vol. 32, No. 3, 1991, pp214-221.

[39] K . Matsuura and Y . Itoh, Estimation of Three-dimensional Grain Size Distribution in

Polycrystalline Material,Mat.Trans., JIM, Vol. 32, N o . l l , 1991, ppl042-1047.

[40] A S T M Standard Designation: E1382-91, 1991 Book of ASTM Standards, Vol.3.01,

A S T M , Philadelphia, PA, 1991, pp972-987.

[41] A S T M Standard Designation: E l 12-88, 1994 Book of ASTM Standards, Vol.3.01,

A S T M , Philadelphia, PA, 1994, pp227-251.

[42] A S T M Standard Designation: E l 181-87, 1994 Book of ASTM Standards, Vol.3.01,

A S T M , Philadelphia, P A , 1991, pp839-852.

[43] F . N . Rhines and B.R. Patterson, Effect of the Degree of Prior Cold Work on the

Grain Volume Distribution and the Rate of Grain Growth of Recrystalhzed

Aluminum, Met.Trans., V o l . l 3 A , 1982, pp985-993.


[44] G. Bockstiegel, A Simple Formula for the Calculation of Spatial Size Distributions

From Data Found by Lineal Analysis, Stereology, Proc.2nd.Int.Cong. for

Stereology, Chicago, Springer-Verlag, Edit. H . Elias, 1967, ppl93-194.

[45] R . G . Cortes, A.O. Sepulveda and W.O. Busch, On the Determination of Grain Size

Distributions from Intercept Distributions, J.Mat.Sci., Vol.20, 1985, pp2997-3002.

[46] K.P. Huang and W. Form, A Simplified Method to Reconstruct a Spatial Grain Size

Distribution from Planar Observations Part II: Application to the Recrystallization

Kinetics of Cold Drawn OF-Cu Bars, Prakt.Met., 27, 1990, pp341-350.

[47] E. Scheil, Z.anorg.Chem., Vol.201, 1931, pp259.

[48] H.A. Schwartz, Metals and Alloys, Vol.5, 1934, ppl39.

[49] H.B. Aaron, R.D. Smith and E.E. Underwood, Spatial Grain-Size Distribution from

Two-Dimensional Measurements, Proc.IstJnt.Cong.Stereology, Vienna, 1963,

paper 16.

[50] D. Shi, A Probabilistic Approach to Estimate Particle Size Distribution Without

Shape Assumption. Part 1: Theory and Testing on Spherical Particles, Mat. Charact.,

Vol.27, 1991, pp35-44.

[51] N . Furushiro, Y . Takayama and S. Hori, Estimation of the Mean Grain Diameter of

Polycrystalhne Materials, Mat.Trans., JIM, Vol.30, N o . l , 1989, pp27-32.

[52] Y . Takayama, T. Tozawa and H . Kato, Linear Intercept Length Distribution in a

Grain Structure Model with Diameter Distribution of Log-Normal Form, Trans.JIM,

Vol.28, No.8,1987, pp631-643.

[53] R.T. DeHoff, The Estimation of Particle-Size Distributions from Simple Counting

Measurements Made on Random Plane Sections, Trans.Met.Soc.AIME, Vol.233,

1965, pp25-29.

[54] S.K. Kurtz and F.M.A. Carpay, Microstructure and Normal Grain Growth in Metals

and Ceramics. Part 2. Experiment, JAppl.Phys., Vol. 51(11), 1980, pp5745-5754.


[55] E . B . Hawbolt and B. Chau, University of British Columbia, B.C., Canada,

Unpublished Research, 1989.

[56] Metals Handbook Desk Edition, Edited by H.E. Boyer and T.L. Gall, American

Society for Metals, Ohio, 1985, pp28.3.

[57] L . E . Samuels, Optical Microscopy of Carbon Steels, American Society for Metals,

Ohio, 1980.

[58] Materials Processing Databook, '83, Metal Progress, 1983, pp91-95.

[59] G.F. Vander Voort, Metallography Principles and Practice, McGraw-Hill, N Y ,

1984.

[60] O.O. Miller, Influence of Austenitizing Time and Temperature on Austenite Grain

Size of Steel, Trans.ASM, Vol.43, 1951, pp260-289.

[61] M . Militzer, A . Giumelh, E.B. Hawbolt and T.R. Meadowcroft, University of British

Columbia, Internal Report, Unpublished, 1994.

[62] M . Militzer, A . Giumelli, E.B. Hawbolt and T.R. Meadowcroft, Austenite and Ferrite

Grain Size Evolution in Plain Carbon Steel, 36th Metal Working and Steel

Processing Conference, Baltimore, 1994.

[63] N . A . Gjostein, H . A . Domain, H.I. Aaronson and E. Eichen, Acta metall, Vol.14,

1966, ppl673.

