Você está na página 1de 408

H E A T T R A N S F E R A N D STRESS G E N E R A T I O N D U R I N G

F O R C E D C O N V E C T I V E Q U E N C H I N G OF S T E E L B A R S
By
Jose Bernardo Hernandez-Morales
B. Sc. (Metallurgical Eng.) University of Mexico, Mexico City, 1983
M . A . Sc. (Metals and Materials Eng.) University of British Columbia, 1988

A THESIS SUBMITTED IN PARTIAL 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

D O C T O R OF P H I L O S O P H Y

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 METALS AND MATERIALS ENGINEERING

W E A C C E P T THIS THESIS AS CONFORMING

T O T H E REQUIRED STANDARD

T H E UNIVERSITY OF BRITISH COLUMBIA

MAY 1996

JOSE BERNARDO HERNANDEZ-MORALES ; 1996


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 Metals and Materials Engineering


The University of British Columbia
2075 Wesbrook Place
Vancouver, Canada

V 6 T 1W5

Date:
Abstract

A n industrial heat treatment based on forced convective quenching of alloyed eutectoid


steel bars has been studied within the framework of microstructural engineering. The pro-
prietary process is used to produce grinding rods with improved abrasion resistance and
toughness. Mathematical models of heat transfer, microstructural evolution and stress
generation have been applied to predict the final microstructure and residual stress dis-
tribution in 38.1 mm-dia. bars, quenched in a laboratory facility. The models are based
on the finite-element method (FEM) and assume axisymmetric geometry. The principle
of additivity was invoked to predict the evolution of the microstructure under continuous
cooling conditions. The kinetics of diffusional and martensitic phase transformations were
characterized using the J M A K and Koistinen-Marburger equations, respectively. The ki-
netic parameters were determined through continuous cooling and isothermal tests using
the Gleeble 1500 thermomechanical simulator. The existing elasto-plastic stress model
incorporates thermal- and transformation-related strains. The mechanical properties
were obtained from the literature. The effect of stresses on transformation kinetics was
not considered; thus, the thermal-microstructural and mechanical models were effectively
uncoupled.
The heat-transfer boundary condition has been characterized by acquiring the thermal
response in instrumented interstitial-free (IF) steel bars quenched from 1000 C in flowing
water. Three water velocities and water temperatures were investigated. Tests with clean
and oxidized bars were also conducted. The surface heat flux, as a function of surface
temperature, was estimated by solving the inverse heat conduction problem (IHCP).
The solution algorithm was based on the sequential function specification technique.
The water temperature was found to have a significant effect on the boiling curves. For

n
the highest water temperature, 75 C, a film boiling stage was identified, while only
transition, nucleate and convective boiling were observed when using water at 50 and
25 C . No significant differences in heat extraction were observed between clean and
oxidized quenched bars, within the parameter range investigated.

The residual stress distribution in selected forced convective quenched IF, 1045 carbon
and alloyed steel bars has been determined by means of neutron diffraction. The axial
and circumferential residual stresses in the IF and 1045 carbon steel quenched bars were
compressive at the surface and tensile at the centre, while the radial component was
always tensile. In contrast, the alloyed eutectoid steel bar showed compressive axial,
circumferential and radial residual stresses at the centre. The final microstructural and
hardness distributions have also been determined in the three bars. The IF steel bar
transformed completely to ferrite, while the alloyed eutectoid steel specimen showed an
essentially martensitic structure. On the other hand, quenching the 1045 carbon steel
bar resulted in an outer ring of martensite and a mixture of diffusional and martenistic
products in the core.
Comparisons between measured and model-predicted thermal response, final mi-
crostructure and residual stress distribution have been made. Fair agreement between
measured and predicted values was observed. It was found that the position of the 'nose'
of the continuous cooling diagram, when a mixture of diffusional and martensitic prod-
ucts was produced, has a significant influence on the predicted final microstructure and,
therefore, on the predicted residual stress distribution. The difference in the measured
residual stress distributions obtained in the alloyed eutectoid steel specimen, when com-
pared with those found in the IF and 1045 carbon steel quenched bars, has been explained
based on the sequence of transformations that took place during the quench. The quench
of a bar under industrial conditions was also simulated.

in
Table of Contents

Abstract ii

Table of Contents iv

List of Tables ix

List of Figures xii

List of Symbols xxxiii

Acknowledgements xxxix

1 Introduction 1

1.1 Microstructural Engineering 1


1.2 Heat Treatment of Grinding Media 2
1.3 Statement of the Problem 3

2 Literature Review 6
2.1 Heat Treatment Design 6
2.1.1 Empirical Approaches 6
2.1.2 Characterization of Quench-Bath Quality 7

2.1.3 Process Modeling 14


2.1.4 Computerized Information Systems 15
2.2 Heat Transfer in Forced Convective Quenching 16

2.2.1 General 16
2.2.2 Surface Heat Flux Characterization 17

iv
2.2.3 Boiling Heat Transfer 19
2.2.4 Forced Convective Quenching 23
2.3 Microstructural Evolution 25
2.3.1 General 26
2.3.2 Isothermal Phase Transformation Kinetics 26
2.3.3 Continuous Cooling Phase Transformation Kinetics 32
2.3.4 Fe-C-X Phase Boundaries 37
2.3.5 Effect of Stresses on the Kinetics of Phase Transformations . . . . 38
2.4 Residual Stresses in Heat Treatments 40
2.4.1 General 40
2.4.2 Residual Stress Measurement Techniques 41
2.5 Mathematical Models of Microstructural Evolution and Stress Generation
in Heat Treating Operations 43
2.5.1 Modeling of Heat Transfer and Microstructural Evolution 43
2.5.2 Modeling of Stress Generation 45

3 Scope and Objectives 61


3.1 Scope of the Research Programme 61

3.2 Objectives of the Research Programme 63

4 Heat Transfer Model 64


4.1 Governing Equation 64
4.1.1 Rate of Heat Evolved 65
4.2 Finite-Element Formulation 67
4.2.1 Finite-Element Equations 68
4.2.2 Solution Algorithm 70
4.3 Verification of Mathematical Model of Heat Flow 72
4.3.1 Infinite Solid Cylinder : q = 0 73

v
4.3.2 Infinite Solid Cylinder : q = f(T) 75
4.3.3 Infinite Solid Cylinder : A i r Cooling of Eutectoid Steel 76

4.4 Summary 76

5 The Inverse Heat Conduction Problem 88


5.1 Solution Algorithm 89
5.2 Modification for Cylindrical Coordinates 92
5.3 Application to Controlled A i r Cooling of Rods 95
5.4 Application to Forced Convective Boiling 95
5.5 Sensitivity coefficients and experimental design 98

6 Stress Model 112


6.1 Stress-Strain Relations 112
6.1.1 Elastic Stress-Strain Relations 114
6.1.2 Plastic Stress-Strain Relations 115
6.1.3 Finite-Element Formulation 120
6.2 Verification of Mathematical Model of Stress Generation . . . . . . . . . 125
6.2.1 Infinite Solid Cylinder : Elastic Thermal Stresses 125
6.2.2 Infinite Solid Cylinder : Elastoplastic Thermal Stresses 127

6.3 Summary 128

7 Laboratory Experiments : Quenching Tests 134


7.1 Quenching Apparatus 135
7.2 Procedure 141
7.3 Metallographic Characterization 143

8 Laboratory Experiments : Transformation Kinetics 161


8.1 Material and Sample Preparation 161
8.2 Procedure 162

vi
8.2.1 Continuous Cooling Tests 162
8.2.2 Isothermal Tests 162

8.2.3 Temperature Gradient in a Gleeble Specimen 163


8.3 Microstructural Characterization 164

9 Laboratory Results and Discussion : Quenching Tests 174


9.1 Measured Temperature Response 174
9.2 Estimated Surface Heat Flux 178
9.3 Metallographic Characterization 185

10 Laboratory Results and Discussion : Transformation Kinetics 215


10.1 Phase Boundaries 215
10.2 Continuous Cooling Tests 216
10.2.1 Alloyed Eutectoid Steels 216
10.2.2 1045 Carbon Steel 218
10.3 Isothermal Tests 219
10.4 Prior-Austenite Grain Size 221
10.5 Ferrite Fraction 222

11 Residual Stress Measurement 239


11.1 Experimental Procedure 239
11.1.1 Campaign 1 . . 239
11.1.2 Campaign 2 241

11.2 Results 2 4 2

12 Mathematical Analysis of Forced Convective Quenching 260


12.1 Verification of the Inverse Analysis 260

12.2 Temperature Response and Microstructural Evolution 264


12.2.1 IF Steel 264

vii
12.2.2 1045 Carbon Steel 266
12.2.3 Alloyed Eutectoid Steel 267
12.2.4 Sensitivity Analysis 269
12.3 Stress Generation 271
12.3.1 Residual Stresses 272
12.3.2 Transient Stresses 275

12.4 Application to Industrial Conditions 277

13 Summary and Conclusions 313

Bibliography 318

A Thermo-Microstructural Model : Finite Element Equations 338

A.l Isoparametric Elements and Numerical Integration 340

B Thermo-Microstructural Model : Semi-Analytical Solution 345

C Stress Model : Finite Element Equations 347

C.l Isoparametric Elements and Numerical Integration 351

D Stress Model : Nonmechanical Strains 354

E Recalescence During Quenching of IF Steel Bars 356

F Isothermal Tests : Data Reduction 359


F.l Linear Regression 359
F.2 Non-Linear Regression 360

vm
List of Tables

2.1 Classification of techniques for measuring residual stresses (modified from


[130]) 53

4.1 Decision tree to determine the sequence of calculations during phase


transformations in a eutectoid steel 77
4.2 Input data used for the comparison of finite-element and analytical so-
lutions under Newtonian cooling conditions 78
4.3 Input data used for the comparison of finite-element and analytical so-
lutions under non-Newtonian cooling conditions 78
4.4 Input data adopted for the comparison of finite-element and finite-difference
simulations of air cooling of eutectoid steel rod 79

5.1 Parameters used for Case 1 : constant surface heat flux 100
5.2 Parameters used for Case 2 : constant heat-transfer coefficient 100
5.3 Parameters used for the simulation of a forced convective quenching ex-
periment 101

5.4 IHCP algorithm : summary of sensitivity runs 101

6.1 Input data used for the comparison of finite-element and analytical so-
lutions for elastic stresses generated in an infinite solid cylinder by a
temperature gradient 129

7.1 Chemical composition of the IF steel used in the forced convective quench-
ing experiments (in weight percent) 144

ix
7.2 Chemical composition of the alloyed steels used in the forced convective
quenching experiments (in weight percent) 144

7.3 Chemical composition of the 1045 steel used in the forced convective
quenching experiments (in weight percent) 145
7.4 Typical operational plant data 145
7.5 Test matrix used for the boiling heat transfer experiments 146
7.6 Quenching conditions for specimens produced for residual stress mea-
surements and metallographic analysis 146
7.7 Quenching conditions to study the effect of surface oxidation 147

8.1 Summary of continuous cooling tests for the alloyed eutectoid steel A . . 166
8.2 Summary of continuous cooling tests for the alloyed eutectoid steel B . . 166
8.3 Summary of continuous cooling tests for the alloyed eutectoid steel C. . 167
8.4 Summary of continuous cooling tests for the 1045 carbon steel 167
8.5 Summary of isothermal tests for the alloyed eutectoid steel A 168
8.6 Summary of isothermal tests for the alloyed eutectoid steel A (bainite
reaction) 169
8.7 Summary of isothermal tests for the alloyed eutectoid steel B 170
8.8 Summary of isothermal tests for the alloyed eutectoid steel C 171

9.1 Estimated critical heat flux (CHF) using raw and pre-filtered data for a
38.1 mm-dia. IF steel bar quenched in water flowing at 2.8 m s _1
at 25 C 1 8 9
9.2 Estimated critical heat flux (CHF) as a function of water velocity dur-
ing forced convective quenching of 38.1 mm-dia. IF steel bars in water
flowing at 3 water temperatures 189

9.3 Average heat flux as a function of water velocity during forced convective
quenching of 38.1 mm-dia. IF steel bars in water flowing at 3 water
temperatures 189

x
10.1 Phase boundaries calculated from the equations by Kirkaldy et al. and
the empirical formula given by Andrews 223
10.2 Measured expansion coefficients for the alloyed eutectoid and the 1045
carbon steel 223

12.1 Quenching conditions simulated in the verification runs of the thermal


model 279
12.2 Thermophysical properties of 1045 carbon steel [188] 280
12.3 Thermophysical properties of alloyed eutectoid steel [188] 281
12.4 Thermomechanical properties of IF steel [205] 282
12.5 Thermomechanical properties of 1045 carbon steel [47] 283

F.l Initial values for n and In 6 used to test the robustness of the weighted
non-linear regression analysis 363

xi
List of Figures

1.1 Thermal, microstructural and stress field interactions during heat treat-
ments [7] 5

2.1 Methods adopted to predict the performance of quenching processes. . . 54


2.2 Schematic representation of a typical cooling curve showing the different
stages of boiling 55
2.3 Schematic representation of a typical boiling curve for saturated pool
boiling [75] 56
2.4 Correspondence between cooling, boiling, and evaporation curves [56]. . 57
2.5 Regimes in boiling heat transfer [87] 58
2.6 Schematic representation of fraction transformed as a function of time,
for a typical nucleation and growth phase transformation 59
2.7 Schematic representation of fraction transformed as a function of tem-
perature, for a typical martensitic phase transformation 59
2.8 Schematic diagram illustrating the application of the additivity principle. 60

4.1 Typical cooling curve showing recalescence (schematic) 80


4.2 Flowchart for the solution of the thermal-microstructural problem. . . . 81

4.3 Details of the model sequence for evaluating the heat released during
austenite-to-pearlite decomposition 82
4.4 Comparison between the finite-element and the analytical solution for
the newtonian cooling (Bi= 0.01) of an infinitely long cylinder 83

xii
4.5 Comparison between the finite-element and the analytical solution at the
centreline for the Non-newtonian cooling (Bi= 0.5) of an infinitely long
cylinder 83

4.6 Function Q used in the semi-analytical solution of Newtonian cooling


0

of a long rod with a uniformly distributed heat source 84


4.7 Semi-analytical solution, at the centreline of a 10 mm-dia rod, for Q ,m 0

= 27.5 and 60.0 M W m ~ r "3 1


84
4.8 Effect of At on the semi-analytical solution, at the centreline of a 10
mm-dia. rod, for Q ,
0 m = 600.0 M W n T s"
3 1
85
4.9 Comparison between the semi-analytical and the numerical ( F E M ) solu-
tion, during cooling of a 10 mm-dia. rod, for Q , 0 m = 27.5 M W m - 3
s . _1
85
4.10 Comparison between the semi-analytical and the numerical ( F E M ) solu-
tion, during cooling of a 10 mm-dia. rod, for Q ,m = 60.0 M W m
0
- 3
s . _1
86
4.11 Finite-difference [188] and finite-element solutions for the thermal re-
sponse at the centreline and surface of a 10 mm-dia. eutectoid carbon
steel 86
4.12 Finite-difference [188] and finite-element solutions for fraction trans-
formed to pearlite at the centreline and surface of a 10 mm-dia. eutectoid
carbon steel 87

5.1 (a) Schematic representation of a one-dimensional, single-sensor IHCP


in a flat plate, (b) Discrete temperature measurements 102
5.2 Piecewise aproximation of the surface heat flux as a function of time. . 103
5.3 Flow chart of the sequential function specification algorithm adopted for
the solution of the IHCP 104
5.4 Comparison of estimated and input surface heat flux for Case 1 : constant
surface heat flux 105

xiii
5.5 Comparison of estimated, using the sequential function specification tech-
nique, and analytical thermal response at the surface of the cylinder for
Case 1 : constant heat flux 105

5.6 Comparison of estimated and input heat-transfer coefficient for Case 2 :


constant heat-transfer coefficient 106
5.7 Comparison of estimated, using the sequential function specification tech-
nique, and analytical thermal response at the surface of the cylinder for
Case 2 : constant heat-transfer coefficient 106
5.8 Estimated heat-transfer coefficient for 8 mm-dia. steel rods air cooled at
22 m/s; calculated by using a sequential matching approach and using
"' the sequential function specification algorithm 107
5.9 The functional q = f(T )s adopted for the finite-element simulation of
the thermal response in a 38.1 mm-dia. cylinder subjected to forced
convective quenching 107
5.10 Calculated thermal response at the centreline, surface, and simulated
thermocouple position (r* = 17.75 mm), obtained when the heat flux
distribution shown in Figure 5.9 was applied at the surface of a 38.1
mm-dia cylinder 108
5.11 Effect of varying the parameter r adopted in the inverse analysis on
the estimated surface heat flux during the simulated forced convective
quenching of a 38.1 mm-dia. cylinder 108
5.12 Effect of varying the parameter r on the estimated surface temperature
response during the simulated forced convective quenching of a 38.1 mm-
dia. cylinder 109

5.13 Effect of varying the value of thermal conductivity adopted in the inverse
analysis on the estimated surface heat flux during the simulated forced
convective quenching of a 38.1 mm-dia. cylinder 109

xiv
5.14 Effect of varying the value of the thermocouple position adopted in the
inverse analysis on the estimated surface heat flux during the simulated
forced convective quenching of a 38.1 mm-dia. cylinder 110
5.15 Effect of the thermocouple position on the estimated surface heat flux
during the simulated forced convective quenching of a 38.1 mm-dia.
cylinder 110
5.16 Sensitivity coefficients at several radial positions in a solid cylinder sub-
jected to a medium of constant heat-transfer coefficient and fluid tem-
perature Ill

6.1 Schematic of loading and unloading paths for a work-hardening material


under a uniaxial load 129
6.2 Schematic representation of the scaling factor, r 130
6.3 Flow chart of the stress solution algorithm 131.
6.4 Comparison between analytical and numerical (FEM) predictions of ther-
mal stress distributions in a 100 mm-dia. cylinder 132
6.5 Comparison between analytical and numerical ( F E M ) predictions of ra-
dial displacements in a 100 mm-dia. cylinder 132
6.6 Comparison between measurements made by Buhler [204] and Mitter
et al. [205] and numerical (FEM) predictions of residual stress distribu-
tions in a 50 mm-dia. pure-iron bar quenched in ice water from 850 C [47]. 133

7.1 Schematic diagram showing the quenching test section 148

7.2 Design of the specimen used for IF steel tests 149


7.3 Design of the specimen used for the alloyed eutectoid and 1045 carbon
steel tests 150

7.4 Design of the specimen adaptor 151


7.5 Design of the tapered extension 152

xv
7.6 Calculated velocity and corresponding Reynolds number as a function of
water flow rate for an annular region with Di = 46 mm 153

7.7 Calculated hydrodynamic entry length as a function of Reynolds number. 154


7.8 Calculated thermal entry length as a function of Reynolds number for
various Prandtl numbers [209] 154
7.9 Flow loop of the quenching apparatus 155
7.10 Design of the aluminum cap 156
7.11 Schematic diagram of the data acquisition configuration . 157
7.12 Magnetic flowmeter calibration : water flow rate and current as a func-
tion of valve setting 158
7.13 Temperature response at the centre of the specimen during heating to
(a) 1000 C and (b) 850 C. 159
7.14 Difference between centre and subsurface temperature prior to the start
of the quench for the IF steel tests 160

8.1 Geometry of the Gleeble specimen used for characterizing the phase
transformation kinetics 172
8.2 Location of the sample blanks taken from the 100 mm-dia. rods 172
8.3 Schematic representation of continuous cooling tests 173
8.4 Schematic representation of isothermal tests 173

9.1 Measured temperature response at the centre and subsurface of two 38.1
mm-dia. IF steel bars quenched with water flowing at 4.8 m s _1
at 50 C.190
9.2 Measured temperature response at the centre of a 38.1 mm-dia. IF steel
bar quenched with water flowing at 2.8 m s" for 3 values of water tem-
1

perature 190

xvi
9.3 Measured temperature response at the centre of a 38.1 mm-dia. IF steel
bar quenched with water flowing at 4.8 m s _1
for 3 values of water tem-
perature 191
9.4 Measured temperature response at the centre of a 38.1 mm-dia. IF steel
bar quenched with water flowing at 6.9 m s _1
for. 3 values of water tem-
perature 191
9.5 Measured temperature response at the centre of a 38.1 mm-dia. IF steel
bar quenched with water flowing at 25 C for 3 values of water velocity. 192
9.6 Measured temperature response at the centre of a 38.1 mm-dia. bar
quenched with water flowing at 50 C for 3 values of water velocity. . . 192
9.7 Measured temperature response at the centre of a 38.1 mm-dia. bar
quenched with water flowing at 75 C for 3 values of water velocity. . . 193
9.8 Measured temperature difference between the centre and the subsurface
as a function of subsurface temperature during forced convective quench-
ing of a 38.1 mm-dia. IF steel bar in water flowing at 25 C for 3 values
of water velocity 193
9.9 Measured temperature difference between the centre and the subsurface
as a function of subsurface temperature, during forced convective quench-
ing of a 38.1 mm-dia. IF steel bar in water flowing at 50 C for 3 values
of water velocity ' 194

9.10 Measured temperature difference between the centre and the subsurface
as a function of subsurface temperature during forced convective quench-
ing of a 38.1 mm-dia. IF steel bar in water flowing at 75 C for 3 values

of water velocity 194


9.11 Cooling rate at the centre as a function of local temperature during forced
convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at

25 C for 3 values of water velocity 195

xvii
9.12 Cooling rate at the centre as a function of local temperature during forced
convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at

50 C for 3 values of water velocity 195


9.13 Cooling rate at the centre as a function of local temperature during forced
convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at
75 C for 3 values of water velocity 196
9.14 Maximum cooling rate at the centre as a function of water velocity during
forced convective quenching of a 38.1 mm-dia. IF steel bar for 3 values
of water temperature 196
9.15 Maximum cooling rate at the subsurface as a function of water velocity
during forced convective quenching of a 38.1 mm-dia. IF steel bar for 3
values of water temperature. . 197
9.16 Effect of surface condition on the temperature response at the centre
during forced convective quenching of a 38.1 mm-dia. IF steel bar in
water flowing at 4.8 m s _1
at 25 C 197
9.17 Effect of surface condition on the temperature response at the centre
during forced convective quenching of a 38.1 mm-dia. alloyed steel bar
in water flowing at 4.8 m s _1
at 32 C 198
9.18 Effect of initial test temperature on the temperature response at the
centre during forced convective quenching of a 38.1 mm-dia. IF steel bar

in water flowing at 4.8 m s _1


at 32 C 198
9.19 Raw and filtered temperature response at the subsurface during forced
convective quenching of a 38.1 mm-dia. IF in water flowing at 2.8 m s _1

at 25 C 199
9.20 Experimentally determined temperature response at the subsurface and
centreline during forced convective quenching of a 38.1 mm-dia. IF steel

bar with water flowing at 2.8 m s _1


at 25 C 200

xvm
9.21 Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar. The raw data

of Figure 9.20 were filtered before being input to the computer program. 200
9.22 Comparison between estimated surface heat flux as a function of sur-
face temperature, using raw and filtered data, during forced convective
quenching of a 38.1 mm-dia. IF steel bar with water flowing at 2.8 m s _ 1

at 25 C 201
9.23 Effect of the parameter r on the residuals obtained during the estimation
of surface heat flux during forced convective quenching of a 38.1 mm-dia.
IF steel bar when raw data was used 201
9.24 Effect of the parameter r on the residuals obtained during the estimation
of surface heat flux during forced convective quenching of a 38.1 mm-dia.
IF steel bar when filtered data was used 202
9.25 Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar with water
flowing at 2.8 m s _1
for 3 values of water temperature 202
9.26 Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar with water
flowing at 4.8 m s _1
for 3 values of water temperature 203
9.27 Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar with water
flowing at 6.9 m s _1
for 3 values of water temperature 203
9.28 Estimated surface heat flux as a function of time, during forced con-
vective quenching of a 38.1 mm-dia. IF steel bar with water flowing at

2.8 m s _1
for 3 values of water temperature 204
9.29 Surface heat flux as a function of wall superheat, during forced convective
quenching of thin platinum wires [218] 204

xix
9.30 Estimated surface heat flux as a function of wall superheat, during forced
convective quenching of a 38.1 mm-dia. IF steel bar with water flowing

at 2.8 m s _1
for 3 values of water temperature 205
9.31 Estimated surface heat flux as a function of wall superheat, during forced
convective quenching of a 38.1 mm-dia. IF steel bar with water flowing
at 4.8 m s _1
for 3 values of water temperature 205
9.32 Estimated surface heat flux as a function of wall superheat, during forced
convective quenching of a 38.1 mm-dia. IF steel bar with water flowing
at 6.9 m s _1
for 3 values of water temperature 206
9.33 Estimated and calculated (film boiling) surface heat flux as a function of
surface temperature, during forced convective quenching of a 38.1 mm-
dia. IF steel bar with water flowing at 2.8 m s - 1
for 3 values of water
temperature. 206
9.34 Estimated critical heat flux (CHF) as a function of water velocity, during
forced convective quenching of a 38.1 mm-dia. IF steel bar for 3 values
of water temperature 207
9.35 Average heat flux as a function of water velocity, during forced convective
quenching of a 38.1 mm-dia. IF steel bar for 3 values of water temperature. 207
9.36 Estimated heat-transfer coefficient as a function of surface temperature,
during forced convective quenching of a 38.1 mm-dia. IF steel bar with
water at 25 C for 3 values of water velocity 208
9.37 Photomicrographs of a 38.1 mm-dia. IF steel bar quenched with water
flowing at 4.8 m s _1
at 25 C, at (a) ~ 3 mm from the surface and (b)
centre. Magnification : 500 X 209
9.38 Photomicrographs of a 38.1 mm-dia. alloyed eutectoid steel bar quenched
with water flowing at 2.8 m s _1
at 75 C, at (a) ~ 3 mm from the surface
and (b) centre. Magnification : 400 X 210

xx
9.39 Photograph of the macrostructure of the 38.1 mm-dia. 1045 carbon steel
bar quenched with water flowing at 2.8 m s _1
at 50 C 211
9.40 Photomicrographs of a 38.1 mm-dia. 1045 carbon steel bar quenched
with water flowing at 2.8 m s - 1
at 50 C. (a) within the martensitic
ring, (b) in the transition zone and (c) at the centre. Magnification :
800 X 211
9.40 Photomicrographs of a 38.1 mm-dia. 1045 carbon steel bar quenched
with water flowing at 2.8 m s _1
at 50 C. (a) within the martensitic
ring, (b) in the transition zone and (c) at the centre. Magnification :
800 X 212
9.41 As-quenched hardness distribution in 38.1 mm-dia. IF, alloyed and 1045
carbon steel bars. Use the right axis to read hardness in the IF steel bar. 213
9.42 Hardness distribution near the ends and at mid-length in a 100 mm-
.dia. alloyed eutectoid steel bar quenched and tempered under industrial
conditions 213
9.43 Outline of prior-austenite grain boundaries in a 38.1 mm-dia. alloyed
eutectoid steel quenched bar, etched in a boiling alkaline sodium picrate
solution. Magnification : 1000 X 214
9.44 Measured prior-austenite grain area distribution in a 38.1 mm-dia. al-
loyed eutectoid steel quenched bar 214

10.1 Experimentally determined cooling curves for the alloyed eutectoid steel B

showing the cooling rate for each test 224


10.2 Experimentally determined dilation-temperature curves obtained for con-
tinuous cooling of the alloyed eutectoid steel B showing the cooling rate
for each test 224

xxi
10.3 Continuous cooling test showing the thermal contraction of the austenite
(high temperature) and the low temperature product phase obtained for

the 17.5 C s _1
cooling of the alloyed eutectoid steel B 225
10.4 Continuous cooling test showing the procedure to determine the start of
transformation for the 17.5 C s _1
cooling of the alloyed eutectoid steel B.225
10.5 Continuous cooling test showing the procedure to determine the end of
transformation for the 17.5 C s~ cooling of the alloyed eutectoid steel B.226
x

10.6 Continuous cooling diagram for the alloyed eutectoid steel A 226
10.7 Continuous cooling diagram for the alloyed eutectoid steel B 227
10.8 Continuous cooling diagram for the alloyed eutectoid steel C 227
10.9 Transformation start temperature as a function of cooling rate for the
alloyed eutectoid steels A , B and C 228
10.10 Transformation start temperature plotted as a function of cooling rate
for the alloyed eutectoid Steel A 228
10.11 Measured and predicted C C T transformation start times for the alloyed
eutectoid Steel A 229
10.12 Continuous cooling diagram for the 1045 carbon steel 229
10.13 Transformation start temperature plotted as a function of cooling rate
for the 1045 carbon steel 230
10.14 Measured and predicted C C T ferrite start time for the 1045 carbon steel. 230

10.15 Measured and predicted C C T pearlite start time for the 1045 carbon steel.231
10.16 Experimentally determined isothermal dilation-time curves for the al-
loyed eutectoid steel A 231
10.17 IT diagram for the alloyed eutectoid steel A 232
10.18 IT diagram for the alloyed eutectoid steel B 232
10.19 IT diagram for the alloyed eutectoid steel C 233

xxii
10.20 Kinetic parameters for the pearlitic transformation in the 3 alloyed eutec-
toid steels estimated with a 5-parameter, non-linear regression analysis
(weighted and non-weighted), (a) n; (b) In b 234

10.21 Kinetic parameter 6, as a function of undercooling, for the pearlitic trans-


formation in the 3 alloyed eutectoid steels estimated with a weighted,
5-parameter non-linear regression analysis, for a constant value of n = h. 235
10.22 Kinetic parameter b, as a function of undercooling, for the pearlitic trans-
formation in 3 alloyed eutectoid steels 235
10.23 Kinetic parameter n for the bainitic transformation in the alloyed eutec-
toid steel A , estimated with a 5-parameter, weighted, non-linear regres-
sion analysis 236
10.24 Kinetic parameter b for the bainitic transformation in the alloyed eutec-
toid steel A as a function of undercooling, estimated with a weighted,
5-parameter non-linear regression analysis, for a constant value of n = h. 236
10.25 Microstructure of etched Gleeble specimen (alloyed eutectoid steel A
cooled at 58 C s ) showing the outline of prior austenite grains. . . . 237
_1

10.26 Measured prior-austenite mean chord length distribution obtained in the


alloyed eutectoid steel A Gleeble specimen cooled at 58 C s _ 1
237
10.27 Ferrite fraction as a function of cooling rate for the 1045 (open circles)

and 1025 (filled circles) [98] carbon steel samples 238

11.1 Specimen orientations with respect to neutron beams to measure (a)


radial, (b) circumferential, and (c) axial strain components in force con-
vective quenched steel bars 249

11.1 Specimen orientations with respect to neutron beams to measure (a)


radial, (b) circumferential, and (c) axial strain components in force con-
vective quenched steel bars 250

xxin
11.2 Schematic diagram showing the cutting sequence used to extract refer-
ence slices for neutron diffraction measurements 251
11.3 Measured diffraction peak profiles in a 38.1 mm-dia. IF steel bar quenched
in water flowing at 4.8 m s _1
at 25 C 252
11.4 Measured diffraction peak profiles in a 38.1 mm-dia. 1045 steel bar
quenched in water flowing at 2.8 m s _1
at 50 C 252
11.5 Measured diffraction peak profiles in a 38.1 mm-dia. alloyed steel bar
quenched in water flowing at 2.8 m s _1
at 75 C showing the position
of the F C C austenite ( A ( ) ) and B C T martensite ( M ( ) ) peaks. Data
in U0

collected during Campaign 1 253


11.6 Measured diffraction peak profiles in a 38.1 mm-dia. alloyed steel bar
quenched in water flowing at 2.8 m s _1
at 75 C. Data collected during
Campaign 2 253
11.7 Measured and fitted diffraction peak profile in a 38.1 mm-dia. IF steel
bar quenched in water flowing at 4.8 m s _1
at 25 C 254
11.8 Measured and fitted diffraction peak profile in a 38.1 mm-dia. 1045
carbon steel bar quenched in water flowing at 2.8 m s _1
at 50 C. . . . 254
11.9 Measured and fitted diffraction peak profile in a 38.1 mm-dia. alloyed
steel bar quenched in water flowing at 2.8 m s _1
at 75 C . Data collected
during Campaign 2 255

11.10 Mean scattering angle as a function of radial position obtained from the
strain-free reference slices extracted from the 1045 carbon steel specimen. 255

11.11 Control volumes adopted for force balance calculations : (a) axial com-
ponent (plan view) and (b) circumferential component 256

11.12 Measured radial, circumferential (hoop) and axial residual strains as a


function of radial position in a 38.1 mm-dia. IF steel bar quenched in
water flowing at 4.8 m s _1
at 25 C 257

xxiv
11.13 Measured radial, circumferential (hoop) and axial residual strains as a
function of radial position in a 38.1 mm-dia. 1045 carbon steel bar
quenched in water flowing at 2.8 m s _1
at 50 C. 257
11.14 Measured radial, circumferential (hoop) and axial residual strains as a
function of radial position in a 38.1 mm-dia. alloyed steel bar quenched
in water flowing at 2.8 m s _1
at 75 C. Data collected during Campaign 2.258
11.15 Measured radial, circumferential (hoop) and axial residual stresses as a
function of radial position in a 38.1 mm-dia. IF steel bar quenched in
water flowing at 4.8 m s _1
at 25 C 258
11.16 Measured radial, circumferential (hoop) and axial residual stresses as
a function of radial position in a 38.1 mm-dia. 1045 carbon steel bar
quenched in water flowing at 2.8 m s _1
at 50 C 259

11.17 Measured radial, circumferential (hoop) and axial residual stresses as a


function of radial position in a 38.1 mm-dia. alloyed steel bar quenched
in water flowing at 2.8 m s _1
at 75 C. Data collected during Campaign 2.259

12.1 Computational domain adopted for the finite-element simulation of ther-


mal/mi crostructural evolution during forced convective quenching. . . . 284
12.2 Schematic representation of the boundary conditions adopted for the
thermal/mi crostructural simulations 284
12.3 Finite-element mesh adopted for the thermal/mi crostructural simulations. 285
12.4 Comparison between predicted and measured temperature responses at
the centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with
water flowing at 4.8 m s _1
at 50 C. The boiling curve shown as a broken
line in Figure 12.5 was adopted as boundary condition 286
12.5 Boiling curves adopted as boundary condition to simulate forced con-
vective quenching of a 38.1 mm-dia. IF steel bar quenched with water
flowing at 4.8 m s- at 50 C
1
286

XXV
12.6 Comparison between predicted and measured temperature responses at
the centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with
water flowing at 4.8 m s _1
at 50 C. The boiling curve shown as a solid

line in Figure 12.5 was adopted as boundary condition 287


12.7 Comparison between predicted and measured temperature responses at
the centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with
water flowing at 6.9 m s _1
at 25 C 287
12.8 Comparison between predicted and measured temperature responses at
the centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with
water flowing at 2.8 m s" at 75 C
1
288
12.9 Effect of varying the value of thermal conductivity by 10 % adopted in
the finite-element simulation on the temperature response at the centre
and subsurface of a 38.1 mm-dia. IF steel bar quenched with water
flowing at 4.8 m s" at 50 C
1
288
12.10 Effect of varying the surface heat flux by 10 % adopted in the finite-
element simulation on the temperature response at the centre and sub-
surface of a 38.1 mm-dia. IF steel bar quenched with water flowing at
4.8 m s" at 50 C
1
289
12.11 Comparison between predicted and measured temperature responses at
the centre of a 38.1 mm-dia. IF steel bar quenched with water flowing

at 4.8 m s- at 25 C .
1
. . . 289
12.12 Comparison between predicted and measured temperature responses at
the subsurface of a 38.1 mm-dia. IF steel bar quenched with water flowing
at 4.8 m s- at 25 C
1
290
12.13 Comparison between predicted and measured cooling rates at the centre
of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m s _1

at 25 C 290

xxvi
12.14 Comparison between predicted and measured cooling rates at the sub-
surface of a 38.1 mm-dia. IF steel bar quenched with water flowing at

4.8 m s- at 25 C
1
291
12.15 Calculated temperature gradients at five different times during forced
convective quenching of a 38.1 mm-dia. IF steel with water flowing at
4.8 m s" at 25 C
1
291
12.16 Calculated microstructural evolution at the centre during forced convec-
tive quenching of a 38.1 mm-dia. IF steel bar with water flowing at 4.8
m s" at 25 C
1
292
12.17 Calculated microstructural evolution at the surface during forced con-
vective quenching of a 38.1 mm-dia. IF steel bar with water flowing at
4.8 m s- at 25 C
1
292
12.18 Comparison between predicted and measured temperature responses at
the centre of a 38.1 mm-dia. 1045 carbon steel bar quenched with water
flowing at 2.8 m s _1
at 50 C 293
12.19 Comparison between predicted and measured temperature responses at
mid-radius of a 38.1 mm-dia. 1045 carbon steel bar quenched with water
flowing at 2.8 m s _1
at 50 C 293
12.20 Comparison between predicted and measured cooling rates at the centre
of a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at

2.8 m s- at 50 C
1
294
12.21 Comparison between predicted and measured cooling rates at mid-radius
of a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at

2.8 m s- at 50 C
1
294
12.22 Calculated temperature gradients at five different times during forced
convective quenching of a 38.1 mm-dia. 1045 carbon steel bar with water
flowing at 2.8 m s _1
at 50 C 295

XXVll
12.23 Calculated (with modified kinetics) microstructural evolution at the cen-
tre during forced convective quenching of a 38.1 mm-dia. 1045 carbon
steel bar with water flowing at 2.8 m s _1
at 50 C 295

12.24 Calculated (with modified kinetics) microstructural evolution at the sur-


face during forced convective quenching of a 38.1 mm-dia. 1045 carbon
steel bar with water flowing at 4.8 m s _1
at 50 C 296
12.25 Calculated (with modified kinetics) final microstructural distribution in
a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at
4.8 m s" at 50 C
1
296
12.26 Comparison between predicted and measured temperature responses at
the centre of a 38.1 mm-dia. alloyed eutectoid steel bar quenched with
water flowing at 2.8 m s" at 75 C
1
297
12.27 Comparison between predicted and measured temperature responses at
the subsurface of a 38.1 mm-dia. alloyed eutectoid steel bar quenched
with water flowing at 2.8 m s" at 75 C
1
297
12.28 Comparison between predicted and measured cooling rates at the centre
' of a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water flowing
at 2.8 m s" at 75 C
1
298
12.29 Comparison between predicted and measured cooling rates at the subsur-
face of a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water
flowing at 2.8 m s _1
at 75 C 298

12.30 Calculated temperature gradients at five different times during forced


convective quenching of a 38.1 mm-dia. alloyed eutectoid steel bar with
water flowing at 2.8 m s" at 75 C
1
299

12.31 Calculated microstructural evolution at the centre during forced convec-


tive quenching of a 38.1 mm-dia. alloyed eutectoid steel bar with water
flowing at 2.8 m s _1
at 75 C 299

xxvm
12.32 Calculated microstructural evolution at the surface during forced convec-
tive quenching of a 38.1 mm-dia. alloyed eutectoid steel bar with water
flowing at 2.8 m s" at 75 C
1
300

12.33 Calculated final microstructural distribution in a 38.1 mm-dia. alloyed


eutectoid steel bar quenched with water flowing at 2.8 m s - 1
at 75 C. . 300
12.34 Effect of varying the value of M adopted in the finite-element simulation
s

on the temperature response at the centre and subsurface of a 38.1 mm-


dia. alloyed eutectoid steel bar quenched with water flowing at 2.8 m s - 1

at 75 C 301
12.35 Effect of varying the value of M adopted in the finite-element simulation
s

on the microstructural evolution at the centre and subsurface of a 38.1


mm-dia. alloyed eutectoid steel bar quenched with water flowing at 2.8
m s _1
at 75 C 301
12.36 Effect of varying the value of the kinetic constant in the Koistinen-
Marburger equation adopted in the finite-element simulation on the tem-
perature response at the centre and subsurface of a 38.1 mm-dia. alloyed
eutectoid steel bar quenched with water flowing at 2.8 m s _1
at 75 C. . 302
12.37 Effect of varying the value of the kinetic constant in the Koistinen-
Marburger equation adopted in the finite-element simulation on the mi-
crostructural evolution at the centre and subsurface of a 38.1 mm-dia.
alloyed eutectoid steel bar quenched with water flowing at 2.8 m s _1
at

75 C 302
12.38 Finite-element mesh adopted to simulate the thermal/mi crostructural
evolution and stress generation during quenching 303

12.39 Computational domain adopted for the finite-element simulation of stress


generation during forced convective quenching 304

xxix
12.40 Schematic representation of the boundary conditions applied for the ther-
mal/mi crostructural simulations 304
12.41 Schematic representation of the boundary conditions applied for the
stress simulations 305
12.42 Comparison between predicted and measured residual stress distribution
(radial component) in a 38.1 mm-dia. IF steel bar quenched with water
flowing at 4.8 m s" at 25 C.
1
305
12.43 Comparison between predicted and measured residual stress distribu-
tion (circumferential (hoop) component) in a 38.1 mm-dia. IF steel bar
quenched with water flowing at 4.8 m s _1
at 25 C 306
12.44 Comparison between predicted and measured residual stress distribution
(axial component) in a 38.1 mm-dia. IF steel bar quenched with water
flowing at 4.8 m s" at 25 C
1
306
12.45 Comparison between predicted and measured residual stress distribution
(radial component) in a 38.1 mm-dia. 1045 carbon steel bar quenched
with water flowing at 2.8 m s _1
at 50 C. Modified kinetics 307
12.46 Comparison between predicted and measured residual stress distribution
(circumferential (hoop) component) in a 38.1 mm-dia. 1045 carbon steel
bar quenched with water flowing at 2.8 m s _1
at 50 C . Modified kinetics.307
12.47 Comparison between predicted and measured residual stress distribution
(axial component) in a 38.1 mm-dia. 1045 carbon steel bar quenched with
water flowing at 2.8 m s _1
at 50 C. Modified kinetics 308
12.48 Comparison between predicted and measured residual stress distribu-
tion (radial component) in a 38.1 mm-dia. alloyed eutectoid steel bar
quenched with water flowing at 2.8 m s _1
at 75 C 308

xxx
12.49 Comparison between predicted and measured residual stress distribution
(circumferential (hoop) component) in a 38.1 mm-dia. alloyed eutectoid
steel bar quenched with water flowing at 2.8 m s _1
at 75 C 309
12.50 Comparison between predicted and measured residual stress distribution
(axial component) in a 38.1 mm-dia. alloyed eutectoid steel bar quenched
with water flowing at 2.8 m s _1
at 75 C 309
12.51 Evolution of radial, circumferential (hoop) and axial stresses at the sur-
face during forced convective quenching of a 38.1 mm-dia. alloyed eutec-
toid steel bar 310
12.52 Evolution of radial, circumferential (hoop) and axial stresses at the centre
during forced convective quenching of a 38.1 mm-dia. alloyed eutectoid
steel bar 310
12.53 Evolution of radial, circumferential (hoop) and axial stresses at the sur-
face during forced convective quenching of a 38.1 mm-dia. 1045 carbon
steel bar 311
12.54 Evolution of radial, circumferential (hoop) and axial stresses at the centre
during forced convective quenching of a 38.1 mm-dia. 1045 carbon steel
bar 311
12.55 Calculated final microstructural distribution in a 100 mm-dia. alloyed
eutectoid steel bar quenched from 850 C with water flowing at 4.8 m

s- 1
at 32 C 312

E.l Start (closed circles) and finish (open circles) temperatures during con-
tinuous cooling of tubular IF steel specimens 358
E.2 Predicted temperature response at the centre and mid-radius of a 38.1
mm-dia. IF steel bar quenched with water flowing at 2.8 m s _1
at 75 C .
The start and end of the transformation are shown as broken lines. . . . 358

xxxi
F.l (a) Exact data corresponding to n = 2, In 6 = 11 and tAv 0.5 s. The
error bars correspond to 0.05 units and are equal at all levels of t. (b)
Exact data linearized using Eq. ( F . l ) 364
F.2 Non-linear regression applied to exact isothermal kinetic data 365
F.3 Residuals plot against estimated values corresponding to the non-linear
regression of Figure F.2 365
F.4 Comparison of weighted and non weighted multiple non-linear regression
applied to exact data : (a) estimated and observed values, (b) residuals
plot against estimated values 366
F.5 Multiple (3 parameter) weighted non-linear regression applied to exper-
imental data : estimated and observed values 367
F.6 Multiple (5 parameter) weighted non-linear regression applied to exper-
imental data : estimated and observed values 367

xxxn
List of Symbols

w e
-- nodal displacement vector, m
A -- area, m 2

4 -- partition coefficient, dimensionless

A -- cross-sectional area, m 2
L

-- frame area, / i m 2
AF

Aj ~- Jeffries number, / m i 2

b -- kinetic parameter in Avrami equation, dimensionless

[B] -- strain-nodal displacement matrix, m _ 1

Bi -- Biot number, dimensionless

[C] -- capacitance matrix, J C _ 1

CR -- cooling rate, C s _1

Cp - heat capacity, J k g - 1
K _ 1

D -- diameter, mm

D 0 -- initial Gleeble specimen diameter, mm

D^ki -- material elastic constant tensor, M P a

D e -- equivalent diameter, m

D -- dilation at the start of the transformation, mm


min

^max
- dilation at the end of the transformation, mm

D 0 -- initial gleeble specimen diameter, mm

E -- Young's modulus, M P a

xxxm
EQAD - .equivalent area grain diameter pim
{/} - load vector (heat transfer), W
{F} - internal force vector, N
F - yield function, M P a
Fo - Fourier number, dimensionless
g - plastic potential function, M P a
G - mass flux, kg m - 2
s _1

h - heat-transfer coefficient, W m - 2
C _ 1

H p - plastic modulus, M P a
AiJ t r a n s f - heat of transformation, J k g - 1

i - current, m A
J 2 - second invariant of the deviatoric stress tensor, (MPa) 2

J 0 - Bessel function of the first kind of order zero, dimensionless


Ji - Bessel function of the first kind of order one, dimensionless
[K] - stiffness matrix,
k - thermal conductivity, J m _ 1
K _ 1
s _1

ki - ternary distribution coefficient, dimensionless


L - characteristic length, m
L h y d - hydrodynamic entry length, mm
L th - thermal entry length, mm
M s - martensite start temperature, C

M f - martensite finish temperature, C


n - kinetic parameter in Avrami equation, dimensionless
[N] - shape function matrix, dimensionless

NQ - number of grains, dimensionless

xxxiv
p - pressure, M P a
P w - wetted perimeter, m
Pr - Prandtl number, dimensionless

q - heat flux, M W m - 2

q avg - average heat flux, M W m ~ 2

c7 chf - critical heat flux, M W m - 2

g m a x - maximum heat flux, M W m - 2

<?rms - root mean square of the error in q, dimensionless


Q - water flow rate, 1 s _1

r - number of future time steps, dimensionless


r
min ~ minimum value of the scaling factor, dimensionless
R - radius, mm
{R} - load vector (stress), N
Re - Reynolds number, dimensionless
5* - least squares function, C 2

t - time, s
A - time increment, s
t* - virtual time, s
tAVCOT - transformation start time (CCT), s
IAV ~ transformation start time (IT), s
T - temperature, C
T A e 3 - Ae3 temperature, C
T A c m - Acm temperature, C

T c - centre temperature, C

xxxv
Tf - fluid temperature, C
T 0 - initial temperature, C
T s u r f - surface temperature, C
T s - transformation start temperature, C
T g a t - saturation temperature, C
T AL - C

TA, - C
AT s a t - wall superheat, C
A r
sub - subcooling, C
{u} - displacement vector, m
u\ - undercooling below T ^ , C
U3 - undercooling below TA , C 3

v - water velocity, m s _1

Wi - weight in regression analysis, dimensionless


X - fraction transformed, dimensionless
Xj - sensitivity coefficient at sensor j, C W _ 1
m 2

Yj - measured temperature at sensor j, C

a - linear thermal expansion coefficient, C _ 1

(3 - ' linear expansion due to transformation, m m"


5ij - Kronecker delta, dimensionless
e - strain, m m _ 1

e e
- elastic strain, m m _ 1

e - 'initial' strain, m m _ 1

p
- plastic strain, m m _ 1

xxxvi
6 th
-- thermal strain, m m _ 1

tr _- transformation strain, m m _ 1

tp _- transformation plasticity strain, m m _ 1

AE _- strain due to variation of elastic properties with temperature, m m


e AF
-- strain due to variation of flow stress with temperature, m m _ 1

Sp - effective plastic strain, m m _ 1

K -- hardening parameter,
A -- wavelength, A

A m
- eigenvalue, m - 1

dA -- loading parameter, dimensionless

-- viscosity, kg m s
- 1 _1
^

v -- poisson's ratio, dimensionless


# -- residual force vector, M P a

p -- density, kg m - 3

a - stress, M P a

o- ~- e
effective stress, M P a
a' -- deviatoric component of stress, M P a
T -- isothermal incubation time, s
6 -- mean scattering angle, degrees

0o -- reference mean scattering angle, degrees

u -- relaxation parameter, dimensionless

xxxvn
Coordinates
r radial, m
z axial, m
9 circumferential (hoop), m
Superindices
e element level

m previous time step

m +1 current time step

xxxvni
Acknowledgements

I would like to thank Dr. E . B . Hawbolt and Dr. J.K. Brimacombe for their interest and
suggestions throughout the course of this work. The assistance of the technical staff at
U B C , in particular Mr. B . Chau, M r . P. Musil and M r . R. Cardeno is greatly appreci-
ated. Thanks are also extended to Mr. J . Lapointe for his help during the construction
of the laboratory quenching facility. The residual stress measurements could not have
been made without the guidance and technical expertise of Dr J . Root and the person-
nel at A E C L (Chalk River Laboratories). The assistance of Prof. J . V . Beck (Michigan
State University) in providing the original code CONTA is gratefully acknowledged.
Discussions with Dr. S. Cockroft and Dr. Y . Nagasaka on the implementation of their
respective computer programs have been also invaluable.
Thanks are also extended to my fellow graduate students for their voluntary assistance
and to my friends who made this a very enjoyable stay. I would like to express my
gratitude to the Universidad Nacional Autonoma de Mexico (Departamento de Ingenieria
Quimica Metalurgica) and the Department of Metals and Materials Engineering of the
University of British Columbia for the financial support given during my professional
studies.

Finally, I would like to thank the members of my family for their understanding and
encouragement.

xxxix
Chapter 1

Introduction

1.1 Microstructural Engineering

Heat treatment processes constitute a very important operation in the processing of engi-
neering components. Through a combination of heating and cooling cycles, the optimum
microstructure required to obtain the desired mechanical properties can be produced. At
the same time, however, residual stresses may be introduced into the component due to
volume changes associated with thermal contraction and phase transformations which
may generate distortion and even cracking. However, residual stresses are not always
detrimental. For example, surface compressive residual stresses contribute to the frac-
ture resistance of case-hardened components [1]. It follows that while distortion must
be minimized, the residual stress pattern produced by a given heat treatment must be
known and optimized.

The need for predicting the mechanical properties and residual stress distribution
resulting from a given heat treatment has long been recognized. More stringent require-
ments of high productivity, performance and low energy consumption have favoured re-
search towards the linking of mechanical properties and process variables. Consequently,
the trial-and-error procedures that have been followed in the past are being replaced by
more fundamental methodologies which incorporate concepts of transport phenomena,
mechanical behaviour, solid-solid phase transformations, and microstructure-properties
relationships, within a paradigm known as 'microstructural engineering'.