[64] H.J. Frost and M . F . Ashby, Deformation-Mechanism Maps, Permagon Press,

Oxford, U K , 1982.

[65] R . V . Hogg and J. Ledolter, Engineering Statistics, Macmillan, 1987.

[66] D . A . Aboav, The Arrangement of Grains in a Polycrystal, Metallography, Vol.3,

1970, pp383-390.
APPENDIX

HR. No. 95% A 95%


Temp Spec. time d-Am
m
n

cond. a grns conf conf

c sec sec |im (%) |im (im 2

(%)

900 As rec. 120 5 506 9.17 17.5 11.87 110.59 17.5

950 As rec. 10 5 733 9.06 4.7 11.38 101.80 4.1

950 As rec. 120 5 492 11.05 8.2 13.90 151.66 9.4

950 As rec. 600 5 491 11.27 3.3 13.91 151.97 4.5

1000 As rec. 10 5 324 11.58 7.6 14.48 164.65 12.3

1000 As rec. 60 5 309 12.81 5.4 14.83 172.65 12.5

1000 As rec. 60 5 61 63.74 87.95 6075.9

1000 As rec. 120 5 286 14.06 12.6 17.23 233.16 23.8

1000 As rec. 120 5 281 71.99 8.88 93.13 6812.4 10.2

1000 As rec. 600 5 229 78.84 13.9 92.28 6687.4 16.3

1000 As rec. 900 5 264 81.93 8.5 96.09 7251.1 13.7

Table A.1 - A36 austenite grain size results obtained by standard Gleeble thermal
treatment and manual measurements, using Heyn's and Jeffries' procedures [41].

124
125

H.R. No. 95% A 95%


Temp Spec. time d-Am
"-m
cond. C/ gms conf conf

c sec sec |lm (%) |im (im 2


(%)

1050 As rec. 1 5 143 15.08 4.6 18.87 279.79 1.65

1050 As rec. 10 5 48 21.14 27.96 614.15

1050 As rec. 10 5 350 64.79 7.2 83.45 5469.4 9.76

1050 As rec. 60 5 14 28.70 44.99 1590

1050 As rec. 60 5 284 76.55 9.3 92.64 6740.4 9.93

1050 As rec. 120 5 265 84.98 9.3 105.1 8668.4 8.30

1050 As rec. 300 5 316 86.58 12.1 110.0 9499.3 11.6

1050 As rec. 750 5 291 97.79 7.9 114.6 10315 13.1

1100 As rec. 1 5 307 73.39 12.7 87.67 6036.3 11.0

1100 As rec. 10 5 328 67.15 8.7 86.20 5836.2 22.3

1100 As rec. 30 5 289 88.68 3.0 115.0 10387 4.0

1100 As rec. 60 5 229 103.6 9.4 129.2 13108 12.9

1100 As rec. 120 5 265 106.4 9.3 131.6 13593 8.3

1100 As rec. 300 5 209 112.6 3.0 135.2 14363 6.7

1100 As rec. 600 5 196 113.5 2.3 139.6 15315 10.1

Table A.l(cont) - A36 austenite grain size results obtained by standard Gleeble thermal
treatment and manual measurements, using Heyn's and Jeffries' procedures [41].
126

Temp Spec. time HR. No. 95% A 95%


d
Am
m
n

cond. c/ grns conf conf

c sec sec |im (%) |im [im 2

(%)

1150 As rec. 1 5 237 77.7 11.3 99.8 7819.1 18.5

1150 As rec. 10 5 318 105.2 7.2 135.3 14366 15.9

1150 As rec. 30 5 293 108.7 8.7 140.9 15592 13.9

1150 As rec. 60 5 281 116.0 4.4 144.0 16287 7.8

1150 As rec. 120 5 258 150.1 9.0 191.6 28821 6.6

1150 As rec. 300 5 295 167.3 13.7 196.1 30189 7.0

1150 As rec. 450 5 222 169.2 5.1 206.6 33506 5.9

1200 As rec. 300 5 172 183.9 10.3 235.8 43685 4.9

1200 As rec. 600 5 253 206.4 6.7 244.4 46916 11.9

950 As rec. 120 50 523 30.5 732.2

950 Sol.trt. 120 50 219 47.2 1748.5

1050 As rec. 120 50 309 54.8 2358

1050 Sol.trt. 120 50 177 72.4 4116.8

1150 As rec. 120 50 132 189.7 28263

1150 SoLtrt. 120 50 72 257.0 51815

Table A.l(cont.) - A36 austenite grain size results obtained by standard Gleeble thermal
treatment and manual measurements, using Heyn's and Jeffries' procedures [41].
127

Temp time H.R. No. Sy 95%

c/ conf

c sec sec grns. \\m |im 2


(%)

1000 10 5 250 14.4 1.59 162.63 21.4

1000 120 5 47.8 2.40 1793.8

1000 600 5 185 81.5 1.63 5217.1 17.5

1150 1 5 193 89.7 1.71 6312.2 23.0

1150 60 5 206 134.6 1.58 14234 8.6

1150 450 5 115 187.0 1.62 27471 9.3

Table A.2 - A36 austenite grain size results obtained by standard Gleeble thermal
treatment and measurement using the image analyzer [40].