1
Introduction 2

1.2 Heat Treatment of Grinding Media

Grinding using either steel rods or balls is the greatest operating cost in mineral process-
ing [2]. Most of this cost stems from rod/ball and energy consumption [3]; thus, the wear
rate of grinding media in rod mills becomes a factor of economic significance, and meth-
ods to reduce grinding media wear, and steel consumption in general, are desirable [4].
By using grinding media of higher quality and, thus, reducing breakage, the grinding rods
can be worn to smaller sizes which can reduce operating costs significantly [5]. Abra-
sion, impact, erosion, and corrosion, contribute to the wear of grinding media under wet
grinding [3]. Typically, grinding media are fabricated of high carbon steel in order to
maximize attainable hardness. The addition of alloying elements improves hardenability
and hardness and, therefore, abrasion resistance, but it may also result in a more brittle
product. A combination of alloying additions and heat treatment is, therefore, required
to accomplish optimum mechanical behaviour during service of grinding media in rod
mills.
Pugh and M a [6] reported a reduction in grinding rod consumption of up to 25 %
when heat treated grinding rods of 100 mm (4 in) dia. were used in production trials
at the Endako Mines Division of Placer Dome Inc. The proprietary heat treatment
process developed at AltaSteel is based on forced convective quenching of steel rods,
followed by self-tempering, and produces a material with a tempered martensitic case
and a bainitic/pearlitic core. The hardness profile ranges from Rc 55 at the surface to
Rc 40 at the core. The mill tests [6] showed a low initial wear rate; the high surface
hardness resulted in an almost constant wear rate until the first 30 mm were worn. By
comparison, the initial wear rate in AISI 1090 rods, tested under the same conditions,
was twice as high as that of the heat treated rods and decreased as the rod wore down.
Clearly, the microstructural distribution plays an important role on wear rate behaviour
of grinding rods, as it affects both the abrasion resistance and the toughness.

Despite improved rod performance, there are production problems that need to be
Introduction 3

addressed. Large transient stresses, responsible for distortion in the final product, are
caused by non uniform cooling and/or phase evolution. In the long (1.8 m) rods required
for grinding mills, end effects produce significant distortion during quenching; the ends
having to be straightened using induction" coils. Variability in the "equalization" tem-
perature, i.e., the maximum temperature reached during self-tempering has also been
observed. Another area of interest for rod producers is that of quality control; at the
moment only crude impact tests are used to asses product quality. Thus, in order to
optimize the process, as well as to assist in future developments, it is desirable to gen-
erate a predictive tool that would obviate the need for expensive and time-consuming
trial-and-error operations. In consequence, fundamental knowledge related to the process
needs to be generated.

1.3 Statement of the Problem

The thermal, microstructural and stress fields obtained during heat treatment interact
in a complex manner as shown in Figure 1.1 [7]. In general, the temperature distribution
upon cooling is nonhomogeneous, which generates thermal stresses (interaction 1). The
microstructural evolution depends, basically, on the cooling history of the component
(interaction 2) and, at the same time, modifies the temperature response through the
latent heat evolved during the transformations (interaction 6). A n unequal volume change
distribution accompanying the phase transformations sets up transformation stresses
(interaction 3) which, in turn, may affect the transformation kinetics (interaction 5).
Finally, the internal stresses generated upon cooling do mechanical work, part of which
is converted into heat (interaction 4); however, the amount of heat generated is negligible
and, therefore, this interaction is usually ignored.

In addition to these effects, material-related properties such as temperature-dependent


thermophysical and thermomechanical properties, prior-austenite grain size distribution
Introduction 4

and carbon concentration gradients (in case hardening) play an important role in deter-
mining the final microstructure and residual stress distributions as well as distortion.

In this study, the concept of microstructural engineering has been applied to study
forced convective quenching of steel bars. The approach taken involved the development
and application of mathematical models of heat transfer, microstructural evolution and
stress generation, coupled with the experimental determination of the active heat-transfer
boundary condition in a laboratory facility designed and built for that purpose. Addition-
ally, phase transformation kinetics were measured in tubular specimens, and the residual
stress distribution and final microstructure and hardness in forced convective quenched
steel bars were determined.
Introduction 5

\ thermal stress

4. heat g e n e r a t e d by
mechanical w o r k
5. stress-induced
transformation. 3. t r a n s f o r m a t i o n
2 t e m p e r a t u r e \ \ 6 . latent stress
dependent p h a s \ V e a t
transformation
metallic structures

Figure 1.1: Thermal, microstructural and stress field interactions during heat treatments
[TI-
Chapter 2

L i t e r a t u r e Review

In this chapter, the literature relevant to the problem is reviewed. First, a general
overview of the methods available to design a heat treatment is presented. Then, mate-
rial related to the kinetics of phase transformations in steels, heat transfer under forced
convective boiling conditions, and methods for residual stress measurement is reviewed.
The final section is dedicated to a discussion of published mathematical models of mi-
crostructural evolution and stress generation during heat treatments.

2.1 Heat Treatment Design

Several approaches have been developed to predict the performance of a given quenching
process. They can be classified as : i) trial-and-error, ii) characterization of quench bath
quality, and iii) process modeling. A diagram showing the conceptual differences among
these techniques is given in Figure 2.1. As can be seen from this figure, the complexity
increases as one moves from the trial-and-error approach to the process modeling method-
ology as does the insight gained on the effect of process parameters. In the following, a
review of these methodologies is presented.

2.1.1 Empirical Approaches

In the past, heat treatment design was conducted exclusively via trial-and-error proce-
dures which are clearly expensive as well as time consuming. They effectively act as a

6
Literature Review 7

'black box' and, therefore, the results obtained can only be applied under a restricted
set of conditions. Factorial design provides a systematic method for conducting experi-
ments, but still fails to give information on the processes that take place during a heat
treatment. After selecting the factors that are considered to have an impact on the final
result, any one of the standard factorial design plans can be applied. In the case of
factorial experimentation (the most commonly used of these standard plans) all levels of
a given factor are combined with all levels of each of the other factors [8]. Two examples
where factorial design has been successfully applied are the heat treatment of nickel-iron-
base superalloys (to optimize for room-temperature tensile properties) [9], and the heat
treatment of gears (to optimize for hardness and distortion) [10].

2.1.2 Characterization of Quench-Bath Quality

The resulting mechanical properties after a heat treatment are linked to the microstruc-
ture which, in turn, is governed by the heat extraction characteristics of the quench
bath. Over the years, a number of tests have been designed to measure the quality of the
quench bath, based on either its hardening power (hardenability, Jominy end quench) or
its cooling power ( G M quenchometer, hot wire test, interval five-second test, and cooling
curve) [11].

2.1.2.1 Quench-Bath Quality Based on Hardening Power

Early attempts to predict mechanical properties were concentrated on the prediction of


the hardness that can be obtained from a quenching operation. Thus, the concept of
hardenability, 'the depth of useful hardness which can be produced for given quenching
conditions' [12] evolved.

The hardenability of a particular steel can be predicted, based on its chemistry, using
the Grossman method [13]. This method characterizes the effect of the amount of each
Literature Review 8

alloying element in a steel on the hardenability as a multiplying factor which is used to


obtain an ideal critical diameter (50 % martensite at the centre of a bar quenched in a
medium of infinite severity); it should be noted that only a linear effect of composition
on hardenability is considered. This method serves only as a tool for comparing steels
of different compositions since the heat transfer characteristics of the cooling media are
described through an average quantity called 'quench severity'. Kirkaldy [14] pointed out
the lack of generality of this method (as evidenced by the number of sets of multiplying
factors that have been reported) and attributed it to the linearity of the function adopted
for the calculation of the ideal diameter. A critical review of hardenability predictors
based on chemical composition alone can be found in [15].
The concept of hardenability is qualitative in nature and, therefore, methods for
describing a quantitative relationship between cooling rates and hardness are required.
The Jominy end-quench test provides such a relationship. In order to be able to predict
the behaviour of a given part, a Jominy test should be conducted under standardized
austenitizing and quenching conditions and criteria to correlate experimental and indus-
trial results should be proposed. Such empirical relationships between end-quench bar
positions and positions in the actual piece are available in the literature for plates [16]
and bars [17].
Hardenability predictors can be very useful in the heat treatment industry. Kirkaldy
et al. [18] reviewed the state of the art in hardenability prediction and concluded that
the Jominy test can be economically replaced by properly calibrated predictors. A n
additional advantage is the possibility of on-line control of the process. One such hard-
enability predictor was developed by Kirkaldy and Venugopalan [14]. Umemoto et al. [19]
used isothermal transformation kinetics information to predict martensite fraction as a
function of distance from the quenched end in a Jominy bar and obtained very good
agreement between predicted and experimental values.

Continuous cooling transformation (CCT) diagrams are also used for designing heat
Literature Review 9

treatments [20]. Again, a criterion is needed to correlate the C C T diagrams obtained in


a laboratory with the response in actual practice. Rose et al. [21] proposed the use of the
time to cool from AC3 to 500 C as the criterion for an equivalence between C C T diagrams
and end-quench tests. Cias [22] found a good correlation when the half-cooling time (the
time to reach the mean temperature between AC3 and room temperature) concept was
used.
It should be noted that Jominy curves as well as C C T diagrams are strictly valid for
the cooling and austenitizing conditions under which they were derived. Also, published
results may contain empirical errors - as pointed out in [18].
For automation reasons it would be convenient to create a description of C C T dia-
grams that would lend itself to automatic retrieving. One possibility (suggested in [23])
would be to save a compilation of C C T curves arranged by composition and grain size
such that upon request the closest C C T curve for a given chemistry and grain size could
be produced. Alternatively, a set of linear regression formulas calibrated to the property
of interest could be deduced. The latter approach has been adopted to generate formu-
lae for the prediction of microstructure and hardness of low-alloy steels in the form of
critical cooling velocities at 700 C necessary to attain a given microstructure [24]. The
austenitizing conditions have been characterized by a thermally activated grain growth
parameter called the austenitizing parameter.

2.1.2.2 Q u e n c h - B a t h Q u a l i t y B a s e d on C o o l i n g Power

While methods based on hardening power are concerned with the ability of a quench
bath to produce a given hardness distribution, methods that measure the cooling power
are directed towards evaluating the amount of heat that can be extracted by a quench
bath.

In the Quenchometer test [25], a 22 mm-dia. nickel sphere is heated to 885 C and
subsequently dropped into a wire basket which is suspended in 200 ml of quench oil kept
Literature Review 10

at a temperature between 20 and 30 C. The time the sphere spends between 885 and
354 C is recorded by first activating a switch when the sphere passes a photoelectric
sensor and falls into the quenching bath and then deactivating it when the nickel becomes
magnetic (at its Curie temperature, i.e., 354 C). Typical Quenchometer values for slow,
medium and fast quenching oils are 15 to 20, 11 to 14, and 8 to 10 seconds, respectively.

Quenchometer results are a good indicator for monitoring oil degradation but it has
not been possible to correlate them to final properties in the quenched part. A major
drawback of this technique is that it covers only a portion of the cooling path. Con-
versely, cooling curves covering the entire range of temperatures can be obtained from
instrumented cylindrical metal probes [11,26]. The probes are commonly made of either
304 stainless steel or Inconel 600 and are instrumented with a single thermocouple lo-
cated in the geometric centre (probe diameter ranges from 12.7 to 50 mm). From the
experimentally obtained temperature vs time curve a cooling rate curve is computed di-
rectly as the first derivative of temperature with respect to time. Bates and Totten [26]
obtained cooling curves and cooling rate curves for a 25 mm-dia., 304 stainless steel
probe quenched in agitated slow, medium and fast quenching oil (with Quenchometer
speeds of 21, 14 and 8.5 s). The computed maximum cooling rate of the slow oil was
higher than for medium speed oil, which could not have been detected by a Quenchome-
ter test alone. Although more informative than the Quenchometer test, the cooling curve
method suffers from the fact that a thermocouple located at the geometrical centre of any
specimen does not reflect accurately the heat transfer at the surface. Tamura [27] recog-
nized this shortcoming and conducted a study involving both the commonly used probe
(made of Inconel 600 and instrumented with a thermocouple in the geometrical centre)
and a Japanese Industrial Standard (JIS) specimen (made of silver and instrumented
with a thermocouple at its surface at mid-length). Both probes are solid cylinders. He
compared the response of probes quenched in oil with varying amounts of additives and
Literature Review 11

showed that the JIS specimen is more sensitive to changes in heat transfer characteris-
tics of the quench bath. A n important question regarding the measurement of a cooling
curve is that of standardisation, as pointed out by Bodin and Segerberg [28]. Several
national standards have been used and a draft proposal for an international standard
for oil quenching is under review; a proposal for a similar test for water-based polymer
quenchants is being formulated.

Once a cooling curve and the corresponding cooling rate curve have been obtained
they can be analyzed in a number of ways varying from visual comparison (i.e., plotting
several cooling curves side by side in order to compare quench baths) to attempts to
correlate cooling curve data to mechanical properties (by including phase transformation
characteristics of the specimen) [11,28]. Among the latter, the quench factor analysis
(QFA) has received a great deal of attention in the aluminum industry due to its ability
to predict the performance of quenching media [29,30]. This method requires a cooling
curve and a T T P C-curve (iso-strength contours plotted in a temperature vs time graph)
1

and is similar to the additivity rule [31]. The quench factor is essentially the summation,
over a temperature range, of fractional times (time spent at a given temperature divided
by the time required to reach the T T P curve isothermally at that particular temperature)
for a given cooling curve. Both, type and depth of corrosion as well as maximum yield
strength have been correlated to quench factor [29]. In general terms, low values of quench
factor are associated with high quench rates and high strengths [11]. Bates [32] applied
the concept of quench factor analysis to the selection of quench baths for aluminum
parts considering attainable tensile properties as well as distortion. In order to minimize
distortion, a quenchant that will produce a heat-transfer coefficient slightly higher than
the equivalent quench factor in the critical section thickness should be selected [32]. Bates
and Totten [33] successfully predicted the as-quenched hardness of 4130, 4140 and 1045

1
Even though a T T P curve follows the typical s-shape associated with T T T and C C T curves, it
differs in that it is calibrated with respect to the property of interest for a given chemistry
Literature Review 12

steels using QFA.


Reti et al. [34] have taken this concept one step further and developed the so-called
Cooling/Transformation Analysis (CTA). The C T A is a computerized technique devel-
oped to quantitatively characterize the quenching performance of a given quenching
medium. Like Q F A , it provides a single figure (the hardening power) that relates the
cooling condition to the resulting microstructure. The starting point for this technique is
a measured, directly or indirectly, surface heat flux history in a series of standard probes
quenched in several media. The heat flux function is characterized by the maximum heat
flux as well as the average heat flux values for the 300 - 400 C, 400 - 500 C , 500 -
600 C , 600 - 700 C and 700 - 850 C ranges. From this information the corresponding
cooling curves and cooling rate curves are obtained. The parameters that characterize
the cooling rate curves are the maximum cooling rate, the temperature corresponding to
this figure, and the mean cooling rates over the 300 - 400 C , 400 - 500 C, 500 - 600 C,
and 600 - 700 C ranges. Experiments based on computer simulations [34] showed that
C T A results are compatible with results produced through cooling curve analysis and
QFA.
It is interesting to note that studies on cooling curves have shown differences between
hardness predictions based on C C T diagrams and those based on cooling curve analysis
[35-37]. The reason for this discrepancy is that C C T diagrams are commonly constructed
from 'linear' cooling conditions with cooling rates that do not vary significantly and,
therefore, the heat extraction history is quite different from one where boiling is present.

Gupta [38] proposed a method for selecting quenching conditions based on the com-
putation of the heat-transfer coefficient for a number of quench baths. The heat-transfer
coefficients were calculated from the solution of the inverse heat conduction problem for
instrumented disks (200 mm-dia. x 20 mm thick) made of 304 stainless steel. To select
the quench bath that produces a part with a specific hardness distribution, the required
thermal history was first obtained from either a C C T diagram or a Jominy curve. This
Literature Review 13

thermal response was imposed at the desired location and a mathematical model to solve
the inverse heat conduction problem was used to compute the required heat-transfer
coefficient vs surface temperature curve. This curve was then compared to the curves
obtained with the instrumented disks to select the appropriate quench bath. The method
was successfully applied to the quenching of a 100 mm-dia. steel grinding ball.

Liscic and Filetin [39] developed software for predicting the hardness distribution
in quenched bars from measured cooling curves. The technique is based on measuring
the temperature at the surface and at an internal point near the surface of cylindrical
specimen instrumented with an specially designed probe. The specimen (50 mm-dia. x 200
mm long) is made of AISI 304 stainless steel to ensure that no heat will be generated
due to phase change. The measured temperature responses are used to compute the
surface heat flux by directly applying Fourier's law, with the conductivity taken as a
constant. The quenching conditions are then described by three functions : T s u r f = f(t),
q = F(t) and q = f(T ).surt A test specimen made of the steel of interest needs then
to be quenched and its hardness distribution determined; this data is transformed into
the equivalent Jominy curve. With this database, the hardness distribution in a bar of
the diameter of interest can be computed by using regression equations from a series of
Craft-Lamont diagrams. Liscic et al. [40] suggested a qualitative description of depth of
hardening and risk of cracking based on the method described above. For this purpose
the maximum flux density (q ) and the heat extracted (as calculated by integrating
max

// q dt) must be calculated from the q = f{T ^) and q = F(t) functions, respectively.
suT

For two given quench baths, the one that results in a higher value of the integral of
surface heat flux with respect to time will produce a greater depth of hardening (since it
corresponds to a higher amount of heat extracted). On the other hand, the quench bath
with the lower value of surface temperature corresponding to a maximum heat flux at
later times will have a higher risk of cracking (since the largest amount of heat would be
extracted at temperatures near or below M J .
Literature Review 14

Wetting kinematics measurements have also been proposed as the basis for hardness
predictions [41]. Wetting time is the time required for the wetting front to arrive at a
given location in the specimen. Since the wetting time irepresents the heat extraction
at the surface, the same hardness is expected in locations having the same wetting time.
The method requires a calibration curve of hardness versus wetting time obtained under
laboratory conditions. After measuring the wetting time under plant conditions the
hardness can be predicted based on the calibration curve. A shortcoming of this method
is the experimental difficulties associated with measuring the wetting time under plant
conditions. Also, the calibration curves are unique to the geometry used to determine
them and cannot be easily extended to other geometries.

2.1.3 Process M o d e l i n g

None of the predictors previously mentioned includes a careful simulation of thermal


histories but rely on either average values of heat transfer conditions or correlations
based on cooling rates at particular stages. Some other simplified assumptions need to
be made in order to keep the computational requirements to a minimum. In contrast, a
process modeling approach can provide detailed thermal, microstructural and mechanical
responses obtained during a particular heat treatment. The methodology involves the so-
lution (through numerical methods) of the partial differential equations that describe the
evolution of the thermal and stress fields. A relatively large data bank containing ther-
mophysical and thermomechanical properties, as well as transformation kinetics, must
be available. Laboratory-scale simulations are usually required to verify the different
mathematical models involved. A vast amount of information regarding the flow of heat
as well as microstructural and mechanical information at all stages of the heat treatment
is generated but, due to the high non-linearities involved in real problems, extensive
computational resources are required. In particularly complex cases, efforts should be
Literature Review 15

devoted towards implementing a practical way to visualize the numerical results. A crit-
ical component of the models is a detailed heat transfer study to characterize the active
boundary condition. Once the microstructural distribution is computed, the mechanical
properties can be predicted from empirical correlations. In this fashion, mathematical
models can be applied to the prediction of mechanical properties from information of
process variables. Crack formation and residual stress distribution can also be predicted
if a stress model is available. The process modeling approach was applied by Campbell
et al. [42-44] to the simulation of controlled cooling of steel rods in a Stelmor line. A
similar approach was adopted by Wallis et al. [45] to optimize the heat treatment of
superalloy forgings and by Persampieri et al. [46] to simulate water spray quenching of a
flat disc and a hollow cylinder of AISI 4335V steel. Another example of the use of this
methodology, as applied to water spray quenching of steel bars, has been presented by
Nagasaka et al. [47]. The commercial package H E A R T S is also built on the microstruc-
tural engineering concept and has been used to model carburized quenching of a gear
tooth [48].
A detailed discussion of published mathematical models developed to simulate mi-
crostructural evolution and stress generation during heat treatments is presented in a
later section.

2.1.4 Computerized Information Systems

The simulation models developed, based on first principles, can be complicated to run and
may require expert knowledge. They can also be expensive. For these reasons, there has
been a strong interest in developing PC-based software for material selection, property
prediction, and design of heat treating operations. A series of computerized information
systems for steel selection and prediction of mechanical properties have been reported
in the literature [49-52]. One of these systems [52] also includes access to data bases
of several steel standards and methods for heat treatments. A comprehensive review
Literature Review 16

of the application of computerized information systems can be found in [53]. Bodin


and Segerberg [54] developed a benchmark testing procedure to evaluate commercially
available, PC-based programs for heat treatment of steel components. They tested three
programs against each other and experimental measurements of quenched cylindrical
specimens of steels of low and high hardenability and found some discrepancies, which
they attributed to differences in the characterization of heat extraction.

2.2 Heat Transfer in Forced Convective Quenching

The resulting hardness profile and residual stress distribution in a heat treated part
depends primarily on cooling rate history. Therefore, characterizing the rate of heat
removed from the surface by the quenching medium is of vital importance. But heat
extraction during quenching is a complicated process due to the presence of boiling. In
this section, the general characteristics of heat transfer, including surface boiling, are
presented, followed by a review of techniques for characterizing surface heat flow. Then,
specific studies of forced convective boiling are reviewed.

2.2.1 General

By recording the temperature response within instrumented specimens and simultane-


ously filming the reactions at the surface, a number of researchers [55-57] have established
the different modes of heat transfer during quenching of metallic parts in typical quench-
ing media. When a component at high temperature is put in contact with a vaporisable
liquid, three distinct stages of cooling can be distinguished from the surface temperature
response (see Figure 2.2). After a short period characterized by intense boiling, the first
stage, known as film boiling, starts. During film boiling a blanket of vapour surrounding
the entire piece is formed, thus creating a barrier for heat transfer; the corresponding low
rate of heat extraction is reflected in a small rate of change in the temperature response
Literature Review 17

in the cooling curve. As the surface temperature drops, the amount of heat reaching
the liquid decreases resulting in a partial collapse of the vapour blanket. During this
transition boiling regime, the breaking up of the vapour blanket provides fresh sites for
metal/liquid contact and, therefore, the heat extraction rate increases. With a further
reduction of temperature, smaller bubbles and an increased metal/liquid contact occur,
which are characteristic of the nucleate boiling stage. This is the most efficient mode
of heat extraction during quenching and results in the maximum surface heat flux. For
surface temperatures near the boiling point of the quenchant and below, boiling stops
and heat extraction is entirely controlled by either natural or forced convection.

In selecting a quenchant for a given application, the complete history of rate of heat
extraction at the surface needs to be considered. Ideally, to minimize thermal stress and
optimize the microstructure required, a quenching medium would produce low cooling
rates at high temperatures, high cooling rates at temperatures where phase transfor-
mations occur (i.e., between A3 and M ), and low cooling rates at the final stages of
s

cooling [58]. Generally speaking, water and brine (aqueous solutions of sodium or cal-
cium chloride) produce the highest cooling rates at virtually all stages of cooling; while
this may result in higher hardnesses, it is also a potential source of severe distortion and
even crack generation. On the other hand, oils produce more uniform, albeit slower, heat
extraction rates as compared to water or brine and are favoured in a large number of
applications. Recently, polymeric aqueous solutions have received increasing attention
due to the variety of heat extraction rates that are possible to obtain with them. Obvi-
ously, the selection of the most appropriate quench bath should also include factors such
as material hardenability, component shape and surface condition, among others.

2.2.2 Surface Heat F l u x Characterization

A cooling curve provides valuable information on the different stages of cooling dur-
ing quenching. However, computer codes designed for the prediction of microstructural
Literature Review 18

and thermomechanical responses in a metallic part require values of either surface heat-
transfer coefficient or surface heat flux, as a function of surface temperature, in order to
characterize the active boundary condition. There are three possible ways to accomplish
this task :

1. Indirect measurement of the surface heat flux.

2. Direct measurement of the surface heat flux.

3. Prediction of the surface heat flux.

Diller [59] has reviewed techniques developed for heat flux measurements and sub-
classified the category of indirect measurement of surface heat flux according to :

1. Measurements based on spatial temperature difference. A temperature difference


in the specimen is measured across a known thermal resistance gauge. The major
advantage in using these gauges resides in the fact that the output signals are
proportional to the surface heat flux.

2. Measurements based on temperature change with time. The thermal response in


the specimen is measured over time with a known thermal capacitance gauge. Given
that only a single temperature measurement is required in this case, these gauges
are the simplest ones to fabricate. Most of the effort is concentrated on recovering
the heat flux - usually through the solution of the inverse heat conduction problem . 2

3. Measurements based on spatial temperature difference in the fluid. The temper-


ature gradient in the fluid next to the surface is measured. This approach also
involves the accurate knowledge of fluid properties as well as fluid flow characteri-
zation and, therefore, it has not been widely used.

2
A detailed discussion on the inverse heat conduction problem is presented in Chapter 5.
Literature Review 19

On the other hand, a direct measurement of the energy leaving or reaching a surface
at steady or quasi-steady state conditions can provide a very accurate measure of the sur-
face heat flux, as long as thermal equilibrium at all temperatures exists. Bamberger and
Prinz [60] have pointed out that the rapid change of heat flux with surface temperature in
transition boiling makes it impossible to attain thermal equilibrium and, therefore, pre-
cludes the direct measurement of heat flux during this stage of boiling. Given that energy
exchange is usually achieved through electrical heating, which has power limitations, this
technique is not normally used in high-heat flux or high-temperature applications.
For either direct or indirect measurements of the surface heat flux, special attention
should be given to the design of the gauge. The goals of any heat flux gauge are twofold :
i) the flow of fluid over the surface should not be affected by the presence of the gauge, and
ii) the conduction pattern in the solid should not be disrupted. Unfortunately, a gauge
always alters the local distribution of both temperature and heat flux, but this disruption
should be kept to a minimum. Another issue in heat flux measurement is calibration.
Although any mode of heat transfer can be applied to calibrate a given gauge, radiation
is used in most of the cases. Two limitations of the calibration procedures are : 1) Most
gauges are calibrated at room temperature, and 2) Calibration of the transient response
is rarely done.
A n alternative to measure the surface heat flux is to predict it theoretically. How-
ever, this is not an easy task, as it involves a coupled heat and fluid flow analysis of
the quenching bath. Moreover, the boiling that accompanies quenching involves several
modes of heat transfer between the surface of the test piece, the bath and bubbles, as
well as kinetics of bubble formation and growth [61-63].

2.2.3 B o i l i n g H e a t Transfer.

The different modes of boiling are classified according to the hydrodynamics of the bath
and the operational temperature of the fluid with respect to its saturation point. If the
Literature Review 20

liquid is quiescent, the boiling mode is termed pool boiling, whereas in forced convective
boiling, the liquid is set in motion by external forces. It should be noted that bubble
dynamics induce mixing near the surface of the test piece in both modes. When the
operational temperature of the liquid is kept below its saturation point subcooled boil-
ing occurs; on the other hand, in the case of saturated boiling, the liquid is kept at a
temperature slightly above saturation.

The bulk of investigations on boiling have been concentrated on saturated pool boil-
ing. In an early study, Nukiyama [64] devised a power-controlled experiment to study the
boiling behaviour of saturated water at atmospheric pressure. He identified the different
stages of boiling on heating and summarized them in a boiling curve : this is a plot of
surface heat flux vs excess temperature (the difference between surface and saturation
temperature), as shown in Figure 2.3. Two points of particular interest in a boiling
curve are the critical heat flux (maximum heat flux and lower temperature bound of the
transition region) and the Leidenfrost point (minimum heat flux and upper temperature
bound of the transition region). The correspondence between the surface temperature
response upon cooling in subcooled pool boiling and the boiling curve is schematically
shown in Figure 2.4 [56].
When boiling is present, Newton's law of cooling takes the following form :

1 = HT -T J
suri s (2.1)

where q is the heat flux, T g a t is the saturation temperature of the fluid and h is the heat
transfer coefficient which is a strong function of the surface temperature, the geometrical
configuration of the test piece, fluid properties and the fluid flow conditions.

Clearly, by inspecting the boiling curve, it is not possible to obtain a single correlation
to encompass all modes of heat transfer with boiling. Instead, most of the studies have
dealt with separate sections of the boiling curve. For example, the heat flux for nucleate
pool boiling on clean surfaces can be predicted using an empirical correlation developed
Literature Review 21

by Rohsenow [65], which includes the effect of surface tension as well as the particular
combination of fluid-surface. On the other hand, film pool boiling correlations have
been developed based on experiments on laminar film condensation. Due to the higher
temperatures associated with film boiling, radiation effects need to be incorporated. The
correlations developed for boiling heat transfer can be classified according to [66] :

1. Correlations of an empirical nature, which make no assumptions about the mech-


anisms of heat transfer, but solely attempt a functional relation between the heat-
transfer coefficient and independent variables.

2. Correlations that recognize that departure from a thermodynamic equilibrium con-


dition can occur and attempt to calculate the actual vapor quality and vapor tem-
perature.

3. Semi-theoretical models, where attempts are made to write equations for the indi-
vidual hydrodynamic and heat transfer processes in the heated channel and relate
them to the heated wall temperature.

Hewitt [67] noted that most correlations developed for boiling are of an empirical nature,
although correlations obtained for pool boiling have, in general, a sounder physical basis;
but even in that number of empirical factors have to be introduced.

A n issue that has received a good deal of attention is the applicability of correlations
obtained under steady-state conditions to applications where transient behaviour occurs
[68-72]. While steady-state boiling curves are commonly obtained by controlled heating
of the test piece, transient measurements are usually conducted by quenching and lend
themselves to obtaining a complete boiling curve - which is not always the case with
steady-state measurements involving fluids having relatively high boiling points (like
water). Bergles and Thompson [68] performed steady-state and transient experiments
with saturated Freon-113, saturated water, and liquid nitrogen, to establish possible
discrepancies between steady-state pool boiling data and quench data. For the water
Literature Review 22

tests, they found several differences between the two kinds of experiments. In particular,
the peak heat flux was characterized by lower values and higher wall superheats in the
transient experiments. The minimum heat flux was also shifted to considerably lower
values. Similar observations were made with Freon and nitrogen. Unfortunately, the
steady-state measurements were conducted with stainless steel, whereas copper was used
for the transient ones. Peyayopanakul and Westwater [69] conducted similar experiments,
but removed the uncertainty caused by using different metals by testing the circular end of
a 5.08 cm copper rod in liquid nitrogen at atmospheric pressure in both, steady-state and
transient experiments. Several quenching times were obtained by varying the thickness
of the specimen. For the fastest quenches (thinner specimens) a departure from steady-
state conditions was observed. A minimum thickness of the copper specimen (in this case
25 mm) was needed to obtain a complete boiling curve equivalent to the steady-state
one. A n interesting observation was that different minimum thicknesses were required to
obtain quasi-steady-state responses for the different regions of the boiling curve. Huang
et al. [70] obtained transient and steady-state boiling curves under forced upflow of
water in a hollow copper cylinder using water as the fluid. To minimize the possible
influence of external factors, they used the same test section and surface condition for
all tests, and devised an in-situ transient calibration of the instrumented specimen. A
single data reduction method was used in all cases. In the range of pressure < 7 bar,
mass flow rate < 300 kg m - 2
s _1
and subcooling < 15 K , good agreement between
steady-state and transient boiling curves was observed. Again, steady-state and transient
experiments affected each stage of the boiling curve differently. The system response in
film boiling was very similar for both modes throughout the parameter range investigated,
possibly due to the low cooling rates experienced in that stage. For the high end of
the parameter range investigated, the minimum film boiling temperature was lower for
the transient experiments. The quench data was consistently lower in the transition
region of the boiling curve, while no difference was found in the nucleate stage. Ishigai
Literature Review 23

et al. [73] measured the surface heat flux, as a function of surface temperature, near
the stagnation point of a plane water jet impinging on a flat surface using steady-state
and transient measurements. The samples were made of stainless steel and instrumented
with a thermocouple welded to the center of the back face. Their results showed slightly
lower values of heat flux in the film boiling region when steady-state measurements
were performed. In the nucleate boiling stage, the boiling curve determined by the
transient experiment was shifted to higher temperatures. It should be pointed out that
the results in the nucleate boiling region of the boiling curve deduced from the steady-
state experiments lay on the extension of the nucleate pool boiling curve.

Other characteristics of the experimental set-ups that may affect the shape of the
boiling curve are [61,62,74] : surface finish, surface contamination, presence of noncon-
densible gases, heater thickness, heater material, method of heating, and flow velocity.

2.2.4 Forced C o n v e c t i v e Q u e n c h i n g

The need for higher cooling rates than those provided by pool boiling has prompted the
development of more efficient cooling schemes using forced convective boiling. Spray
cooling is used in 'press quenching' of aluminum alloys because it enhances the heat
transfer rate in all stages of cooling and increases the Leidenfrost temperature (which
could possibly eliminate the film boiling regime entirely) [75]. Impinging liquid jets have
been widely used either in submerged or free surface mode; Wolf et al. [74] have recently
reviewed the literature available on jet impingement boiling. A number of studies have
been conducted on forced flow of boiling helium, which is commonly used for cooling
superconducting magnets [76]. Water jet cooling is applied in run-out tables to obtain
high heat extraction rates; this accelerated cooling refines the ferrite grain size, producing
steels of higher strength [77]. During continuous casting of steel, heat exchange between
the copper mould and the cooling water takes place by forced convection. Under ideal
operating conditions, the temperature of the flowing water is kept well below its boiling
Literature Review 24

point; but, if the mould is hot enough, boiling may occur which can lead to defects in the
casting. Samarasekera and Brimacombe [78] computed boiling behaviour diagrams (plots
of water velocity vs water pressure, showing areas of boiling and no boiling) based on
published correlations for fully developed pool boiling and incipient boiling, and used the
results to show that intermittent boiling could occur during the casting of high-carbon
steel grades. Several studies have been conducted on the heat transfer during a hypothet-
ical loss-of-coolant accident (LOCA) in nuclear power plants [79-85]. Both, reflooding
(low pressure) and recovery (high pressure) stages of L O C A s have been considered. The
hydrodynamic conditions investigated are falling film of water on the outside surface
(relevant to B W R and S G H W R spray cooling) and upflow of water (relevant to B W R
and P W R core reflooding). It should be noted that most of the studies related to heat
transfer in L O C A s have been conducted at low mass fluxes (< 2000 kg m - 2
s ) and
_1

relatively low surface temperatures (< 500 C)


While fluid flow in pool boiling is primarily driven by bubble motion, bulk motion
as well as buoyancy effects are responsible for fluid flow in forced convective boiling [86].
This results in higher rates of heat extraction (that increase with fluid velocity and
subcooling) as shown, schematically, in Figure 2.5 [87]. Near the critical heat flux the
forced convective boiling curves for various fluid velocities and subcoolings merge into a
single curve known as the fully developed boiling curve. In some systems, this curve lies
on an extension of the corresponding nucleate boiling curve for pool boiling.

During forced convective boiling a fraction of energy is transported from the heated
wall to the liquid through turbulent forced convection, as in single-phase convection. In
addition, mechanisms are available that increase heat transport by : 1) creating addi-
tional near-wall convection (therefore, augmenting single-phase forced convection) or 2)
providing additional heat transfer routes.

Vandervort et al. [88] studied the mechanisms for heat transfer in very high heat
flux ( > 10 W m ) subcooled boiling that can be achieved through subcooled forced
7 - 2
Literature Review 25

convective boiling of water. They focused on phase distribution measurements at high


mass fluxes (> 5000 kg m - 2
s ) inside small diameter tubes (0.3 - 2.5 mm-dia.) and
_1

its relation to heat transfer, in particular C H F . It was observed that the maximum
heat flux attainable increases with mass flux rate and decreases with tube diameter,
with q > 10 W m
8 - 2
for a flow of 40,000 kg m - 2
s _1
in a tube of 0.3 mm diameter.
The accompanying pressure drop suggested that the vapor fraction is low compared
with typical values at lower velocities and subcoolings and this was confirmed by visual
inspection. They also presented an analysis of forces acting on bubbles at the wall and
on departed bubbles. At high heat fluxes, Marangoni forces are likely to dominate,
since very large thermal gradients exists near the wall. These large gradients are also
responsible for bubble condensation occurring near the wall since the superheated layer
extends only a few microns from the wall. For their particular configuration, Vandervort
et al. [88] found the bubbles to grow to diameters up to approximately 3 microns and form
a bubble boundary layer of approximately 10 microns, i.e., only a few bubble diameters.
A number of mechanisms for heat transport were also discussed : when boiling occurs,
additional near-wall turbulent motion is created by bubble growth and detachment, and
new mechanisms for heat transfer are provided by superheated liquid vaporization, liquid
quenching of exposed surface, bubble-induced microconvection and latent heat transport.

Recently, Celata et al. [89] assessed the applicability of existing correlations and mod-
els to predict the critical heat flux in subcooled forced convective boiling in the range of
interest of fusion reactors (0.1 < p < 8.4 M P a , 0.3 < D < 25.4 mm, 0.1 < L < 0.61 m,
2 < G < 90 M g m - 2
s , 90 < A T
_1
s u b < 230 K ) . They noted that a major obstacle in
establishing the applicability of these kinds of correlations is the scarcity of experimental
data available.
Literature Review 26

2.3 Microstructural Evolution

The prediction of microstructural evolution constitutes the essence of any effort to model
heat treatments because the final microstructure will define the mechanical properties in
the heat treated material. A vast amount of information regarding isothermal transfor-
mations in steels, both empirical and fundamental, has been generated over the years;
however, for practical industrial heat treatment applications, the kinetics of the trans-
formations under continuous cooling/heating conditions are required. In this section,
general characteristics of phase transformations occurring during heat treatment of steels
are first presented. Then, the kinetic equations describing isothermal transformations in
metals are discussed, followed by their implementation to continuous cooling processes.
Finally, the effect of stresses on transformation kinetics is presented.

2.3.1 General

In general terms, solid-solid phase transformations in alloys can be classified, based on


nucleation and growth mechanisms, as 1) thermally activated and 2) athermal. The
former are also known as nucleation and growth reactions and are controlled by diffusional
mechanisms in which atoms are added to an original embryo via random atom motion.
The transformation rate is dependent on temperature (with the rate increasing with
undercooling below the equilibrium temperature for the transformation); and the fraction
transformed varies with time spent at a given temperature. On the other hand, athermal
transformations are characterized by a coordinated atom movement. The transformation
interface moves at the order of the speed of sound in solids and, consequently, the fraction
transformed does not depend on time spent at a temperature, but solely on temperature.
Literature Review 27

2.3.2 Isothermal Phase Transformation Kinetics

A schematic plot of fraction transformed as a function of time spent at constant temper-


ature for a nucleation and growth transformation is shown in Figure 2.6. The transfor-
mation curve takes the typical sigmoidal shape. It can be readily appreciated that the
transformation does not start immediately, but that there exists a finite incubation pe-
riod. Also, the rate of transformation (the derivative of fraction transformed with time)
increases rapidly in the early stages of transformation, then reaches a nearly constant
value and, finally, decreases to zero. There are two factors contributing to the reduction
of the growth rate : 1) at the late stages of the transformation the diffusion fields overlap
causing a reduction in the carbon gradient at the interface, diminishing the driving force
for diffusion (this phenomenom is known as soft impingement and is characteristic of
long range diffusion-controlled transformations) and 2) the growth of a region is stopped
when it reaches the growing interface of another region (i.e., hard impingement). Because
of these characteristics of thermally activated transformations, their reaction kinetics at
constant temperature cannot be described by simple first-order reaction models.
Johnson and Mehl [90] incorporated hard impingement into a kinetic equation by
defining the so-called extended volume. Following this assumption, the transformation is
allowed to occur as if nucleation and growth could take place in the untransformed as
well as in the already transformed regions. With this treatment it is possible to separate
the geometrical aspect of hard impingement. B y making the following assumptions :

nucleation occurs randomly;

the transformation is isotropic and, therefore, the product has a perfectly spherical

shape;

nucleation rate as well as growth rate are constant; and

an increment of extended volume is formed by totally random nucleation,


Literature Review

the Johnson-Mehl equation is obtained through integration of a differential increment in


volume transformed :
X = 1 exp (2.2)
3

Avrami [91-93] extended this treatment to include a variable nucleation rate based on
the number of nucleation sites available. For the case of most of the nucleation sites being
consumed in the early stages of the transformation (early site saturation), the following
equation was obtained :-

X = 1 - exp ( - y N G t^j
0
s
(2.3)

This equation can be generalized to :

X = 1 - exp(-6i ) n
(2.4)

where b is related to the magnitude of the nucleation and growth rates and is a func-
tion of temperature, and n is nearly independent of temperature and represents geom-
etry (dimensionality) and growth mode (linear, etc.). This expression is known as the
Johnson-Mehl-Avrami-Kolmogorov ( J M A K ) [90-94] equation and is extensively used for
predicting isothermal phase transformation kinetics in steels. The parameters n and b are
deduced from experimental measurements of dilation as a function of time at constant
temperature.

Cahn [95] derived rate laws describing transformations in which nucleation takes
place on grain boundary surfaces, grain edges, or grain corners. He assumed steady-state
conditions, i.e., growth rate and nucleation rate were independent of time, and considered
the case where only one kind of site was active. Furthermore, the new phase was assumed
to grow with constant radial velocity until it impinged on another growing area. Cahn
adopted the extended volume fraction concept proposed by Johnson and Mehl [90], and
defined an extended volume fraction (for grain corner nucleation), an extended area
fraction (for grain boundary nucleation) and an extended length fraction (for grain edge
Literature Review 29

nucleation). The results were presented in the form of curves of l o g l o g ( l / l X) vs logi.


At low nucleation rates, the rate laws approached the form of Eq. (2.2).

In early studies, published isothermal transformation kinetics were obtained under


conditions that did not warrant a careful characterization of the heat exchange between
the sample and the cooling medium. When this kind of information was applied in
a mathematical model of quenching of eutectoid steel rods, poor agreement between
measured and predicted centerline temperatures was found [96]. Recognizing that a more
careful characterization of the kinetics of the transformations was needed, Hawbolt et al.
performed detailed studies on the austenite-to-pearlite transformation in eutectoid carbon
steel [97] and the austenite-to-ferrite and austenite-to-pearlite transformations in 1025
carbon steel [98]. To accomplish this task, an apparatus based on a diametral dilatometer
was designed to measure temperature and dilation simultaneously on resistively-heated
tubular samples. For the eutectoid steel, the parameter n was found to be nearly constant
when the experimental data was fitted to the J M A K equation obtained by setting the
transformation start time, t = 0, at the start of the transformation (tAv)\ but n varied
widely when t = 0 was set to the time to reach T Al The reason for this discrepancy is that
the J M A K equation was developed without any consideration of the incubation period
and, therefore, the correct procedure to determine the kinetic constants n and b should
only consider the transformation event. The effect of varying the cooling rate to reach the
isothermal transformation temperature on the transformation start time was also studied
and found to be small : an increase of only 7.5 s (from 14 to 21.5 s) was obtained for
a change in cooling rate from 108 C s _1
to 14.65 C s . Similar conclusions regarding
_1

the calculation of n and b were drawn from the study of transformations in the 1025
carbon steel. By comparing the thermal response of a thin wafer specimen heat treated
in a molten salt bath with the results obtained from the resistance-heated dilatometer,
these investigators found that, at the desired test temperatures, isothermal conditions
were not attained quickly in the salt bath.
Literature Review 30

The prior-austenite grain size plays an important role in the kinetics of phase trans-
formations. In general, the rate of transformation is inversely proportional to the grain
size. This effect can be modelled using a modified form of the Avrami equation proposed
by Umemoto et al. [19] :
(-bt \
n

X(0 = l - e x p ^ J (2.5)
where d is the prior-austenite grain diameter and the exponent m takes values from 0 to
3 depending on the geometry of the nucleation site.
In contrast to nucleation and growth reactions, the kinetics of athermal transforma-
tions do not depend on time spent at a given temperature, but on temperature alone.
The evolution of martensite fraction, as a function of temperature, is schematically shown
in Figure 2.7. For the case of martensitic transformations in steels, an empirical equation
obtained by Koistinen and Marburger [99] has been found to predict the kinetics of the
transformation adequately :
X = e x p [ A ( M - r)]
s (2.6)

where A = 0.011 C _ 1
for carbon steels and M is the composition sensitive martensite
s

start temperature.
Information relating the temperature and time required for the start and end of vari-
ous phase transformations for a particular steel is conveniently summarized in isothermal
transformation (IT) diagrams. Unfortunately, the incorrect assumption of instantaneous
attainment of bath temperature in salt bath tests and the fact that the vast majority
of the heat treatments involve continuous heating and cooling limit the usefulness of IT
diagrams to such heat treatments as isothermal annealing, austempering and martem-
pering [22] (which usually involve salt bath processing).

A characteristic feature of IT diagrams is the ' C shape of the iso-transformation


contours for reactions controlled by nucleation and growth. Assuming that the velocity
Literature Review 31

of phase-boundary propagation for nucleation and growth reactions in steels can be de-
scribed based on the diffusion of carbon atoms across the boundary, Zener [100] derived
the following relationship between the rate of advance of the boundary and the diffusivity
and diffusion distance :

f^th b ^ d
a 1 1 0 6
^ (^^ff ^
118 011
distance)(diffusivity) (2.7)

where the diffusivity is inversely related to temperature through an Arrhenius-type equa-


tion and the diffusion distance is inversely proportional to undercooling. This relation-
ship provides a basis for a qualitative explanation of the characteristic C-shape of IT
diagrams. Slow kinetics are obtained at high temperatures (low undercoolings) where
diffusivity is high but the diffusion distance is large; fast kinetics occur at intermediate
temperatures; and slow kinetics appear again at low temperatures (high undercoolings)
where the diffusion distance is small but the diffusivity is low.

Based on classical nucleation theory concepts, Russell [101] obtained a dependence of


the rate of establishing steady-state nucleation on driving force and mobility :

rate of establishing (driving force) (mobility)


. oc ( -o)
2

steady-state nucleation T

The driving force increases with undercooling (decreases with increasing temperature)
while the mobility exhibits the opposite behaviour. Assuming that the rate of estab-
lishing steady-state nucleation is inversely proportional to the incubation time, Eq. (2.8)
predicts large incubation times at high and low temperatures and short incubation times
at intermediate temperatures, i.e., a C-shape for the transformation start time with a
diffusion controlled transformation.

Empirical relationships to compute IT diagrams have been reported by Sakamoto


Literature Review 32

et al. [102]. A more fundamental approach has been adopted by Kirkaldy and co-workers
[14,103,104] to derive semi-empirical formulas for predicting IT curves based on Zener-
type equations.

2.3.3 Continuous C o o l i n g Phase Transformation Kinetics

Phase transformations during most heat treatments occur under continuous cooling con-
ditions. Experimentally determined times to reach various stages of transformation under
non-isothermal conditions are presented in continuous cooling transformation (CCT) dia-
grams. Typically, they can be obtained from dilatometer records of length vs temperature
that are verified through metallographic analysis [20]. The hardness value at the end of
the cooling path is also reported. A major drawback of these diagrams is that they are
strictly valid only for the specific cooling conditions and the austenite grain size used
for their experimental determination. Even though they provide a means of visualiz-
ing a great deal of information at a glance, their graphical nature makes it difficult to
incorporate the data into a computerized system.
The fact that phase transformation kinetics under continuous cooling conditions de-
pend not only on steel composition and prior-austenite grain size but also on the cooling
path, makes the task of predicting the microstructural evolution during heat treatments
a difficult one. Since the cooling rate in a part varies from point to point, deriving
the necessary number of empirical relationships between phase distribution and cooling
path would require a very large number of measurements, specially for a part of complex
geometry. Thus, a more general approach is desired.

The first step in modeling microstructural evolution involves the prediction of the
incubation time. A common practice consists of obtaining empirical correlations between
incubation time and cooling rate [43]. However, this method cannot be generalized easily
and, therefore, several researchers have attempted to develop methods to predict the
start of austenite decomposition for any arbitrary cooling path.
Literature Review 33

Scheil [31] proposed that the amount of undercooling and the incubation time are
related. B y assuming-that the continuous cooling path to a given transformation tem-
perature can be discretized, Scheil postulated that a fraction of the isothermal incubation
time is consumed at each 'isothermal' stage and that when the sum of all these fractions
reaches unity, the transformation begins. Thus, if is the time spent at a given 'isother-
mal' stage and Tj is the isothermal incubation time corresponding to that temperature,
the fraction consumed isothermally is given by :

and, upon summation, :

(2.10)

In the limit where A T > 0 and assuming a constant cooling rate :

which is known as the additivity rule or Scheil equation.


In their work described previously, Hawbolt et al. [97, 98] computed the incuba-
tion time for transformations in both hypoeutectoid and eutectoid carbon steels using
Eq. (2.11) and compared it with the experimentally determined start times. They found
that by using the additivity rule the start times were consistently overestimated.

In an attempt to improve upon the ability to predict the transformation start time,
Pham [105] derived 'ideal' T T T curves based on measured transformation kinetics under
3

isothermal and non-isothermal conditions. In the latter case, the inverse of the Scheil's
additivity rule was applied to back-calculate the 'ideal' isothermal start times. Predicted

3
The 'ideal' T T T curve corresponds to a hypothetical infinitely rapid cooling rate to each of the test
temperatures and, therefore, can be defined in terms of the steel composition and austenite grain size
alone.
Literature Review 34

start times, based on the 'ideal' IT curve, for the austenite-to-pearlite transformation in
a eutectoid carbon steel subjected to continuous and semi-continuous (Stelmor) cooling
were in good agreement with measured values. For non-continuous cooling a reasonable
agreement was found, with the predicted times being consistently early.

The most plausible reason for the observed discrepancy between theoretical and mea-
sured incubation times is that real systems do not comply with the implicit assumption of
'proportional consumption' [106], which requires that a given fraction of the incubation
time at a given temperature always corresponds to the same fraction of the total incuba-
tion time under non-isothermal conditions for all temperatures involved. Experimental
evidence supporting this hypothesis has been reported by Moore [107]
As mentioned earlier, transformation kinetics under continuous cooling conditions
cannot be characterized from first principles. A methodology to apply semi-empirical de-
scriptions of isothermal transformations kinetics to a transformation event under continu-
ous cooling conditions has been successfully applied by, among others, the microstructural
engineering research group at U B C [96,108,109]. The calculations during the transfor-
mation stage are based on the assumption that the rate at which the transformation
develops is a function only of the current temperature and fraction transformed, and not
of the history of the microstructural evolution; in other words, this assumption, known
as the additivity principle, proposes that the material does not have structural 'mem-
ory' and thus the continuous transformation can be approximated by a series of additive
fractional transformations.