Quench Spec. cool No. 95% d-Am A 95%


m n

Temp, cond. time, grns conf conf

c sec |im (%) |im iim 2


(%)

1120 AR l 470 38.7 6.0 49.7 1938 in


1110 AR 5 285 49.7 6.0 63.8 3196 1A

1100 AR 10 269 50.7 8.3 65.7 3386 5.5

1080 AR 20 263 51.2 5.6 66.4 3463 7.4

1030 AR 45 243 55.0 4.9 69.1 3748 9.95

1000 AR 60 228 55.3 7.1 71.3 3995 4.6

Table A.3 - A36 austenite grain size results obtained by continuous cooling on the
Gleeble, and measurement by Heyn's and Jeffries' procedures [41].
128

Temp, Spec. time, No. 95% d-Am 95%

cond. gms conf conf

c sec (im (%) (lm (im 2

(%)
1050 AR 1 371 46.4 9.2 55.9 2455.1 15.8

1050 AR 10 251 47.0 2.8 60.8 2903.1 9.9

1050 AR 30 232 55.0 3.8 63.2 3140.8 5.9

1050 AR 120 406 62.5 8.0 76.2 4564.4 12.7

1050 AR 750 330 71.9 14.0 84.6 5615.6 16.4

1100 AR 1 427 38.9 10.3 46.6 1706.5 23.0

1100 AR 10 222 59.4 9.6 72.3 4102.9 12.1

1100 AR 30 135 68.0 5.3 82.9 5397.5 4.5

1100 AR 120 226 73.0 10.6 91.4 6559.8 14.2

1100 AR 600 251 78.9 10.6 97.0 7383.0 18.3

Table A.4 - A36 austenite grain size results obtained by stepped Gleeble thermal
treatment and manual measurement by Heyn's and Jeffries' procedures [41].
129

Temp time No. d-Am A 95%

grns Jeff. Im.An. Im.An. conf

c sec |im |Xm |im 2


(%)

1000 0 352 29.0 27.2 580.7 9.8

1000 50 395 30.6 29.7 691.3 6.8

1000 122 356 32.3 31.3 767.9 1.1

1000 245 266 33.4 32.5 827.9 7.7

1000 412 256 34.0 33.3 868.3 1.4

1050 0 397 30.6 28.9 654.2 12.0

1050 61 311 30.9 29.4 679.9 20.3

1050 118 276 32.8 31.1 758.8 15.8

1050 236 298 35.3 33.5 879.6 6.8

1050 420 226 36.2 33.8 897.0 9.1

1100 0 261 37.4 34.8 952.8 10.2

1100 68 299 45.2 42.8 1436.8 10.2

1100 124 249 66.8 64.0 3219.8

1100 244 295 79.0 72.8 4160.4

1100 424 368 101.9 87.3 5101.7 19.8

1150 0 246 124.6 112.4 9917.5 9.9

1150 55 271 148.0 135.8 14486 10.6

1150 113 218 165.0 145.5 16618 11.6

1150 227 335 168.0 152.8 18342 11.6

1150 410 298 178.1 163.1 20894 9.6

A.5 - DQSK austenite grain size results obtained by tube furnace thermal treatment
and measurement using Jeffries' procedure [41] and the image analyzer [40].
130

Temp time No.

gms Jeff. Jeff.

c sec (im (im 2

1050 60 173 37.5 1107

1100 60 462 78.3 4813

1100 120 183 113.5 10126

1150 60 197 155.3 18937

Table A.6 - D Q S K austenite grain size results from standard Gleeble thermal
treatment and measurement by Jeffries' procedure [41].

Temp time No. 95% d


Am 95% d
Am 95%

grns conf Jeff. Jeff. conf Im.An Im.An. conf

c sec |im (%) fim |im 2


(%) |im |im 2
(%)

1050 0 1313 43.8 2.2 53.6 2258.2 6.6 50.3 1953.7 7.5

1050 120 1055 49.3 4.0 59.8 2810.4 6.2 55.3 2405.6 7.4

1050 420 756 58.5 5.8 70.7 3922.0 9.2 64.7 3273.9 8.2

Table A.7 - 1080 steel austenite grain size results from furnace thermal treatment
applied in previous work [55], and measurement by Heyn's and Jeffries' procedure
[41], and by the image analyzer [40].

S-ar putea să vă placă și