The application of the additivity principle is illustrated in Figure 2.8. If X\ is the


fraction transformed at (T\,ti), then an additional increment of fraction transformed,
AX, after a time increment, At, (At 2 ^ 1 ) spent at temperature T2, can be computed
by using the kinetic information corresponding to T^. The methodology includes the
evaluation of a virtual (fictitious) time that corresponds to the time required to attain X\,
at T2, then adding the incremental growth at T2 associated with additional incremental
Literature Review 35

time at T . A n expression for this virtual time can be obtained from Eq. (2.4) :
2

(2.12)

where X\ is the fraction transformed at time ti and denotes the inverse of the
function X(t). Thus, the effective time increment adopted for the calculation is A t =
t t\ (instead of t ti).
2 2

It is readily seen that the additivity principle is similar to the additivity rule except
that it is applied to the transformation, as opposed to the incubation period.
The additivity principle not only provides a tool to apply isothermal data to contin-
uous cooling processes, but it also matches the numerical techniques used to solve the
energy equation that describes the thermal field evolution. Irrespective of the details of
the particular technique employed, the continuous cooling curve is always replaced by
a series of small isothermal steps; a large number of such steps results in a more accu-
rate simulation, but at the cost of larger memory requirements and a greater number of
mathematical operations.
Because of the enormous potential of the additivity principle as a tool for applying
isothermal data to the prediction of continuous cooling events, it is very important to
determine the limits of its applicability. Avrami [91] proposed that the ratio of nucle-
ation rate to growth rate (N/G) should be used as a parameter to predict whether the
additivity principle can be applied. The criterion he proposed, called the isokinetic con-
dition, requires that the ratio remains constant over the temperature range of interest.
Cahn [95,110] suggested an expanded set of conditions based on the assumption that
the kinetics of the transformation are controlled by the growth rate. Christian [111] pro-
posed a somewhat different approach considering that the rate of transformation must
be described by the ratio of two independent functions, one dependent on temperature
Literature Review 36

and the other on fraction transformed (i.e. R = J^Q)- More recently, Kuban et al. [112]
conducted an experimental program that included the measurement of nucleation and
growth rates of pearlite in a eutectoid, plain-carbon steel and suggested an 'effective site
saturation' criterion, where early growth is again the controlling mechanism.
There have been few studies concerned with mathematical formalisms of the additivity
principle. Verdi et al. [113] introduced the concept of degree of advancement of the
transformation, [ix-, and showed that the concept of additivity, as applied to the evolution
of phase transformations, holds if the rate of degree of advancement of the transformation
depends on X and T alone.
Due to the nature of the research on both the additivity rule and the additivity
principle, some confusion exists in the literature regarding the nomenclature as well as
the equivalence between different descriptions and conditions for additivity. Recently,
Mukunthan [106] has presented an excellent review aimed at clarifying these points.
Experimental success at using the Avrami equation and applying the additivity prin-
ciple to the austenite-to-ferrite transformation in 1010 and 1020 plain carbon steel, was
reported by Kamat et al. [114]. By conducting stepped isothermal transformation tests,
these researchers clearly showed that the simulated continuous cooling transformation to
ferrite could be adequately described using the additivity principle. Using a fundamental,
carbon diffusion-controlled mathematical model for ferrite growth kinetics, they showed
that the duration of the unsteady-state effects caused by the rapid change of bound-
ary conditions and diffusion coefficient is very short, which would explain the observed
experimental additivity of the proeutectoid ferrite. Since the Avrami equation was not
developed to describe the 'soft impingement' conditions associated with the austenite-
to-ferrite transformation, it is used here as an empirical equation, having the correct
form.

Finally, the bulk of the work that has been done regarding transformation kinetics
has been directed towards the cooling of steels. Since heat treatment involves the initial
Literature Review 37

heating and austenitizing of the steel, a systematic study of the applicability of the addi-
tivity principle to the kinetics of the heating reversion to austenite should be conducted
in order to be able to model a full heat treatment. This has been done by Riehm for a
eutectoid carbon steel and confirms that the reversion can be adequately described using
equations having a similar form [115].

2.3.4 F e - C - X Phase B o u n d a r i e s

The need to accurately model the multicomponent equilibrium Fe-C-X diagram in hard-
enable steels is twofold : 1) typically, parameters in empirical and semi-empirical kinetic
equations are evaluated based on the degree of undercooling, and 2) regression equations
for calculating equilibrium fractions transformed from the lever rule are required in the
case of hypo- and hypereutectoid steels.
Empirical formulae to compute TA , 3 TA I,
C and TA 3 have been obtained by Andrews
C

[116] using a database comprised of published data on 150 steels. It is important to note
that all the formulae show a linear dependence of the respective temperature with the
percentage of alloying elements.
Theoretically-based formulae have been given by Kirkaldy and co-workers [14,104,117]
to predict phase boundaries of the Fe(X,-)-(Fe, X J 3 C multicomponent phase diagram for
carbon contents up to ~ 2 w/o C. The thermodynamic formalism to describe the 7-a
equilibrium is a generalization of the classical 'depression of the freezing point' relation,
applied to the Fe-Fe C diagram. The temperature deviation is derived, from the equality
3

of chemical potentials. The calculations for the carbide equilibria followed a somewhat
different approach due to the fact that there is insufficient knowledge of the solution
thermodynamics of the alloy-enriched cementite. Starting with the equality of chem-
ical potentials, as given by the Gibbs-Duhem equations, the isothermal concentration
deviation of the 7
7+ FeaC boundary is obtained.
Literature Review 38

Once the equilibrium lines have been calculated, the eutectoid composition and tem-
perature in an alloyed steel can be computed from the intersection of the two equilibria.
By comparing predicted and measured equilibria in low-alloy steels [104], the limits of
applicability of this formulation were found as follows : % C < 2 w/o, % M n < 3 w/o, %Si
< 1 w/o, %'Ni < 3 w/o, %Cr < 2.5 w/o, %Mo < 2 w/o, % C u < 3 w/o, others < 1 w/o.

For the prediction of IT curves in hardenable steels, the values of interest are the
ferrite, pearlite, and bainite asymptotes, i.e., the temperatures to which the start curve
of the respective reaction extrapolates at very long times. The ferrite asymptote always
corresponds to the A e temperature [104]. In contrast, the pearlite asymptote depends
3

on the partitioning behaviour of alloying elements. Kirkaldy and Venugopalan [14] have
proposed an interpolation formula to compute the pearlite asymptote based on the cal-
culation of equilibrium (full partitioning) and paraequilibrium (no-partitioning) phase
boundaries. Rigorous thermodynamic formulae for equilibrium, paraequilibrium, and
no-partition local equilibrium have been given by Hashiguchi et al. [118]. No theoret-
ical description of the bainite asymptote has been developed and, therefore, empirical
equations are commonly adopted, as they are for the composition invariant martensite
transformation [116,119].

2.3.5 Effect of Stresses on the Kinetics of Phase Transformations

The influence of external stresses applied prior to phase transformations is well known. In
the case of the martensitic transformation, strain-induced nucleation has been recognized
as an important factor in increasing the M temperature [120-122]. A n increase in the
g

nucleation rate has also been reported for the austenite-to-pearlite transformation when
the austenite has been previously deformed [123].

In contrast, very few studies on the effect of stresses during the transformation are
available. Dubrov [124] conducted a high-temperature metallographic analysis of the
Literature Review 39

bainitic transformation under applied tensile stresses in a 40Kh2N3SGV steel. He moni-


tored the growth of a needles and observed an increased number of nucleation sites when
the transformation developed under stress. The influence of stresses on the kinetics of
the transformation varied with the stage at which they were applied. Violle [125] studied
the effect of a uniaxial compressive stress on the kinetics of the transformation of /3 to
a in U-Cr alloys. For the isothermal case he reported shorter incubation periods and
higher rates of transformation at all temperatures; in the case of continuous cooling, the
applied stress (up to 25 kg m m ) resulted in a shift in the actual transformation temper-
- 2

atures toward higher values. Gautier [126] studied the influence of applied tensile stresses
during the pearlitic transformation in eutectoid carbon steels by measuring dilation and
electrical resistivity. She reported shorter times for the beginning and end of the isother-
mal transformation. This shift in the kinetic curves was a function of both the level
of applied stress and the test temperature. It was concluded that the growth rate was
not significantly modified and, therefore, the transformation kinetics were accelerated
through a change in the nucleation rate alone. It was observed that the nucleation sites
remained unchanged but they were saturated in shorter times. Gautier also investigated
the influence of applied tensile stresses on the kinetics of the martensitic transformation
in a 60 N C D 11 steel. The applied stresses resulted in an increase in the M s tem-
perature. Fernandez et al. [127] reported a similar shift in the transformation kinetic
curves of eutectoid steels subjected to an applied stress during the austenite-to-pearlite
transformation.

Andre et al. [128] measured the hardness distribution in massive and implanted cylin-
ders (34 mm-dia. by 102 mm length) of 60SC7 spring steel quenched in water and oil.
The implanted samples were fabricated by drilling an axial hole in the specimen and
inserting a cylinder made of the same steel in the shaft. Good thermal contact between
the two cylinders was ensured by pouring liquid tin between them. Due to the presence
of liquid tin, the stresses generated in the outer shell were not transmitted to the inner
Literature Review 40

cylinder; also, negligible thermal gradients were generated within the inner cylinder. The
authors concluded that internal stresses generated by quenching resulted in an acceler-
ation in the pearlite and bainite transformations. A review of the influence of internal
stresses on the kinetics of the pearlite transformation during continuous cooling has been
presented by Denis et al. [129].

2.4 Residual Stresses in Heat Treatments

The residual stress distribution generated during the manufacture of engineered compo-
nents, in combination with the loads experienced in service, determines the components'
life. Additionally, mathematical models of stress generation during heat treatments can
only be verified against measured residual stresses. Thus, techniques for measuring resid-
ual stress distributions are a very important tool in structural analysis. In this section,
the concept of residual stress is defined together with current classifications of residual
stresses. Then, the generation of residual stresses during quenching is briefly discussed,
and the techniques for their measurement are reviewed.

2.4.1 General

Residual stresses are those stresses that would exist in a body in the absence of external
loads [130]. They are also referred to as internal stresses, initial stresses, inherent stresses,
reaction stresses, and locked-in stresses. Residual stresses are elastic only and the residual
stress system in a body must be in static equilibrium, i.e., the resultant force and the
resultant moment produced by the system of stresses must be zero.

According to the characteristic dimension involved, residual stresses can be classified


into three subgroups [131] : 1) macrostresses, which vary continuously over macroscopic
dimensions larger than the average grain size, 2) microstresses, which act over dimensions
of the order of a grain dimension and are distributed non-homogeneously on a macro
Literature Review 41

scale, and 3) pseudo-macrostresses, which are the average of microstresses taken over
a representative volume. K i m [132] noted that the difference in residual stress among
phases in a microstructure would be a more meaningful characteristic of residual stresses.
Accordingly, he proposed the following classification : 1) macrostresses, same as above,
2) 'phase macrostresses', which are the average residual stress in a particular phase in a
microstructure , and 3) microstresses, defined as variations of macrostresses and 'phase
4

macrostresses' over small distances.

Residual stresses may be generated during manufacturing or during use of parts, and
are caused by nonhomogeneous partitioning of inelastic strains both at the micro- and
macroscopic scale. The residual stress distribution generated during quenching is due to
thermal and transformation volume changes that interact in a complex manner as the
part is being cooled from the austenitizing temperature. Thermal contraction occurs
at all times, while volume expansion associated with phase transformations occurs only
during the transformation. Clearly, the hardenability of the steel and the heat extraction
characteristics of the quenching medium, as well as the thermophysical properties of the
part, will dictate the final stress profiles. Residual stresses cause distortion and affect
the performance of parts. Distortion is caused by the non-uniform nature of the residual
stress field inside the body. Two examples where residual stresses can affect a structure
adversely are fracture and stress-corrosion cracking. Residual stresses however can be
beneficial as in the case of compressive residual stresses near the surface of shot-peened
parts. The effect of residual stresses on the behaviour of parts is more important at low
levels of applied stress.

4
The weighted average of the 'phase macrostresses' would be identical to the macrostresses.
Literature Review 42

2.4.2 R e s i d u a l Stress M e a s u r e m e n t Techniques

Several techniques can be used to measure residual stresses. They are classified, according
to the principle used [130], in : 1) stress-relaxation techniques, 2) diffraction techniques,
3) techniques using stress sensitive properties, and 4) cracking techniques. Some examples
are given in Table 2.1.
In the stress-relaxation techniques the residual stresses are computed by directly
measuring the strain released by either cutting the specimen into pieces or removing
a piece from the specimen. It is assumed that no further plastic damage is caused
during the stress relaxation (cutting) operation. Despite the fact that stress-relaxation
methods are destructive, they are widely used [133-137] due to the relatively simple
equipment involved. The accuracy and reliability of the measurements require that the
data reduction assumptions are valid. The most commonly used method is the hole-
drilling strain-gauge method described in the A S T M standard E837 [137]. In all cases,
the method is calibrated in a uniform uniaxial stress field of known magnitude.
In contrast, diffraction techniques for residual stress measurements are non-destructive.
The elastic strains associated with the residual stress field are obtained by measuring the
scattering angle for a given plane (or set of planes) when the specimen is irradiated with
X-rays [1,138,139] or neutrons [140-143] and comparing it to a suitable reference angle -
commonly measured in a stress-free specimen. The measurements can be made in a small,
precisely located area, but the techniques are relatively slow. X-rays can only penetrate
small distances in metallic specimens and, therefore, only surface residual stresses can
be determined using X-ray diffraction. On the other hand, neutrons are approximately
a thousand times more penetrating than X-rays [141]; thus, neutron diffraction can be
applied to determine residual stresses throughout the entire part. However, the data
acquisition time for neutron diffraction is longer, due to the loss of beam intensity with
depth of beam penetration.

Rather than measuring strains directly, some methods have been developed based
Literature Review 43

on the measurement of stress-sensitive properties, such as ultrasonic velocity [144] and


Barkhausen noise [145]. The complexity of the interaction of the propagating ultrasonic
wave and a lack of consistency in the performance of transducers are important problems
related to ultrasound measurements [146]. Barkhausen noise analysis can only be applied
to ferromagnetic materials. In addition, it has a narrow range of stress sensitivity, and
is restricted to near-surface probing. None of these techniques are used in field work.
Residual stress distributions can also be characterized through hydrogen-induced and
corrosion cracking [147]; the results are, however, qualitative in nature.

2.5 Mathematical Models of Microstructural Evolution and Stress Genera-

tion in Heat Treating Operations

Mathematical models of heat treatments are made up of two components : 1) a heat


transfer and microstructural evolution module, and 2) a stress generation module. The
majority of the modeling efforts produced so far have treated them independently, solving
for the heat transfer and microstructural evolution first and then using those results as
input to stress generation models. In this section, mathematical models developed to
predict microstructural evolution and stress generation in heat treating operations are
described.

2.5.1 Modeling of Heat Transfer and Microstructural Evolution

The thermal response and microstructural evolution during cooling are linked through
transformation kinetics and the heat evolved during the phase change. Early models
[7,148] did not include the heat of transformation and assumed constant thermophysical
properties. More recent work was based on adopting a modified value of specific heat
in the temperature range where phase transformations take place (see, for example,
Literature Review 44

[149]). This approach-eliminates the need for calculating the microstructural evolution
to compute the rate of heat evolved and, therefore, the calculations are simplified.

Agarwal and Brimacombe [96] modelled the microstructural evolution during water-
quenching of 8.5 mm-dia. eutectoid steel rods. They computed the pearlite fraction
transformed, as a function of time, by adopting an empirical two-parameter equation
(similar to the J M A K equation) and invoking the additivity principle. The parame-
ters in the empirical kinetic equation were estimated from the corresponding isothermal
transformation (IT) diagrams; the transformation was assumed to commence as soon as
the local temperature dropped below T ^ . The numerical procedure adopted to solve
the heat balance equation was based on the finite-difference method. Iyer et al. [108]
improved upon that model by separating the incubation period (as characterized by an
experimentally measured start time) from the transformation event. The kinetic pa-
rameters in the J M A K equation were experimentally determined rather than calculated
from published T T T diagrams. Campbell et al. [43] extended the phase transformation
kinetics database to hypoeutectoid and eutectoid steels and incorporated the results in
a mathematical model of the Stelmor process.
Kumar Singh and Mazumdar [150] compared three methods to model the heat source
upon heating and cooling : 1) the fraction transformed vs time relationship was derived
from IT diagrams, 2) the fraction transformed vs time relationship was derived from the
iron-carbon equilibrium diagram, and 3) the heat source term was assigned a value of
zero and a modified heat capacity was adopted. A control volume-based finite-difference
procedure was implemented for the numerical solution. The computed results were com-
pared against analytical solutions, as well as measured temperature response during
heating, air-cooling of hypoeutectoid steel, and water-quenching of eutectoid steel. The
only transformation considered, even for the hypoeutectoid grade, was from austenite to
pearlite (or vice-versa). For the calculations, a constant density and specific heat were
assumed together with a temperature-dependent thermal conductivity. The boundary
Literature Review 45

condition was characterized by either a combined radiative plus convective heat flux (for
heating and air-cooling) or a constant heat-transfer coefficient (for water-quenching). For
the experimental conditions investigated, the authors found that the procedure adopted
to model the source term was not critical; they attributed this observation to the rel-
atively small contribution of the heat source. However, due to the simplicity of the
computational algorithms used in methods 2 and 3, they recommended using them over
method 1.

The effect of prior-austenite grain size has been included in a recent model of induction
hardening [151] by shifting the time scale of the IT diagram. The incubation time, as well
as the kinetic parameters in the J M A K equation, were calculated by modifying their base
value through a grain size-dependent parameter. For the martensitic transformation, the
M temperature was explicitly given as a function of prior-austenite grain size.
s

Few mathematical models have included the effect of internal stresses on the kinetics
of the transformation. Fernandes et al. [127] recognized the importance of including
this effect and modelled it by using a mean value of the equivalent stress between the
surface and the centre of an infinitely long cylinder. Experimental values of IT curves for
various stress levels were then used to obtain the modified kinetics. A good agreement
between calculated and experimental values was found. Inoue et al. [152] modelled the
kinetics through a Johnson-Mehl type equation modified for the presence of stresses.
Sjostrom [153] used a linear relationship suggested by Denis [154] to account for the
effect of stress on the martensite start temperature.

Regarding the active boundary condition, a significant number of investigators have


adopted a measured surface or near-surface temperature response, while fewer researchers
have estimated the heat-transfer coefficient from either available correlations or through
the solution of the inverse heat transfer problem.
Literature Review 46

2.5.2 M o d e l i n g of Stress G e n e r a t i o n

It has already been stated that volume changes due to thermal and microstructural
responses interact in a complex manner to give rise to transient and residual stresses in
quenched parts. In addition, the level of stresses generated upon cooling are such that
plastic flow always occurs. In those instances where low cooling rates are encountered
(e.g. oil quenching of large specimens), viscous effects may also be important.
The mechanical behaviour of engineered components is commonly defined through
a constitutive material law. This equation describes the relation between stresses and
strains [155] and, together with the equilibrium equations and the compatibility equa-
tions, define the problem. A mechanistic approach, based on crystal plasticity, including
all the interacting micro-mechanisms responsible for plastic deformation of real materials,
is desirable. However, such a formulation is extremely complicated and difficult to apply
in real situations [156]. Instead, virtually all stress analyses are carried out by adopt-
ing the so-called phenomenological (classical) theory of time-independent plasticity [157].
The basic elements of the theory include the assumption of a mathematical continuum,
the existence of a plastic potential, the change in the yield surface with plastic flow as
described by a hardening rule, and the associated flow rule.
A relationship between plastic strain and deviatoric stress has been proposed, both
on a total plastic strain (Hencky's equations) and an incremental plastic strain (Prandtl-
Reuss equations) basis [158]. However, the total plastic strain theory gives inconsistent
results [159] (except in the case of 'proportional loading') and, therefore, all mathematical
models of stress generation are based on the Prandtl-Reuss equations of incremental
plastic strain.

Furthermore, it is possible to separate the incremental strain into its elastic and
plastic components. Although strictly valid only for infinitesimal strain increments, this
assumption holds true for finite strain increments as well. When the nonmechanical
strain increments (such as thermal, transformation, etc.) are explicitly considered, their
Literature Review 47

values can also be computed independently and then added to obtain the total elastic and
plastic strain increments. The general form of the strain increment, de, then becomes :

d
= ^elastic + Aplastic + ^thermal +transf. +
de
(- )
2 13

Early work was devoted to analytical solutions of structural problems and, there-
fore, only the simplest material behaviour, i.e., perfectly elastic-perfectly plastic, was
considered. With the advances in computer technology and the development of suit-
able numerical techniques, constitutive equations that reflect the behaviour of actual
components can now be adopted.
Among those numerical techniques, the finite-element method has gained popularity
as a tool for solving structural problems, due to its ability to handle complex geome-
tries. Two commonly adopted methods to include material nonlinearities are [160] : 1)
the tangent stiffness approach (where the nonlinearities are included in the equilibrium
equations in terms of a 'tangent stiffness'), and 2) the pseudo-forces approach (where the
nonlinearities are included in the equilibrium equations as 'pseudo-forces').
One of the first efforts to model stress generation during quenching is due to Weiner
and Huddleston [161]. They assumed an elastic-perfectly plastic material obeying the
Tresca yield criterion, to compute the transient and residual stresses in solid and hollow
cylinders. A n analytical solution was obtained for a cylinder cooling with a negligible
thermal gradient and a phase transformation that occurred instantly when a critical
temperature was reached.

Early models obviated the need for a detailed description of the microstructural evo-
lution during quenching by using experimentally measured 'effective' thermal expansion
coefficients to characterize a combined thermal-transformation strain. In this formula-
tion, the 'effective' thermal expansion coefficient was assumed to be an explicit function
of either temperature alone [161,162], or temperature and a parameter that character-
izes the cooling rate [7]. Inoue and Raniecki [163] have pointed out that the former
Literature Review 48

approximation would only hold for through-hardened specimens, given that the rate of
transformation in diffusional reactions does not depend on temperature alone.
Fujio et al. [148] modelled the generation of stresses in a 50 mm-dia. 1045 carbon steel
cylinder quenched in water by assuming a constant heat-transfer coefficient at the surface
of the piece. Calculations performed with the chosen value of 5800 W m~ K 2 _ 1
showed
large differences with respect to the measured thermal response 1 mm below the surface
at early times, but a reasonable agreement otherwise. The authors justified their selection
of a constant heat-transfer coefficient by suggesting that the maximum cooling rate is
the most important parameter in describing the austenite-to-martensite transformation.
They did not include the latent heat of transformation in the analysis. The results of the
model were compared with experimentally determined residual stresses obtained using
Sachs' technique. Despite the simplifications, the computed residual stresses agreed well
with the experimentally measured values.
Inoue and Tanaka [7] studied a similar case, i.e., quenching of a 60 mm-dia. bar
made of 0.43 % carbon steel in water. In this investigation, the 'effective' coefficient
of thermal expansion was assumed to vary with temperature and cooling rate, and the
microstructural evolution was estimated from the maximum cooling rate in the part. The
latter assumption resulted in a non-continuous approximation of the phase distribution
during the quenching process. The latent heat of transformation was not included in the
model and the thermophysical properties were assumed to be independent of temperature.
In contrast with the work of Fujio and co-workers, the rate of heat extraction at the
surface was assumed to vary with temperature following the boiling curve. The residual
stress distribution was measured using the Sachs' boring-out technique. Despite having
used a more complex model than Fujio et al. [148], good agreement between predicted
and measured values was again observed.

Inoue and Raniecki [163] presented a more realistic model where the kinetics of the
transformations were taken into account via empirical relationships obtained from IT
Literature Review 49

diagrams. Since the microstructural evolution could be tracked as the quench progressed,
an explicit relationship for the volumetric change accompanying the transformation could
be developed. The model was applied to the quenching of a semi-infinite body; the results
were presented as residual stress and martensite fraction, as a function of distance from
the quenched end, for various cooling rates characterized by a Biot number. Based on
their results, the authors established that the residual stresses would vanish on a plane
containing 30 to 35 % martensite.
A topic that has received a good deal of attention is the modeling of strain hardening
during quenching. Denis et al. [164] simulated the water quenching of a 35 mm-dia.,
60NCD11 steel bar in water at 20 C. For this cooling regime, the only reaction taking
place was the austenite-to-martensite transformation. The authors assumed a perfectly
plastic behaviour within the martensitic transformation interval and an isotropic linear
hardening for the austenitic phase. Isotropic hardening was also assumed by Nagasaka
et al. [47] in their simulation of water spray quenching of 75 mm-dia., 1035 carbon steel
and Ni-Cr alloy steel bars, Buchmayr and Kirkaldy [165] to model the Jominy end-quench
test, and Habraken [166] to simulate a combination of air cooling of beams followed by
water quenching. The case of a 50 mm-dia. 11.6 % N i alloy steel cylinder cooled from
900 C in ice-water was studied by Rammerstorfer et al. [149]. The model allowed for
either pure isotropic or pure kinematic strain hardening; the case of perfectly plastic
behaviour was also included for comparison. The computed results showed a better
agreement with measured residual stresses at the centre of the cylinder when kinematic
hardening was adopted to model strain hardening; the calculated values at the surface
were very similar in both cases. Sjostrom [153] modelled the quenching of a 60 mm-dia.
1050 steel cylinder in water at 20 C and a 50 mm-dia. 11% N i alloy steel cylinder
in ice-water. He found that the choice between isotropic and kinematic hardening was
not critical for the 1050 steel, whereas kinematic hardening gave better agreement with
measured residual stresses for the alloyed steel. Other models [167-169] have included a
Literature Review 50

mixed (isotropic-kinematic) linear hardening rule.


The related topic of loss of memory of prior deformation through phase transformation
was also investigated by Sjostrom in the above-mentioned work [153]. Again, the results
were material-dependent, with the refinement being relevant only in the case of the
through-hardened specimen.

The stress and microstructural fields interact in two ways : 1) the stress field inside
a component modifies the transformation kinetics, and 2) the transformation induces
the so-called transformation plasticity strain. Reference has previously been made to
the treatment of stress-modified kinetics of diffusional and martensitic transformations
and their inclusion in mathematical models (see subsection 2.3.5). A description of
transformation plasticity and its numerical implementation follows.
It has been experimentally observed that macroscopic plastic flow can occur during
solid-solid phase transformations (of both pure metals and alloys) for externally applied
stresses well below the yield stress of the phases involved when an internal stress state
exists in the material. This macroscopic behaviour is the result of localized microscopic
plastic flow and is known as transformation plasticity.
Early researchers had suggested that a transient loss of cohesion between atoms while
they move to new positions in the lattice during the transformation was responsible for
this loss of strength, but more modern investigations suggest two basic mechanisms : 1)
the weaker of the two phases flows plastically in order to accommodate the strain gen-
erated by the transformation front [170], and 2) the product phase orients preferentially
in order to minimize the total energy of the system [171]. The first mechanism was pro-
posed for transformations that involve a significant change in density while the latter was
introduced to explain microplasticity results observed in martensitic transformations.

Several quantitative models have been proposed to incorporate transformation plas-


ticity in calculations of internal stresses during heat treatments. In general terms, they
fall into two categories : 1) adopting an artificially lowered elastic limit during the time
Literature Review 51

the transformation takes place, and 2) introducing an additional strain increment term
which needs to be determined experimentally. However, it should be noted that a lower-
ing of the yield strength implies that the plastic deformation can increase indefinitely and
independently of the phase distribution; also, there is not enough experimental evidence
to support this treatment.

The majority of the expressions to evaluate transformation plasticity in mathematical


models have been developed based on the Greenwood and Johnson mechanism [170].
Their original equation suggested a linear relationship existed between the transformation
plasticity strain, e , and the applied stress :
tp

(2-14)

Abrassart [172] modified this equation to include the amount of phase transformed :

(2.15)

Note that for /, = 1

which is very similar to Eq. (2.14).

From dilatometric experiments under uniaxial constant stress, Desalos and Guinsberg
[173] have proposed the following model :

dfi
3K{l-f )s (2.17)
dt
z

where / ; is fraction transformed and Sij is the deviatoric component of the stress tensor.
Giusti [174] and Leblond [175] derived the following theoretical expression for the time
Literature Review 52

derivative of the transformation plasticity strain under a multiaxial state of stresses :

# = K'sn gUi) f (2-18)

where g(fi) is a normalized function of the proportion of phase transformed.


Leblond et al. [176] have given the following heuristic arguments that justify the
previous relationship :

the equation for e will be incremental, i.e., it will give k .


tp tp

e tp
should be proportional to /, and must be zero for /, = 0 (since transformation
plasticity occurs only if there is a transformation).

e should be proportional to a.
tp

i ip
should not involve volume variation, and should thus be related to the stress
deviatoric component.

In contrast, very little attention has been given to the mechanism put forward by

Magge, whose equation is now given :

etp i 0

iJ{0)dO
de
sin 20 + ^ e ( l + cos 26) (2.19)
Je

The vast majority of the work on modeling the mechanical response during cooling of
metallic components has been based on elasto-plastic formulations. However, few studies
have been conducted using an elasto-viscoplastic approach to model stress generation in
casting [177-179] and quenching [180]. One advantage of elasto-viscoplastic models is
that they can be used to describe rate-dependent phenomena, such as creep. Also, the
steady-state solution of the viscoplastic problem is identical to the corresponding static
elasto-plastic solution [181] and, therefore, efficient codes may be written for certain ap-
plications. A general formulation for the solution of the elasto-viscoplastic problem can
Literature Review 53

be found in Zienkiewicz and Cormeau [178]. The main justification given by the authors
for adopting this methodology is that, experimentally, creep and plasticity cannot be ef-
fectively separated. In a similar fashion to that adopted in the elasto-plastic formulation,
the total strain is separated into an elastic strain, an 'initial' strain, and a viscoplastic
strain. The latter is characterized by an strain rate which is nonzero only for stresses
above a certain yield value, and is defined in terms of a plastic potential and a fluidity
parameter [178].
Literature Review 54

Table 2.1: Classification of techniques for measuring residual stresses (modified from
[130]).

Stress-relaxation techniques using Sachs boring-out technique.


electric and mechanical strain gauges
Centre-hole (blind-hole) drilling technique.

Rosenthal-Norton sectioning technique.

Stress-relaxation techniques using Brittle coating-drilling technique.


apparatus other than electric and
mechanical gauges Photoelastic coating-drilling technique.

Grid system-dividing technique.

Diffraction techniques X-ray diffraction.

Neutron diffraction.

Techniques using stress-sensitive Ultrasonic techniques.


properties
Hardness techniques.

Magnetic techniques.

Cracking techniques Hydrogen-induced cracking techniques.

Stress corrosion cracking techniques.


Literature Review

Trial-and-Error

Processing
Properties
Conditions

Quench Bath Quality

Processing
Structure Properties
Conditions

Process Modelling

Processing
Structure Properties
Conditions

Transport Materials
Phenomena Science

Figure 2.1: Methods adopted to predict the performance of quenching processes.


Literature Review 56

Figure 2.2: Schematic representation of a typical cooling curve showing the different
stages of boiling.
Literature Review

vaoor
Blanket

Natural Nucleate Transition Film


Convection Boiling Boiling Boiling
Regime Regime Regime Regime

L
Isolated J Jets and
Bubbles Columns

r Critical Heat Flux (CHF)

x
3

a
X
a
* Leidenfrost Point

L
Onset of Nucleate Boiling (ONB

Wall Superheat log A T s a t

Figure 2.3: Schematic representation of a typical boiling curve for saturated pool boi
[75].
Literature Review 58

S u r f a c e temperature (8,)

LPs.
/
/
/
/
/
/

\ bp /
\1 /
/
/
>
/

(c) V _ /
/

S u r f a c e temperature (9S)

Film boiling
(2) Transition boiling
(3) Nucleate boiling
@ Convection (non-boiling)

Figure 2.4: Correspondence between (a) cooling, (b) boiling, and (c) evaporation curves
[56].
Literature Review 59

log (Tw-Tszt)

Figure 2.5: Regimes in boiling heat transfer [87].


Literature Review 60

Time

Figure 2.6: Schematic representation of fraction transformed as a function of time, for a


typical nucleation and growth phase transformation.

Figure 2.7: Schematic representation of fraction transformed as a function of tempera-


ture, for a typical martensitic phase transformation.
Literature Review 61

Figure 2.8: Schematic diagram illustrating the application of the additivity principle.
The non-isothermal reaction follows the path marked by the arrows. The isothermal
transformation kinetic curves at two temperatures are labeled T i and T .
2
Chapter 3

Scope and Objectives

3.1 Scope of the Research Programme

The goal of the present study was to apply the microstructural engineering approach to
generate a basis for understanding the microstructural and mechanical response during
forced convective quenching (heat treatment) of steel rods. It is part of an ongoing generic
project aimed at the prediction of microstructure and mechanical properties obtained
from thermal and thermo-mechanical processing operations.
To achieve this objective, mathematical models of thermal/microstructural evolution
and stress generation were developed. Once debugged and tested, the mathematical
models were validated by comparing calculated results with values measured in the lab-
oratory. The effect of internal stresses on the transformation kinetics is relatively small;
however, the computing costs increase dramatically when this interaction is considered.
Also, there is uncertainty as to how to best apply the results of simple mechanical tests
to a complex stress state. Thus, in this work, the effect of internal stresses on transfor-
mation kinetics was not considered. As a consequence of this assumption, the models
could be effectively uncoupled, i.e., the results of the thermal/microstructural model were
adopted as input to the stress generation model.

The problem at hand is highly nonlinear and, therefore, a numerical solution needs
to be implemented. In structural analysis, the finite-element method ( F E M ) has be-
come a de facto standard (mainly due to its ability to handle components of complex

62
Scope and Objectives 63

geometry) and was adopted for the analysis. Based on geometric considerations, an ax-
isymmetric formulation was deemed adequate. A finite-element program developed to
simulate thermal stresses in fused-cast monofrax-s refractories [182] was used as a basis
for the transient thermal/microstructural model while a computer program for the time-
independent, elasto-plastic analysis of stress evolution in water spray-quenching of steel
bars [47] was adopted to model internal stress generation.
Due to difficulties in instrumenting samples under industrial conditions, a laboratory
quenching facility was designed and built. Measurements made with this apparatus
were used to characterize heat transfer in forced convective quenching as well as provide
samples for residual stress measurements (by neutron diffraction) and microstructural
characterization to validate the mathematical models.
A critical component of the models is the heat transfer boundary condition. Due
to a lack of data in the literature for the conditions of interest, the surface heat flux,
as a function of surface temperature, was estimated from the measured temperature
response of instrumented samples through the solution of the inverse heat conduction
problem (IHCP). A n existing computer program that implements the sequential function
specification technique was modified for this purpose [183].
A characteristic of mathematical models of heat treatments is the large database
that is required. For this study, the thermophysical and thermomechanical properties
were taken from the literature while the transformation kinetics under isothermal and
continuous cooling regimes were measured in the laboratory using a Gleeble 1500 ther-
momechanical simulator.
Scope and Objectives 64

3.2 Objectives of the Research Programme

The objective of the present study was :

To formulate, develop and verify mathematical models capable of


solving the transient thermal, microstructural and stressfieldsob-
tained during forced convective quenching of steel bars.

To accomplish this goal, the following subtasks were undertaken :

To characterize the surface heat flux, as a function of surface temperature, in forced


convective quenching.

To formulate, develop and verify a computer program to solve the inverse heat
conduction problem.

To measure transformation kinetics under isothermal and continuous cooling con-


ditions.

To characterize the final microstructure and hardness in the quenched samples, for
comparison to model predictions.

To measure the residual stress distribution in the quenched samples via neutron
diffraction, for comparison to model predictions.
Chapter 4

Heat Transfer Model

To model the evolution of the thermal and microstructural fields, a continuum mechanics
approach was adopted. As discussed earlier, the thermal and microstructural fields need
to be solved simultaneously, while ignoring both the effect of elasto-plastic deformation
on phase transformation kinetics and the heat generated by the deformation.

In this chapter, the equation governing heat transfer during quenching of steel, to-
gether with the relationships between the kinetics of phase transformation and heat
evolved, are presented. The principle of additivity was invoked to compute the advance
of the transformations taking place. Then, a solution to this problem, based on the
finite-element method, is presented.

4.1 Governing Equation

The governing equation for the heat transfer phenomena in an isotropic body, including
heat generated due to phase evolution, is as follows

V-fc VT + q = pC p (4.1)

where, in general, T = T(x{,t) and the rate of heat generation, q = q(xi,t). For an

axisymetric problem, this equation reduces to :

(4.2)

65
Heat Transfer Model 66

Ignoring the end effects, the above equation reduces to :

For an infinitely long, solid cylinder cooling in water, one can express the boundary
conditions (B.C.'s) for the heat transfer problem by the following

dT
B . C . 1: ^ = 0 at r = 0, f > 0 (4.4)
or

and

dT
B.C. 2 : -k =-h(T-Tf) at r = R, t > 0 (4.5)
or

i.e., symmetry at the centreline and heat transfer by convection at the surface, respec-
tively.
The bar is assumed to be at a uniform temperature before the start of quenching.
Then, the initial condition (I.C) is

I.C.: T(r,t)= T0 0 < r < R, t = 0 (4.6)

4.1.1 Rate of Heat Evolved

The rate of heat evolved during a transformation constitutes the link between the thermal
and microstructural fields. At each time increment, At, it is directly proportional to the
rate of transformation :

9= ^ ^ ^ (4-7)

In order to compute q, the new fraction transformed (which is a function of temper-


ature and cooling path) must be known. It follows that an iterative scheme must be
Heat Transfer Model 67

implemented.

For diffusional transformations, the fraction transformed during continuous cooling

can be computed using kinetic information obtained under isothermal conditions by in-

voking the additivity principle as follows : 1

The 'virtual time', the time required to reach the old fraction transformed, X, m
at

the current iterated temperature, m + 1


T*, is obtained by solving the J M A K equation for

time, i.e., obtaining the inverse of the J M A K equation :

In' 1

x
l+1
6* = ( r o + 1
X(T,*)) (4.8)

where n and b are kinetic parameters in the J M A K equation. The new fraction trans-
formed is then calculated by adding At to m+1
0* and substituting in the Eq. (2.4) :

m+l
X* = 1 - exp [ - & * ( m+1
0 * + At) "] n
(4.9)

The kinetic parameters n, b and A c c t are computed via empirical equations that are

developed based on isothermal (n and b) and continuous cooling (transformation start,

^AVCCT) tests. It should be noted that, in the model, tAv CCT is calculated at each node

ignoring recalescence, i.e., it is assumed that each node cools independently of the rest . 2

The kinetics of the martensitic transformation are described by the Koistinen and

Marburger [99] equation (Eq. (2.6)). In this case, the fraction transformed is only de-

pendent on the current temperature.

Given that the kinetic relationships are usually expressed in terms of undercooling

below the thermodynamic transformation temperature, the relevant sections of the phase

"Til the following equations


m
() and () denote values at the end and beginning of the current time
+ 1 m

interval, respectively. An asterisk, ()*, denotes a quantity evaluated at the guessed temperature T*. m+1

2
Based on the fact that the experimental values of tAvccr obtained from a set of independent
a r e

experiments conducted at several cooling rates.


Heat Transfer Model 68

diagram need to be calculated. To this end, a code based on the algorithm proposed by

Kirkaldy et al. [104] was developed. Once the Ae3 and A c m lines were computed, the

eutectoid temperature and composition were calculated as their intersection; instead

of solving the corresponding equations simultaneously a trial and error procedure was

implemented, with a convergence criterion given by

e= T A e S
~ T A c m
< 0.05 (4.10)
-L Acm

The code calculates the A c m and Ae3 lines as 2nd and 3rd order polynomials as a
function of % C , respectively. For hypo- and hyper-eutectoid steels, the % C vs T functions
for temperatures below the eutectoid can be obtained by inverting the polynomials,
which was accomplished by using reversion formulae [184] based on the Taylor series
expansion. It was found that 2nd and 3rd order reversion formulae are adequate to
invert the 7Fe C+7 and 7 a + 7 lines, respectively.
3

4.2 Finite-Element Formulation

The heat evolved during the transformation, the boundary conditions and the variation of

thermophysical properties with temperature, make this a highly non-linear problem and,

therefore, an analytical solution is not an option. Instead, one has to resort to a numerical

solution. In particular, the finite-element method was chosen to solve Eq. (4.3). The basic

components of the finite-element formulation are : 1) the finite-element equations and 2)

the solution algorithm for the nonlinear equations.


Heat Transfer Model 69

4.2.1 Finite-Element Equations


r
In the finite-element method the thermal field at the element level is approximated by
the following relationship :
{T} {f} = [N]{aY (4.11)

where {T} = [T] represents the thermal field, [N] is the shape function matrix , and
T 3

{a} is the nodal temperature vector. There is one degree of freedom at each nodal point.
e

The gradient of the thermal field is expressed as

V T = [L]{T] = [L][N]{aY (4.12)

where [L] is the following linear operator (for an axisymmetric problem) :

^-{!} w
The [B] matrix is defined by
[B] = [L][N] (4.14)

then

VT = [B]{a} e
(4.15)

and the heat flux by conduction is computed from the thermal gradient as follows :

{q}T
= - f c V T = -k[B]{a}e
(4.16)

B y adopting the Galerkin method, a weak form of Eq. (4.3) can be derived, from

3
Linear-order (4-node) and quadratic order (8-node) shape functions applicable to two-dimensional
isoparametric elements can be found in Ref. [185].
Heat Transfer Model 70

which the following equivalent expression of the governing equation is obtained 4

[C] {aY + [K] {aY


e e
= {RY (4.17)

where, for the axisymmetric case, the capacitance matrix [C] and the conduction matrix e

[K] at the element level are given by


e 5

[C] e
= 2TT / [N} p C r[N] dr dz
T
p

2wr f [N] p C [N] dr


T
P (4.18)

and

[K] e
= [K ] rr
e
+ [K ] zz
e
+ [K ] cv
e
(4.19)

[K ]rr
e
= 2n f -^-[N] kr-^-[N]drdz T
(4.20)
J A* or or
w 2nf f [N] kr[N]drdz
T

JA E
or or
[K ] zz
e
= 2*1 ^-[N] kr^-[N]drdz
T
(4.21)
JA" OZ OZ
2nf ( \NYk-\N\drdz
JA' oz dz
[K ]cv
e
= 2TT / [N]hrdrdz (4.22)
Jc e

w 27rf / [N]hdrdz
Jc e

respectively.
Considering boundary conditions of convection and/or specified surface heat flux, and

Details of the derivation are given in Appendix A.


4

5
T o facilitate the integration, the quantities that depend on the coordinates are evaluated at the
centroidal point (r = J2i ^t*t) d
r 7 a n
denoted by Q.
a r e
Heat Transfer Model 71

a spatially varying volumetric heat source, the elemental load vector has three compo-
nents :

{/} e
= iUY + {hvY + {f Y q (4.23)

where

{/} e
= 2TT / [N]q r drdz

Pa 2?rf / [N}qdrdz (4.24)


JA e

{/,}" = 2;r / q r dC
s

2?rr / ? s dC7 (4.25)

if*,}* = 2TT / /*7> dC7


2?rf / hTt dC (4.26)

The system of equations given by Eq. (4.18) represents the equivalent finite-element
form of the governing differential equation and must be assembled to give a system of
equations at the global level before solving. These equations are then solved by adopting
the following three-point recurrence scheme [186], as described in [182] :

. 4 + A ^ - J ( -^J
4

with a Crank-Nicholson approximation adopted to start the algorithm.

4.2.2 Solution A l g o r i t h m

Due to kinetic factors, phase transformations in actual processes do not start at the tem-
peratures indicated in a phase diagram. Also, the rate of heat released by the austenite
decomposition reactions may be larger than the rate of heat removed by the cooling
medium, producing recalescence. These two factors combine to produce cooling curves
Heat Transfer Model 72

like the one shown, schematically, in Figure 4.1.


For each of the sections shown in the figure the heat released was evaluated in a
different way :

For conditions under which the transformation either has not yet begun or has
already been completed, a value of zero was assigned to the heat generation term.

When the transformation is taking place, a definite amount of heat (related to the
current fraction transformed) was released. A n iterative scheme was necessary in
order to evaluate this contribution. A first approximation for the thermal field was
obtained assuming that no heat was released. With this information, the additivity
principle was then used to evaluate the fraction transformed and the total heat
released, q, was then calculated from Eq. (4.7). The thermal field was re-calculated
using this value of q, and the process was repeated until the following convergence
criterion was met :

(4.28)
v / (
E m + i )2
r

where JYL{ T*mJrl


m + 1
T) is the Euclidean Norm [187] and e is a predetermined
2

convergence criterion.

Depending on the cooling rate, the parent phase may transform to a mixture of equi-
librium (pearlite, ferrite) and non-equilibrium (bainite, martensite) phases and, there-
fore, several possible scenarios needed to be incorporated in the computer code. This
was achieved through decision trees, as illustrated in Table 4.1. The progress of the
transformations was monitored at all times, and an increment of fraction transformed
was computed at every time step. The start of the austenite-to-ferrite, -pearlite, or -
bainite transformation was determined by either calculating the time spent below the
corresponding transformation temperature and comparing it to the measured incuba-
tion time, or by using empirical correlations of start time as a function of cooling rate.
Heat Transfer Model 73

The austenite-to-martensite transformation was assumed to start as soon as the local


temperature dropped below M . g

A flow diagram of the numerical procedure is shown in Figure 4.2. As an example,


details of the sequence for evaluating the heat evolved during the austenite-to-pearlite
transformation are shown in Figure 4.3.
In order to speed up the solution, a relaxation technique was implemented. The
temperature field was under-relaxed after every iteration following :

T = toT* + (1 - u)T (4.29)

where T* is the current iterated thermal field and T is the thermal field that will be
used for the next iteration. A value of 0.5 was used for the under-relaxation parameter,
u>. Realizing that the very first iteration under-estimates the actual thermal field, the
relaxation factor was set to u < 0.5 for that particular iteration. Typically, less than 10
iterations were needed to achieve convergence.

4.3 Verification of Mathematical Model of Heat Flow

The finite-element code developed by Cockroft [182] to simulate heat transfer in the
Epic-3 Monofrax-S casting process was modified to model microstructural and thermal
evolution during heat treatment. The verification of the computer code was carried
out in three stages. First, numerical results for heat conduction, without considering
heat generation, were compared against analytical solutions. Secondly, a semi-analytical
solution was carried out for the cooling of a cylinder with a uniformly distributed heat
source of known strength in order to verify the solution algorithm (without including the
kinetics of the phase transformation). Lastly, a finite-difference simulation of thermal
and microstructural responses for controlled air cooling of eutectoid steel rod [188] was
used to test the complete finite-element code. In all uniform initial temperature
Heat Transfer Model 74

distribution was assumed.

4.3.1 Infinite Solid C y l i n d e r : q 0

The governing equation for the case of no generation nor consumption of energy is

1 & ( n dT\ dT . .

where T = T(r,f). For constant thermophysical properties, Eq. (4.30) reduces to

dT
2
IdT IdT .
Or 2
r or a ot

4.3.1.1 Newtonian C o o l i n g .

A n extreme case of Eq. (4.31) is that of Newtonian cooling (Bi < 0.1), where it is assumed
that no thermal gradients exist inside the solid cylinder, i.e., the thermal resistance due
to convection is much higher than that due to conduction. After performing a heat
balance, the ordinary differential equation that describes the problem is [189]

- Vp C ^ p - hA(T - T )f (4.32)

with T = T(i) only.


After integration, the thermal response, for Tf = 0, in the cylinder is given by,

T
= ^(~MV)
T (433)

The input to the analytical and F E solutions is given in Table 4.2. Figure 4.4 compares
the F E M and the analytical solution for a 1 mm-dia. rod cooled under Newtonian
conditions. Since the numerical solution reflects the fact that a thermal gradient does
Heat Transfer Model 75

exist, the temperature plotted corresponds to values at r/R = 0.5. Excellent agreement
can be observed.

4.3.1.2 Non-Newtonian Cooling.

The more general case for non-Newtonian cooling of an infinite solid cylinder by con-
vection to a medium of constant temperature with a constant heat-transfer coefficient,
was also considered. In this case, Eq. (4.31) applies, subject to the following boundary
conditions

B . C . 1: T= 0 at r = 0, i > 0 (4.34)

B.C. 2 : -k^j- = -h(T f -T) at r = R, t > 0 (4.35)

where T = T(r, t).

The analytical solution to Eqs. (4.31), (4.34) and (4.35) (for 7> = 0) is, using the
method of separation of variables [190],

r ( r , t) = 2
-^t i+ffr],
n exp(-aA *)
m m
2
(4.36)
K
m=l( + A
H )J (\ R)
2
0 m

where
H
= \ ( 4
' 3 7 )

and the eigenvalues, A , are the positive roots of either


m

\ J' {\ R)
m 0 m + HJ (X R)
0 m =0 (4.38)

or

A Ji{X R) m m - HJ {\ R)
0 m =0 (4.39)

with J {x), J' (x) and J\{x) being the Bessel function of the first kind of order zero, its
0 0
Heat Transfer Model 76

first derivative, and the Bessel function of the first kind of order one, respectively.
Eqs. (4.36) and (4.39) can be rewritten in non-dimensional terms (Fo = at/R ,Bi 2
=
HR) as :

T
=4(WWXM t , j (
- ( U ) ! F o )
'
(4 40)

and
f(X R)
m = X RJi(X R)
m m - B\J {X R)
0 m =0 (4.41)

The data used as input are given in Table 4.3. A comparison between the numerical
and the analytical solution for cooling at the centreline of a rod in which non-Newtonian
conditions apply (Bi = 0.5) is shown in Figure 4.5. Again, very good agreement can be
seen.

4.3.2 Infinite S o l i d C y l i n d e r : q = f(T)

A semi-analytical solution for the Newtonian cooling of an infinitely long cylinder with
a uniformly distributed heat source, q = Q (T), was carried out (see Appendix B ) . As
0

is the case with virtually all analytical solutions, the thermophysical properties were
assumed to be independent of temperature.

The specific function, Q (T), used to test the numerical scheme is shown in Figure 4.6;
0

it was chosen because it resembles the shape of the heat source history in a typical quench.
Figure 4.7 shows the thermal response at the centreline for a 10 mm-dia. rod for values
of Q ,m
0 of 2.75 x l O and 6.0 x l O W m ~ s
7 8 3 - 1
(typical values for heat evolved during the
austenite-to-pearlite transformation range from 5 x 10 to 3 x 10 W m ~ s ) . The curve
7 8 3 _1

for Qo,m = 0, i.e., no heat evolved, is included for comparison. Note the recalescence
produced by the heat generation term.

As described in Appendix B , the semi-analytical solution was achieved in a step-wise


fashion. Results for a large value of Q ,m 0 obtained using different time intervals, At,
showed instabilities when a large value of At was chosen (Figure 4.8). This observation
Heat Transfer Model 77

points out the importance of a proper selection of the time interval, A t .


The numerical and semi-analytical solutions were compared for values of Q 0ym =
2.75 x 10 and 6.0 x l O W m ~ s
7 8 3 _1
(see Figures 4.9 and 4.10, respectively). Excellent
agreement was found for both cases . 6

4.3.3 Infinite Solid Cylinder : Air Cooling of Eutectoid Steel

The predicted thermal response of an infinitely long, 10 mm-dia. rod of eutectoid carbon
steel obtained with a finite difference scheme [188], assuming constant thermophysical
properties, was also used to verify the mathematical model. The properties adopted are
given in Table 4.4. Figure 4.11 shows the results of both the finite difference and the
finite element analyses for the 10 mm-dia. rod; very good agreement is evident between
comparing the predicted temperatures at both the centre and the surface of the rod.
Figure 4.12 shows the corresponding results for microstructural evolution.

4.4 Summary

A computer program based on the finite-element method has been developed to simulate
heat transfer and microstructural evolution during forced convective quenching of steel
bars. Given that the solutions of the thermal and microstructural fields are coupled,
through the heat of transformation, an iterative procedure was implemented. To accel-
erate the convergence, a relaxation technique was applied. The principle of additivity
was invoked to compute the progress of the transformation under continuous cooling
conditions. The J M A K equation was used to describe the kinetics of diffusional trans-
formations, while the Koistinen and Marburger equation was adopted to compute the
martensitic transformation. The code has been verified by comparing results of numerical
and analytical solutions.

6
No relaxation was applied for these calculations
Heat Transfer Model 78

Table 4.1: Decision tree to determine the sequence of calculations during phase transfor-
mations in a eutectoid steel.

Product Condition Action


/.'.OLD > 0 A/* = 0 p

/B,OLD >
0 A/p = 0
P /P.OLD > 0-995 A/p = 0
/p,OLD =
0 Check start of transformation
0 < /P.OLD < -995 Calculate A / p
/',OLD > 0 A/* = 0
B

/P.OLD > 0-995 A/* = 0 B

B /B.OLD > 0-995 A/* = 0 B

/B,OLD =
0 Check start of transformation
0 < /B.OLD < 0-995 Calculate A / B

/P,OLD + /B,OLD 0-995


>
A/* , = 0
A

a' /a',OLD > 0-995 A/;< = o


/<*',OLD =
0 Check start of transformation
0 < /.'.OLD < 0-995 Calculate A / * ,

Note : fi = normalized fraction transformed

Note : / * = true fraction transformed.


Heat Transfer Model 79

Table 4.2: Input data used for the comparison of finite-element and analytical solutions
under Newtonian cooling conditions. The corresponding Biot number was B i = 0.01.

Variable Value
R 0.001 m
k 20 W / ( m K )
P 10000 k g / m 3

Cp 200 J/(kg K )
h 200 W / ( m K )
2

T 0 1000 C
Tf 0 C
At 1.0 s

Table 4.3: Input data used for the comparison of finite-element and analytical solutions
under non-Newtonian cooling conditions. The corresponding Biot number was B i = 0.5.

Variable Value
R 0.050 m
k 20 W / ( m K )
P 10000 k g / m 3

Cp 200 J/(kg K )
h 200 W / ( m K )
2

To 1000 c
Tf 0 c
At 250.0 s
Heat Transfer Model 80

Table 4.4: Input data adopted for the comparison of finite-element and finite-difference
simulations of air cooling of eutectoid steel rod [188].

Variable Value
k 25 W / ( m K )
P 7650 k g / m
3

625 J/(kg K )
h 160 W / ( m K )
2
(simulates a cooling rate of 9.4 C s )
_1

T0 850 C
Tj 20 C
A# _
7 P 88200 J k g " 1

In np 2.2
In bp -41.49(7^ - r ) - e x p [ - 0 . 0 3 6 4 3 ( ( r - T)]
O 7 2 3
A l

In tAvccr 29.73 + 0.05816(r - T) - 8.622 log(T - T)


Al Al

+ 6.27 (XC + XMJG.O)

Note : The thermophysical properties : k,p and C , of both austenite and pearlite were
p

assumed to be equal.
Heat Transfer Model
Heat Transfer Model

( START )

Read input data


Initialize variables

{W <- { T } m

g = 0

Calculate {T*}

{T }^Lo{T:}
prop + (l-u){T }prop

Calculate {T*} QTRNSF-MICRO

Convergence ? y

* <- t + A t

t>t l F ( END

n
{T m + 1
} <- {T*}

{r} <- {/ > m+1

Figure 4.2: Flowchart for the solution of the thermal-microstructural probL


Heat Transfer Model 83

( QTRNSFJVIICRO )

T > ? RETURN
n
f m
> 0.995 ? ^ RETURN

n
t* = t t Al

t* > tAVCCT ?
>( R E T U R N j
n
Compute f m+1
(nodal)

A / - / m + 1
- / m

Compute <j (nodal)

RETURN

Figure 4.3: Details of the model sequence for evaluating the heat released during austen-
ite-to-pearlite decomposition.
Heat Transfer Model 84

-100 1
1 1 1 1
1

0 10 20 30 40

Time, s

Figure 4.4: Comparison between the finite-element and the analytical solution for the
newtonian cooling (Bi= 0.01) of an infinitely long cylinder.

1100

Analytical
O FEM

Centerline
4>
H

- e e e e o
_i i i
500 1000 1500 2000 2500 3000

Time, s

Figure 4.5: Comparison between the finite-element and the analytical solution at the
centreline for the Non-newtonian cooling (Bi= 0.5) of an infinitely long cylinder
Heat Transfer Model 85

100 200 300 400 500 600 700 800 900


o
Temperature, C

Figure 4.6: Function Q used in the semi-analytical solution of Newtonian cooling of a


0

long rod with a uniformly distributed heat source, where Qo, is the maximum value of m

the function QQ.

-3 -1

900 - Q = 0.0 MWm s


o. m -3 -1
Q = 27.5 MWm s
^o, m -3 -1
\ Q. = 60.0 MWm s
800 - ^o, m

o
. 700 -
u

* 600 -
<D
o.
S
<>
i 500 -
E-

400 -

300 h

200 I 1 1 1 1 1 1

0 20 40 60 80 100 120

Time, s

Figure 4.7: Semi-analytical solution, at the centreline of a 10 mm-dia rod, for Q , 27.5 0 m

and 60.0 MWm" r . The curve for Q = 0 M W m s is included for comparison.


3 - 1 - 3 _1
0
Heat Transfer Model 86

850

At = 1 0 ms
At = 1 0 s
825

o
o
800
in
3

Cll

P.
775
= 600 MWm s 3 1

s
a>
750

725
20 40 60 80 100 120

Time, s

Figure 4.8: Effect of At on the semi-analytical solution, at the centreline of a 10 mm-dia.


rod, for Q , = 600.0 M W m " s" .
0 m
3 1

900

Analytical
FEM (centre)
FEM (surface)

400
20 40 60 80

Time, s

Figure 4.9: Comparison between the semi-analytical and the numerical ( F E M ) solution,
during cooling of a 10 mm-dia. rod, for Q ,m = 27.5 M W m s .
0
- 3 - 1
Heat Transfer Model 87

Figure 4.10: Comparison between the semi-analytical and the numerical ( F E M ) solution,
during cooling of a 10 mm-dia. rod, for Q , = 60.0 M W m
0 m s . - 3 _1

500
10 20 30 40 50

Time, s

Figure 4.11: Finite-difference [188] and finite-element solutions for the thermal response
at the centreline and surface of a 10 mm-dia. eutectoid carbon steel. The cooling rate at
728 C was 9.4 C s . _1
Heat Transfer Model 88

0 10 20 30 40 50
Time, s

Figure 4.12: Finite-difference [188] and finite-element solutions for fraction transformed
to pear lite at the centreline and surface of a 10 mm-dia. eutectoid carbon steel. The
cooling rate at 728 C is 9.4 C s .
-1
Chapter 5

The Inverse Heat Conduction Problem

The accurate determination of the heat transfer boundary condition is a crucial com-
ponent of mathematical, models aimed at predicting the evolution of the thermal, mi-
crostructural and stress fields during the heat treatment of a metal. This task is difficult
to accomplish experimentally due to the high temperatures involved, the uncertainty
associated with surface temperatures measured with thermocouples, and, in the case of
quenching, the presence of boiling heat transfer. A viable alternative consists of estimat-
ing the heat flux and surface temperature from a temperature response measured inside
the body; this is known as the inverse heat conduction problem (IHCP).
In most heat treatment applications of inverse analysis the determination of surface
heat flux constitutes a function estimation problem, i.e., the unknown function is es-
timated without any prior knowledge of the actual functional form of the variation of
surface heat flux with time. To estimate this function, a large number of surface heat
flux components need to be determined. From a mathematical point of view, the I H C P
can be classified as an ill-posed problem, i.e., it does not satisfy the requirements of
existence, uniqueness and stability of the solution. This translates into inverse solutions
being highly sensitive to small fluctuations in the measured temperatures. Current so-
lution algorithms for solving the IHCP have been designed to overcome this problem.
The method of solution can be either analytical or numerical. It may be applied either
sequentially, i.e., by estimating a single component of the surface heat flux function at
each time step, or to the whole time domain, in which case all components are estimated

89
The Inverse Heat Conduction Problem 90

simultaneously. Due to the restrictive nature of analytical methods (chiefly their inability
to handle nonlinearities), practically all IHCPs need to be solved numerically.

In this chapter, the application of the sequential function specification algorithm to


the solution of the IHCP is presented, with special emphasis on the modification (to
accommodate cylindrical geometry) of an existing computer code.

5.1 Solution Algorithm

The one-dimensional IHCP is schematically illustrated in Figure 5.1. A plate of thickness


2L is initially at a uniform temperature, T . For times t > 0, an unknown, time-dependent
0

heat flux, q(t), is applied at the boundary x = L, while the boundary at x = 0 is that of
symmetry with respect to the thermal gradient. In order to estimate q(t), temperature
measurements are made at a position inside the body, denoted by x\. It is assumed
that no information is available regarding the specific form of q(t) and, therefore, a large
number of components of q(t) need to be estimated.
In the past, 'brute force' algorithms based on a trial-and-error procedure have been
used to solve the IHCP. At each time step, a guessed value of the heat flux is adopted for
the calculation of the temperature at the sensor position, and the iterations continue until
the difference between the calculated and measured values falls within a predetermined
tolerance. Current methods adopted to solve this function estimation class of IHCPs fall
into one of two broad categories [191] : 1) function specification and 2) regularization.
The algorithm adopted in this investigation was proposed by Beck et al. [192] and can
be described as a sequential function specification algorithm. In this algorithm, the
surface heat flux is estimated by assuming temporarily that heat flux components up to
(r 1) time steps ahead of the current time are constant (see Figure 5.2). For a single
thermocouple, the estimated temperature exactly matches the measured value at each
time step, if r is set to 1. However, the solution becomes unstable and very sensitive to
The Inverse Heat Conduction Problem 91

measurement errors (specially when small time steps are adopted).


The unknown heat flux at the surface is obtained from temperature measurements
spanning several future time steps, by minimizing the following least squares expression
with respect to the heat-flux component at time t = IM '

S
=E E (Yf *-4 1
- T^ - )1 1 2
(5.1)
=i i=i

where y^ +'
M -1
J the measured temperature at the jth sensor at time tM+i-i,
s T^ +l_1

is the corresponding calculated temperature, and r is the number of future time steps
adopted for estimating q . M
It is assumed that the heat flux is known for t < M - I -
Through the use of the future data concept, stable solutions, even for small time steps,
can be obtained. It should be noted that by minimizing the least square norm, instead
of making it zero, the uniqueness of the solution is guaranteed [193].
Differentiating Eq.(5.1) with respect to q , replacing q M M
by q M
(the estimated heat
flux at time M), and setting the expression equal to zero, one obtains :

2 E ( C + i
- 1
- 7 f +
- ) ( a^if) . = (5-2)
8=1 j=i \ " / '

Assuming that the future temperature at the sensor position j can be calculated from
a Taylor series expansion about q ~ : M l

,, . , *M+i-l * M+i-1
Tf^~ = T l
+(q -q ~ )X 3
M M 1
3 (5.3)

where the asterisk implies that the T and X functions are evaluated using the thermal
properties and surface heat flux values at time t\i-i- The quantity X^ +t_1
is called the
sensitivity coefficient and is defined by

QrpM + i-l
x r - 1
= - f a - <->
5 4
The Inverse Heat Conduction Problem 92

Note that, by using this definition, the partial differential equation that describes the
sensitivity coefficient distribution is obtained as

d dx dx
k
dxi dxi = pC dt p (5.5)

with the following boundary conditions :

8X
B.C. 1 : = 0 at X l = 0, t > 0 (5.6)
Ox i
dX
B.C. 2 : -k-z = 1 at X l =X u t >0 (5.7)
ox i
representing an insulated (or, equivalently, symmetric) boundary and a specified heat
flux condition for the inactive and active boundaries, respectively. It is evident that
Eq. (5.5) has exactly the same form as the differential equation for the thermal field,
the only difference between the two problems, from a mathematical point of view, being
the exact form of the boundary conditions. It follows that the same algorithm adopted
to solve the direct heat conduction problem can be applied to compute the sensitivity
coefficient distribution, resulting in a more efficient code.
Introducing Eq. (5.4) into Eq. (5.3) and solving for q M
gives

1 r J
/ *M+i-l\ * M+i-l

dM
=^ +T ^ E (Y?*- 1
~T 3 ) X 3 (5.8)
M i = l j = 1 V /

where
JL-L/* M+i-l\ 2

AM =E E ^ (5-9)
;=ij=i v J

Note that estimating T . . . T


M M + r
- 1
and X M
...X ~ M+r 1
by adopting the thermo-
physical properties, k and pC , which correspond to the previous time step, the problem
p

has been linearized and Eq. (5.8) becomes explicit in q . M


The justification for this as-
sumption is that, for a small time step, At, the thermal properties change little at a given
The Inverse Heat Conduction Problem 93

location from one time to the next even though there may be a large variation in prop-
erties from one end of the body to the other; an iterative procedure is then not needed,
even for a problem with varying thermal properties. The algorithm is summarized in the
flowchart shown in Figure 5.3.
Once the response of the heat flux and temperature at the surface have been calcu-
lated, they are used to estimate the heat-transfer coefficient. It should be noted that the
algorithm calculates both q and T , but they are calculated at slightly different times : q
s
M

is best associated with ijvf-1/25 while T corresponds to tjvf- Therefore, the heat-transfer
s

coefficient at time M is estimated from :

(5.10)

The computer program C O N T A [183] incorporates the sequential function specifi-


cation algorithm described above, for the case of a one-dimensional, planar IHCP. This
code uses the Crank-Nicholson implementation of the finite-difference method to solve the
partial differential equations describing the thermal field and the sensitivity coefficient
distribution. It has been successfully applied to estimate the heat-transfer coefficient
during pool quenching of stainless steel disks [194] and to characterize interfacial heat
transfer during metal solidification [195,196].

5.2 Modification for C y l i n d r i c a l Coordinates

The code C O N T A was modified [197] to incorporate cylindrical geometry - appropriate


for the quenching of bars. Once coded and debugged, the program was tested by compar-
ing results obtained using analytical solutions for two cases : 1) a solid cylinder subjected
to a constant surface heat flux and 2) a solid cylinder subjected to a medium of constant
heat-transfer coefficient and fluid temperature. In both cases the direct problem was
The Inverse Heat Conduction Problem 94

first solved analytically to obtain the thermal response at a given point in the domain,
which was then adopted as input to the numerical solution of the IHCP to estimate both
surface heat flux and surface temperature.

Case 1 : Solid Cylinder Subjected to a Constant Heat Flux at the Surface.

This problem is described mathematically, assuming constant thermophysical properties,

by Eq. (4.31), subject to the following boundary and initial conditions :

dT
B.C. 1 : = 0 at r = 0, t > 0 (5.11)
dr

, dT
B.C. 2 : at r = R, t > 0 (5.12)

LC. : T(r,t)= T0 0 < r < R, t = 0 (5.13)

for which an analytical solution, for T = 0, is available [198] :


0

. 2q at q Rf r2
1 ~ / a\ H\ J {r\ /R)
(5.14)
r r 0 0 m 0 m

kR k \2R 2
4
m=l

where A are the roots of the transcendental equation


m

Wn) = 0 (5.15)

The results at the centreline (T(0,t)) were used as input for the program and the
surface heat flux was recalculated. The input data are given in Table 5.1.

Figure 5.4 shows a plot of estimated surface heat flux versus time as well as the input
heat flux; good agreement can be observed except at early times. The thermal response
at the surface was also computed and compared to the analytical solution; very good
agreement was obtained, as shown in Figure 5.5.
The Inverse Heat Conduction Problem 95

Case 2 : Solid Cylinder Subjected to a Medium of Constant Heat-Transfer

Coefficient.

Eq. (4.31) is again applicable in this case, with the following boundary and initial con-
ditions :

dT
B.C. 1 : = 0 at r = 0, t > 0 (5.16)
dr

dT
B.C. 2 : at r = R, t > 0 (5.17)

I.C. : T(r,0) = T 0 0 < r < i?, t = 0 (5.18)

The analytical solution is given in [190] as

2HT 00
Jo{X r)
^E( A
exp(-aA i)
m
T(r,t) (5.19)
0 2
m
R 2
+ W)J (\ R)
0 m

where H h/k and A are the roots of the transcendental equation


m

XmJ' {XmR) + HJ (\ R)
0 0 m =0 (5.20)

The program was run, using the thermal response at the centreline calculated with
the analytical solution as input, and the heat-transfer coefficient was back-calculated.
The data used is shown in Table 5.2. Figure 5.6 shows good agreement between the
estimated and the input heat-transfer coefficients, except at early times. The estimated
thermal response at the surface also agrees well with the analytical solution, as shown in
Figure 5.7.
The Inverse Heat Conduction Problem 96

5.3 Application to Controlled Air Cooling of Rods

A n additional test, using experimental information, was conducted to verify the modified
code [197]. Measured temperatures at the centreline of rods subjected to forced air
cooling were used as input to estimate the corresponding heat-transfer coefficients. The
experiments, reported by Campbell et al. [42], involved a range of steel grades, rod
diameters and air velocities, and were designed to simulate the Stelmor process. The
specimens were long cylinders ( L / D > 25 ), thus ensuring one-dimensional heat flow in
the radial direction. Assuming a uniform air flow around the specimen, the boundary
conditions for the direct problem were symmetry at the centreline (r = 0) and heat
transfer by convection and radiation at the surface (r = R). The rods were initially kept
at a uniform temperature.
To illustrate the calculations, the estimated heat-transfer coefficient plotted against
computed steel surface temperature for an 8 mm-dia. rod, air-cooled at 22 m/s, is shown
in Figure 5.8, together with results obtained by Campbell et al. [42] using an iterative
procedure based on an implicit finite-difference approximation of the direct problem with
exact matching of experimental and computed temperatures (see [197] for thermophysical
data used in the calculations). Good agreement can be seen, although it is important
to note that the calculations based on the sequential function specification algorithm
exhibit less scatter. This result is a consequence of the more stable solution achieved
through the use of the concept of future time steps.

5.4 Application to Forced Convective Boiling

In order to study the behaviour of the inverse analysis code under similar conditions to
those expected in the laboratory trials, the thermal response of a thermocouple located
near the surface of a 38.1 mm (1.5 in) bar subjected to forced convective quenching was
The Inverse Heat Conduction Problem 97

modelled using a finite-element program. Constant thermophysical properties (repre-


sentative of steel) were adopted in the calculations; it was assumed that no heat was
generated inside the bar. The form of the surface heat flux functional used in the simu-
lations was :
' 1 + 0.003 (T, - 900) if T > 900
s

4 - 0.0075 (T - 500) if 900 > T > 500


s s

q=\ (5-21)
1 + 0.0075 (T - 100) if 500 > T > 100
s s

0.0133 (T, - 25) if 100 > T > 25


s

where q was in M W m - 2
and T in C. This form of q, as shown in Figure 5.9, was
s

adopted for the sensitivity analysis because of its similarity with preliminary laboratory
results. The data used in the calculations is given in Table 5.3. A total of 38 8-node
isoparametric elements were used with a time step, A t , of 0.01s; the results were printed
every 0.1 s. The calculated thermal response at the centre, surface, and a subsurface
position are shown in Figure 5.10. The solution at the subsurface position was adopted
as input to the IHCP algorithm to simulate an errorless measured thermal response. For
the inverse analysis, the specimens were subdivided into two regions of 1.5 and 17.55 mm,
discretized by 5 and 15 nodes, respectively. The time step adopted for the calculation
was 0.05 s.

A sensitivity analysis was conducted by varying the number of future time steps
(parameter r), the thermal conductivity, and the thermocouple position. The results are
summarized in Table 5.4. The last column in the table gives the difference, in percentage,
between the input and estimated maximum heat flux. The difference between input and
estimated heat flux can also be characterized by the root mean square (rms) of the error :

where q and q are the actual and estimated surface heat flux values, respectively.
The Inverse Heat Conduction Problem 98

The estimated surface heat flux, as a function of surface temperature, is shown in


Figure 5.11 for three values of r : 2, 4, and 6. Good agreement can be seen, except at
early times and for the peak value of heat flux. The maximum heat flux is underestimated
in all cases, with the difference increasing as r increases (from a -4.25 % difference for
r = 2 to - 11.75 % for r = 6). The peak heat flux is underestimated because in the
vicinity of the surface temperature corresponding to q max the algorithm anticipates the
drastic change in slope sign and starts to decrease earlier. However, the sudden change
in the surface heat flux is followed well. It should be noted that only the results obtained
with r 2 showed the early decreasing trend in the heat flux at high temperatures (T > s

900 C ) . For the case of r = 1, i.e., an exact matching of the thermal response at the
thermocouple position, an unstable solution was observed. In contrast to the heat flux
results, the surface temperature at which the maximum heat flux is observed differs by no
more than 3 C (a difference of only 0.6 %) for the three values of r used; the estimated
surface temperature, as a function of time, is shown in Figure 5.12.
To study the sensitivity of the analysis with respect to the thermophysical proper-
ties, the thermal conductivity adopted for the inverse analysis was allowed to vary from
- 20 % to + 20 % of the value used in the finite-element simulation, while using the ther-
mal response obtained with the base value as input to the program for all cases. The
estimated surface heat fluxes (for r = 4) are shown in Figure 5.13. A higher value of
the thermal conductivity results in a higher maximum heat flux and the peak position
was shifted to higher surface temperatures, while the opposite was observed when the
thermal conductivity was lowered.

The effect of using a wrong estimate of the thermocouple position was also studied.
To this effect, three values of thermocouple position were used : 16.55, 17.55, and 18.55
mm (the base value used in the finite-element simulation was 17.55 mm); the simulated
response using the base value was adopted for input to the inverse analysis in all cases.
The estimated surface heat flux, as a function of surface temperature, for the three cases
The Inverse Heat Conduction Problem 99

(with r = 4) is shown in Figure 5.14. The effect of adopting a thermocouple position


closer to the surface than the actual one, is similar to using a higher value of thermal
conductivity; but, for the parameter levels adopted, is less pronounced.

Finally, the simulated thermal response at 3 positions in the cylinder (15.55, 16.55,
and 17.55 mm from the center) was used to estimate the surface heat flux. The results
are shown in Figure 5.15. The results show that a better estimate is obtained as the
sensor is positioned closer to the surface. The rms of the error in the estimated heat flux
decreases from 0.0552 at 15.55 mm to 0.0402 at 17.55 mm.
From the previous results, it is evident that care should be taken in the selection of
parameters such as future time steps, thermophysical properties and sensor position. A
discussion on filtering noisy data before applying the inverse analysis is left for a later
chapter.

5.5 Sensitivity coefficients and experimental design

The sensitivity coefficients, Xj, are calculated as part of the solution to the IHCP. They
represent a quantitative measure of the sensitivity of the thermal response to changes
in the unknown surface heat flux (see Eq. (5.4)), and can, therefore, be used to design
experiments. In general, one is interested in large, uncorrelated values of Xj, which
means that more non-repetitive information can be extracted from the experimentally
determined thermal response. Figure 5.16 shows calculated sensitivity coefficients at sev-
eral positions across the radius of a solid cylinder subjected to a constant heat-transfer
coefficient and fluid temperature. It can be seen that the largest value of the sensitivity
coefficient occurs at the position closest to the surface of the cylinder; for this reason, ev-
ery effort should be made to place the thermocouples as close to the surface as practically
possible. Lambert and Economopoulos [199] established a linear relationship between the
logarithm of the mean error of the determined heat-transfer coefficient and the distance
The Inverse Heat Conduction Problem 100

of the point of measurement below the surface for cylindrical samples. They obtained a
value of the mean error as high as 10 % for a position 2 mm below the surface, which
decreased to less than 1% for a measurement taken at 1 mm below the surface, when long
cylinders of 20 mm-dia. were used. This is a direct result of the lagging and damping
effects characteristic of transient diffusion problems. In quenching, where steep thermal
gradients are present at the early stages, the damping of the transient temperature re-
sponse would have an even more adverse effect on the estimation of the surface heat flux
from temperature measurements.
The Inverse Heat Conduction Problem 101

Table 5.1: Parameters used for Case 1 : constant surface heat flux.

Parameter Value
R 5 xlO m- 2

a 0.00125 m s- 2 1

50 W m - 2

k 25 W m " K - 1 1

Cp 200 J k g " K - 1 1

P 100 kg m - 3

Table 5.2: Parameters used for Case 2 : constant heat-transfer coefficient.

Parameter Value
h 160 W m ~ K - 2 1

Tf 20 C
T 0 850 C
R 5 mm
k 25 W m - K 1 _ 1

Cp 625 W m - K " 1 1

P 7650 kg m ~ 3
The Inverse Heat Conduction Problem 102

Table 5.3: Parameters used for the simulation of a forced convective quenching experi-
ment.

Parameter Value
R 19.05 mm
k 25 W m - K - 1 1

P 7650 kg m - 3

Cp 625 J k g - K - 1 1

T 1000 C
* 17.55 mm
r

Table 5.4: I H C P algorithm : summary of sensitivity runs. The base values for thermal
conductivity and thermocouple position were k =25 W m K and r* = 17.55 mm,
0
_ 1 _ 1

respectively.

r k T / C position ^max T ^rms % A 9 m a x

(W m K - ) 1
(mm) (MW m" ) 2 s,max
2 k 0 K 3.83 497.0 0.0203 -4.25
4 k 0 K 3.67 500.3 0.0402 -8.25
6 k 0 K 3.53 502.5 0.0616 -11.75
1 k 0 < - - - -

4 k - 20 %
0 K 3.44 471.6 0.0979 -14.0
4 k + 20 %
0 K 3.82 523.8 0.0527 -4.5
4 k 0 K-i 3.63 501.6 0.0491 -9.25
4 k 0 K + l 3.64 527.0 0.0471 -9.0
4 k0 r* = 16.55 3.65 498.8 0.0463 -8.75
4 k0 r* = 15.55 3.64 475.8 0.0552 -9.0
The Inverse Heat Conduction Problem 103

B.C. 1 B.C. 2

Y(t)
q(t) ='
X
1

x= 0

(a)

o
o
OO o o

(b)
Figure 5.1: (a) Schematic representation of a one-dimensional, single-sensor IHCP in a
flat plate of thickness 2L; the sensor is located at position x\. The boundary conditions
are : at x 0, symmetry (dT/dx = 0); and at x = L, unknown time-dependent heat
flux, (b) Discrete temperature measurements (Y(ti)) at position x\.
The Inverse Heat Conduction Problem 104

LL
CO
CD
I
CD
O
CO

CO

0 t. t t M + r-1
M-1 M

Time

Figure 5.2: Piecewise aproximation of the surface heat flux as a function of time. The
constant heat flux function between tu-\ and tM+r-i is adopted to calculate q in the
M

sequential function specification algorithm [192].


The Inverse Heat Conduction Problem 105

( START )

Read input data and


initialize variables

*M+i-i ^_ Q ) \<i< r

Ti > 1 < i<r

* M+t-l
Xj , 1 < %< r

rpM

t >t 7
( END

n
Af <- Af + 1

Figure 5.3: Flow chart of the sequential function specification algorithm adopted for the
solution of the IHCP.
The Inverse Heat Conduction Problem 106

70
i 1 1 1 r

60 h

o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o

input
o estimated

0 J I L

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4


Time, s

Figure 5.4: Comparison of estimated and input surface heat flux for Case 1 : constant
surface heat flux.

150

Figure 5.5: Comparison of estimated, using the sequential function specification tech-
nique, and analytical thermal response at the surface of the cylinder for Case 1 : constant
heat flux.
The Inverse Heat Conduction Problem 107

200

160
.o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o-o

120

'a
80
Input
o Estimated
40

0*
10 15 20 25
Time, a

Figure 5.6: Comparison of estimated and input heat-transfer coefficient for Case 2
constant heat-transfer coefficient.

900

Analytical
850 o Estimated

800

V 750

700

650

600
10 15 20 25
Time, s

Figure 5.7: Comparison of estimated, using the sequential function specification tech-
nique, and analytical thermal response at the surface of the cylinder for Case 2 : constant
heat-transfer coefficient.
The Inverse Heat Conduction Problem 108

150 L - ' ' ' ' '


1 1 1 1 1 1 1 1 1 t 1 1 1 1

600 650 700 750 800

Figure 5.8: Estimated heat-transfer coefficient for 8 mm-dia. steel rods air cooled at 22
m/s; calculated by using a sequential matching (SM) approach (open circles) and using
the sequential function specification (SFS) algorithm (closed circles). The solid curve is
a best-fit curve using all the points [197].

4.5 h

o
T urf>
S C

Figure 5.9: The functional q = f(T ) adopted for the finite-element simulation of the
s

thermal response in a 38.1 mm-dia. cylinder subjected to forced convective quenching.


The Inverse Heat Conduction Problem 109

1000,

800

600

0> 400

200

100

Figure 5.10: Calculated thermal response at the centreline, surface, and simulated ther-
mocouple position (r* = 17.75 mm), obtained when the heat flux distribution shown in
Figure 5.9 was applied at the surface of a 38.1 mm-dia cylinder.

surf
1

Figure 5.11: Effect of varying the parameter r adopted in the inverse analysis on the
estimated surface heat flux during the simulated forced convective quenching of a 38.1
mm-dia. cylinder.
The Inverse Heat Conduction Problem 110

1100

FEM -

o r = 2
V r = 4
o r = 6 -

30

Figure 5.12: Effect of varying the parameter r on the estimated surface temperature
response during the simulated forced convective quenching of a 38.1 mm-dia. cylinder.

i 1
i 1
r

r = 4

o input
k = k Q

-- k = k - 20 %
k = k + 20 %
o
_i_ _1_
200 400 600 800 1000
T
surf C

Figure 5.13: Effect of varying the value of thermal conductivity adopted in the inverse
analysis on the estimated surface heat flux during the simulated forced convective quench-
ing of a 38.1 mm-dia. cylinder. The base value was k = 25 W m K . 0
_ 1 _ 1
The Inverse Heat Conduction Problem 111

0 200 400 600 800 1000

T
surf C

Figure 5.14: Effect of varying the value of the thermocouple position adopted in the
inverse analysis on the estimated surface heat flux during the simulated forced convective
quenching of a 38.1 mm-dia. cylinder. The base value was r = 17.55 mm.0

1000

surf

Figure 5.15: Effect of the thermocouple position on the estimated surface heat flux during
the simulated forced convective quenching of a 38.1 mm-dia. cylinder.
The Inverse Heat Conduction Problem 112

Figure 5.16: Sensitivity coefficients at several radial positions in a solid cylinder subjected
to a medium of constant heat-transfer coefficient and fluid temperature.
Chapter 6

Stress M o d e l

To model the development of the stress field during the quenching operation, a continuum
mechanics approach was adopted. The analysis was restricted to small deformations,
based on the fact that the deformations caused by thermal and transformation strains
are likely to be moderate and no external forces are applied. Given that the bars spend
a relatively small time at high temperatures, rate-dependent effects were not considered.
In this chapter, the stress-strain relationships for inviscid elastic-plastic materials
with work-hardening, including thermal and microstructural effects generated during
quenching, are first derived. To obtain the equations, the classic incremental theory of
plasticity (a loading-path-dependent formulation) is invoked. Then, the finite-element
implementation of these relationships is discussed and two instances of code verification
are presented.

6.1 Stress-Strain Relations

To illustrate typical material behaviour in the plastic regime, the case of uniaxial loading
shall be examined first. Referring to Figure 6.1, the material behaves elastically upon
loading until the yield stress (point A) is reached; beyond the yield stress there no
longer exists a linear relationship between stress and strain but instead the slope of
the stress-strain curve decreases monotonically until, eventually, fracture occurs. If a
test is conducted where a specimen is loaded up to point B and then unloaded, the
unloading path (BC) will have a slope similar to that of the elastic path but the strain

113
Stress Model 114

at zero stress has a finite value. This residual, or irrecoverable strain, is termed the
plastic strain (OC), whereas the recoverable strain (CD) is the elastic strain component
of the total strain. Upon reloading (CBE) the material will behave elastically until the
previous stress corresponding to point B is reached (this is termed the subsequent yield
stress) after which the plastic deformation occurs according to path BE. From the figure
it can be observed that the subsequent yield stress of a real material increases as induced
plastic strain increases; this effect is known as strain hardening or work-hardening.
From the previous discussion it is clear that the mechanical response is a function of
both the current state of stress and the loading history, or loading path. Thus, the classic
incremental flow theory of plasticity is adopted to describe the stress-strain relationship
during plastic loading under a multiaxial state of stress.
From the experimental evidence presented above, the fundamental assumption that
the individual components of the incremental strain can be separated is expressed as

d e ^ d ^ + d ^ + de?- (6.1)

where

deij = total strain increment

de?-
- elastic strain increment

del = plastic strain increment

de non-mechanical strain increment

For cooling during heat treatments, the non-mechanical strain increment has contribu-
tions from thermal- and transformation-related strains. In particular, strains due to
volume changes associated with cooling (de*^) and phase transformations (de*^), as well
as those arising from transformation plasticity (de'J), variation of the elastic constants
Stress Model 115

with temperature (defj ), and variation of the flow stress with temperature(deg ) need
E F

to be considered :

de?- = def + de% + de% + deg* +


3 teff (6.2)

The first four strain increments are present in both the elastic and plastic regimes while
the last one is nonzero only during plastic flow. Accordingly, they can be grouped as
follows :
ds = def + de?
l3 (6.3)

where
def = d e j + de* + de* + deg* (6.4)

and
def = de?/ (6.5)

In the following, the fundamental characteristics of material behaviour in a uniaxial


load test, as described above, are generalized to predict the mechanical response under a
multiaxial stress state, taking into account thermal- and transformation-related strains.
The discussion is separated into elastic and plastic regimes.

6.1.1 Elastic Stress-Strain Relations

To model elastic loading, the material is assumed to be homogeneous and isotropic.


Further, a linear relationship between stress and strain is adopted. Such a relationship,
for uniaxial loading, is provided by Hooke's law. In the more general case, the so-called
generalized Hooke's law is invoked :

Oij = Dijki (e ! - e f)
k k
(6.6)
Stress Model 116

or, in incremental form


do-;j = D^ki (de i - de f) k k (6.7)

where Dijki is the material elastic constant tensor (or stiffness tensor) given by

E I 2u SijSki + SikSji + SijSjk (6.8)


2(1+1/) [(1 -2u)

where E is the Young's modulus and v is Poisson's ratio.

6.1.2 Plastic Stress-Strain Relations

The elastic limit or yield stress under all possible combinations of stresses is defined as a
yield function :
F(aij,k k ,...)
u 2 - 0 (6.9)

where ki,k ,... 2 are material constants to be determined experimentally and the state of
stress, o~ij, is defined in terms of the invariants of the deviatoric stress tensor.
In particular, the von Mises yield criterion is widely adopted :

F(J ) = J -k
2 2
2
=0 (6.10)

where J is the second invariant of the deviatoric stress tensor and k is the yield stress in
2

pure shear. Geometrically, a plot of the yield function in stress space results in a surface,
while the stress state is represented by a point.
The yield function, Eq. (6.10), is modified to reflect the effects of temperature and

1
The Kronecker delta is denned as

l if i = j
0 if i j
Stress Model 117

transformation changes and includes isotropic work hardening as follows :

F = F(a ,K,T,X )
ij k (6.11)

where n(e ) is a hardening parameter and e is the effective plastic strain.


p p

In order to apply the loading function to a real problem, the hardening parameter
must be related to experimental uniaxial stress-strain curves. This is done through the

definition of the effective stress and the effective plastic strain. A n expression for the

effective stress is derived from the results of a uniaxial test :

(6.12)

The effective plastic strain is given by

p = \ l 4 4
d D D
( -
6 1 3
)

The effective stress, <7, is related to the effective plastic strain by


e

cr = a (e )
e e p (6-14)

Differentiating :
dcr = H (a )de
e p e p (6.15)

where Hp(e ) = da /de


p e p is the plastic modulus. For a generalized state of stress, the

plastic modulus is related to the rate of expansion of the yield surface, while in a uniaxial

test it represents the slope of the stress-plastic strain curve at the current value of a . e

The effective strain path gives the strain history of the material. Its length at any given

point is given by
p=
I dp=
Iint) -
(6 16)
Stress Model 118

The plastic strain increment under a generalized state of stress can be computed by
invoking the flow rule :

de?- = d A ^ (6.17)
every-
where g(o~ij) is the plastic potential function and dA is a positive proportionality constant,
termed the loading parameter, which is nonzero only when plastic deformation occurs.
The direction of the plastic strain increment tensor, de?-, is defined by the gradient of
the plastic potential surface, Bg/Baij, while its magnitude is determined by the loading
parameter. A special case of the flow rule is the so-called associated flow rule, in which
the yield function and the plastic potential function coincide, i.e., g = F. Thus

dF
de*- = d A - (6.18)
Oaij

Adopting the associated flow rule, the elastoplastic stress-strain relation is derived as
follows. Differentiating F in Eq. (6.11) results in

BF BF BF BF
d F = ^-daa + ^-d + ^-dT + ^rdX (6.19)
Y
K k
daij BK BT d X k

Given that the hardening parameter, K, can be represented by the amount of plastic work
done to the solid during plastic deformation the chain rule can be used to obtain

BF BF BK
J^ = d^W,^'
iK 1
'
(6 20)

Following the consistency condition 2

dF = 0 (6.21)

2
The consistency condition ensures that the stress state remains on the yield surface.
Stress Model 119

then
dF , dF 8K , _ dF :rT1 ^ dF n
(6.22)
d^- '' /3K del V d T

The stress increment based on the elastic strain increment is :

der,- Dijki de (6.23)

D.jW (deij - de ? - de th Apt r

= ij U f c
i j
de'? de-)

Invoking the associated flow rule Eq. (6.18), substituting into the consistency condition

and solving for dA results in :

dF
dA = S-1
D ijkl (de - de% - de% - de% - de
i3
daij
U
(6.24)
dT

where
dF du dF
< i - d
nF d F
(6.25)
o doij L>ijkl da^ du de^j da^

A loading criterion under a multiaxial state of stresses can now be derived. Referring

to the consistency condition

n
Dijki (A de -j A) -+L dT
[deij - A A T +X \^
- AY
-dX SdX = 0 (6.26)
d F D F d F

8 k

a.

For a work-hardening material it can be shown that the scalar function S is always
positive. Given that dA is a non-negative quantity (for plastic loading), it then follows

that S'dA is always positive. Plastic flow begins when the stress point coincides with the

yield surface and continues, if and only if, the stress incremental vector, daij, is directed

outwards from the current elastic region. If the stress increment vector is parallel to the

yield surface, no additional plastic deformation will occur (i.e., neutral loading); while
if it points inward from the current yield surface, unloading will occur. Defining the
Stress Model 120

loading criterion function, L , by

L = L Diju ( d , _ d 4 / ) + | f dT + E -AX k (6.27)

the following loading criterion is obtained :

> 0 loading > elasto-plastic constitutive eqn.


= 0 neutral loading > elastic constitutive eqn. (6.28)
< 0 unloading > elastic constitutive eqn.

Finally, the elasto-plastic thermo-microstructural constitutive equation is obtained :

= de - A7 ( D
4 + D
4 + D
? + K) E

-Dljkl (j^dT + | ^ d
^ ) ( -29)
6

Note that the last term is different from zero for plastic deformation only.
In E q . (6.29),

D% = D M ijkl - D p
jkl (6.30)

is the elasto-plastic matrix and

Dijkl = HH S- a' Dij


D l
i3 kl a'ij (6.31)

is the plasticity matrix.

The constitutive equation can be written in the following, more compact, form

d(T - = D?
0 itA (de,, - de?-) (6.32)

where
de?- = de# + de* + deg + de?* + d e , f (6.33)
Stress Model 121

d e g = a* (D^Y
F 1
(^dT + ^-dX ^j
k (6.34)

and
0 if elastic
a = < (6.35)
1 if plastic

Note that

Dijki ^ Diju {o-ij) (6.36)

while

l%i = Oik (^) (- )


6 37

6.1.3 Finite-Element Formulation

Due to the material nonlinearity in Eq. (6.32), an analytical solution for real problems is
very difficult to obtain. Instead, a numerical method must be adopted. In this study, a
computer program based on the finite-element method was used. The basic components of
the finite-element formulation are : 1) finite-element equations and 2) solution algorithm
for the nonlinear equations.

6.1.3.1 Finite-Element Equations

In the finite-element solution of stress problems, the displacement field at the element
level is approximated by the following relationship :

{u} {u} = [N]{ay (6.38)


Stress Model 122

where {u} = [u,v] is the displacement vector for a two-dimensional problem, [N] is the
T

shape function matrix and {a} is the nodal displacement vector. Note that, in contrast
3 e

with the thermal analysis, there are now two degrees of freedom (u and v) at each nodal
point.

Also, the strains and nodal displacements are related by

{e} = [L]{u} = [L][N]{ay (6.39)

where, for a solid of revolution,

d_
dr
8_
[L] 1/7 (6.40)
d_ f
dz dr

Let us define the strain-nodal displacement matrix by

[B] = [L][N] (6.41)

then
{e} = [B}{aY (6.42)

By invoking the principle of virtual displacements, a weak form of the equilibrium


equations can be obtained, from which, upon substituting Eq. (6.42) and introducing a
general constitutive equation, the following equivalent form of the governing equation is
obtained 4

[ATW = e
{RY (6.43)

3
Linear-order (4-node) and quadratic order (8-node) shape functions applicable to two-dimensional
isoparametric elements can be found in Ref. [185].
4
Details of the derivation are given in Appendix C.
Stress Model 123

where, for the axisymmetric case, the stiffness matrix, [K] , and the load vector, e
{R} , e

at the element level are given by 5

[K] e
= 2TT / [B) [D][B]r dr dzT

JA e

2irr [ [B] [D][B] dr dzT


(6.44)
JA e

and

{R}e
= 2TT I [B] [D}{e}r T
dr dz
JA e

27rf / [B] [D]{e } dr dz


T 0
(6.45)
Me

For the problem in hand, there are five components of the nonmechanical strain
(Eq. (6.2)). Appropriate expressions for each of these terms are given in Appendix D.

6.1.3.2 Solution A l g o r i t h m

Due to the nonlinear relationship between stress and strain, the governing equation
(Eq. (6.43)) is nonlinear on the nodal displacements. Moreover, since the elasto-plastic
constitutive equation depends on deformation history, the variation of displacement,
strain and stress should be traced along with the loads.
The current load vector is computed from 6

m+1
{R} = {R}
m
+ {AR} (6.46)

The balance between internal forces, { F } , and applied loads, {R}, is expressed in the

5
T o facilitate the integration, r is used instead of r.
6
In this subsection m + 1
() and () m
denote values at the current and previous load increment,
respectively.
Stress Model 124

equilibrium equation :

{F}
77V
{R} (6.47)

where

771+1
{F} = j [B] {a}dV
v
Tm+1

2nf f [B] {a}dr Tm+1


dz (6.48)
JA

In general, Eq. (6.47) is not satisfied, but rather a residual, ^ , is obtained :

*( m + 1
M) 771 + 1
{F( {u})}- {R}
m+1 m+1
=o (6.49)

The elastoplastic constitutive equation derived previously (Eq. (6.32)) provides a rela-
tionship between an infinitesimal strain increment and an infinitesimal stress increment,
and is adopted to compute the stress tensor at the Gauss points. In a finite-element
solution, however, a finite load increment (as opposed to an infinitesimal load increment)
is applied at each load step, which results in finite increments of stress and strain. Con-
sequently, the incremental constitutive relation given by Eq. (6.32) has to be integrated,
usually employing a numerical technique.
In virtually all the algorithms used to solve elastoplastic problems, the solution of
Eq. (6.47) at each loading step is carried out in two stages :
7

Stage 1 The current load increment, m+1


{AR}, is computed based on the thermal
and microstructural states at the end of the current loading step. The corresponding
displacement increment at the global level is obtained by solving Eq. (6.43) at the global
level.

Stage 2 A trial stress increment, {ACT }, is computed assuming elastic behaviour and
6

7
A loading step in the stress problem is equivalent to a time step in the thermo-microstructural
solution.
Stress Model 125

is used to determine the loading state. In the most general case where the Gauss point
has entered a plastic state from an elastic one, the trial stress increment is split into
its elastic and elastoplastic components by determining an scaling factor, r, such that
f( {cr} + r{Aa }, k)
m e m
= 0, see Figure 6.2. Then, the elastoplastic fraction of {A<r } is e

computed by integrating the elastoplastic constitutive equation (Eq. (6.32)).


In the first stage, the load increment is used to predict the stress state (assuming
elastic behaviour) while a 'correction' is applied in the second stage to obtain the actual
stress increment. For this reason,, the method described above is commonly termed
predictor-corrector [200].
In this study, the computer program developed by Nagasaka et al. [47] based on
the algorithm proposed by Yamada et al. [201] has been adopted. The algorithm is
summarized in the flowchart shown in Figure 6.3. The stress increment at each Gauss
point was calculated by first selecting r m i n (the minimum value of the scaling factor r)
and multiplying the load, nodal displacement, stress and strain increments by r min to
obtain the deformation state after the previously elastic element with r = r m i n has just
yielded. From then on, within the load increment, this element was treated as plastic.
The predictor-corrector method was applied in this manner for all elements (at which
point J2 i
r
m n = 1-0). Then, the deformation state was updated and the next loading step
calculated.

The scaling factor, r, can be calculated either analytically or numerically [159]. Fol-
lowing Yamada et al. [201], r was computed using :

r = (6.50)
2({Aa}Y

where

r = ({Aa}) - 2a Aa
2
e e - {Aa f
e
(6.51)
Stress Model 126

and

{ A ^ } = y^{A<7'}{A<7'} (6.52)

In the equations above, Acr , denotes the increment of effective stress induced by the
e

load increment, {Ai?}.


In order to speed up the computations, elements just prior to yield (<r > 0.995a"i?)
e

are considered to be plastic in the next cycle of calculation.

6.2 Verification of Mathematical Model of Stress Generation

As mentioned above, the finite-element code developed by Nagasaka et al. [47] to simulate
stress generation during spray cooling of steel bars was adopted to model stresses in forced

convective quenching. The elastic component of the code was verified by comparing
numerical predictions of thermal stresses in an infinite cylinder with constant mechanical

properties against an analytical solution. A verification of the elastoplastic component

of the computer code has been reported elsewhere [47].

6.2.1 Infinite Solid Cylinder : Elastic Thermal Stresses

In absence of body forces, and considering axial symmetry in the temperature distribu-

tion, the thermal stresses in an axisymmetric body are obtained from the equations of

equilibrium, which reduce to [202] :

?ZL + JLZL = o (6.53)


or r

with the following boundary conditions :

B.C. 1 u = 0 at r = 0 (6.54)
Stress Model 127

B.C. 2 : -4 =
L
0 at r = R (6.55)

The stresses are generated by a temperature gradient which is symmetrical about the z
axis and independent of the axial coordinate, z. Assuming constant thermomechanical
properties, an analytical solution for the radial (oy), circumferential (erg) and axial (a ) z

stress distributions can be derived by applying the method of strain suppression [ 2 0 3 ] to


obtain

= r^(f<
-->'
'-?r<
<-> rr d r 66
5

> = T ^ ( % 0 - ^ T T R D R
-( - ^) T T
<' >
6 58

The corresponding radial displacement is

u = [ ^ a ((1 - 2u) J\T - T r e f )rdr+ - 1


J\T - T r e f ) r dr) (6.59)

For a parabolic temperature gradient of the form

T(r) = T- 0 (Arf (6.60)

and taking T r e f = T , i.e., zero thermal strain at T T , the above equations reduce to
0 0

aE
a r = i ( f ^ y ( 4 A V - AR) 2 2
(6.61)

aE
cr = a
A4A r -A R -A r )
E 2 2 2 2 2 2
(6.62)
4(1 - u)
e A

aE
, ( 4 A r - 2vA R ) 2 2 2 2
(6.63)
4(1-u)

u = " ^ T ^ A
( ( 1
" 2
" ) J R 2 R + R 3
) ( 6
- 6 4 )
Stress Model 128

To assess the performance of the model, the elastic thermal stresses in an infinitely
long cylinder of 0.1 m diameter generated by a parabolic temperature gradient (T(r) =
100 (50r) ) were calculated. The cylinder was assumed to be initially at a uniform
2

temperature T = 100 C. The thermomechanical properties were assumed constant and


0

are given in Table 6.1. The finite-element mesh consisted of 10 4-node isoparametric
(2D4) elements. The results of the comparison between the radial, circumferential and
axial stress distributions predicted by the model and by the analytical solution are shown
in Figure 6.4. The applied temperature gradient is also shown in the figure. A very good
agreement between numerical and analytical predictions was observed. The correspond-
ing radial displacements are shown in Figure 6.5. The variation of radial displacement
with radial position was also properly predicted by the model.

6.2.2 Infinite Solid C y l i n d e r : Elastoplastic T h e r m a l Stresses

There are no analytical solutions for the complete quenching problem or even for elasto-
plastic stresses generated by thermal gradients without considering phase transforma-
tions. Thus, only comparisons between measured and predicted residual stresses can
be used to verify the models. The ability of the stress model to predict elastoplastic
thermal stresses was investigated by Nagasaka et al. [47]. In that work , the residual 8

stresses predicted by the model were compared with measured (using a boring-out tech-
nique [204,205]) residual stresses in a 50 mm-dia. pure iron bar quenched from 850 C in
ice-water. Since pure iron was used, the stresses were generated by thermal strains only.
The results are shown in Figure 6.6. The predicted residual stress distributions agreed
well with the published measured values. It should be noted that Mitter et al. [205]
found that plastic flow had occurred during their measurements, which would explain
discrepancies between their values and those reported by Buhler et al. [204].

8
For details of the calculations, the reader is referred to the original reference [47].
Stress Model 129

6.3 Summary

A computer program developed by Nagasaka et al. [47] has been adopted to simulate
stress generation during forced convective quenching of steel bars. The code is based
on a finite-element formulation of the thermal-microstructural elastoplastic problem, in-
cluding non-mechanical strains associated with temperature gradients, microstructural
evolution, changes in elastic properties and flow stress with temperature and phase com-
position, and transformation plasticity. The thermal-microstructural elastoplastic con-
stitutive equation was derived adopting the classical incremental theory of plasticity,
considering isotropic hardening and a J2 (von Mises) material behaviour. The code has
been verified by modeling 1) elastic thermal stresses in an infinitely long cylinder sub-
jected to an axisymmetric temperature gradient (assuming constant thermomechanical
properties), and 2) thermal stress generation in a pure-iron bar quenched in ice-water [47].
Stress Model 130

Table 6.1: Input data used for the comparison of finite-element and analytical solutions
for elastic stresses generated in an infinite solid cylinder by a temperature gradient.

Variable Value
R 0.1 m
E 1.25 x l O M P a
5

V 0.3
a 2.0 x l 0 ~ C
5 _ 1

T(r). 100 - (50r) 2

0 C D

Figure 6.1: Schematic of loading and unloading paths for a work-hardening material
under a uniaxial load.
Stress Model 131

Figure 6.2: Schematic representation of the scaling factor, r [159].


Stress Model 132

( START )

Read input data


Initialize variables

Read 7 7 1 + 1
T, 7 7 1 + 1

Initialize variables for r m i n calc"

Calculate m+1
{AR}

Solve [K]m+1 m + 1
{Aa} = m+l
{AR}

Compute m+1
{Aa } e

Calculate r

Compute r^r+HAi?}]

Compute 7 7 1 + 1
{Aa}

n r = l ?

t <- < + At

n
* > 'end ?

( END

Figure 6.3: Flow chart of the stress solution algorithm.


Stress Model 133

-40 60
0.00 0.02 0.04 0.06 0.08 0.10

Radial Position, m

Figure 6.4: Comparison between analytical and numerical ( F E M ) predictions of stress


distributions in a 100 mm-dia. cylinder, produced by the thermal gradient shown as a
broken line.

0.00

o
-o.oi

d
V

a
o -0.02
CO

Analytical
O FEM
-0.03

0.00 0.05 0.10

Radial Position, m

Figure 6.5: Comparison between analytical and numerical ( F E M ) predictions of radial


displacements in a 100 mm-dia. cylinder, produced by the thermal gradient shown as a
broken line in Figure 6.4.
Stress Model 134

Figure 6.6: Comparison between measurements made by Buhler [204] and Mitter et al.
[205] and numerical ( F E M ) predictions of residual stress distributions in a 50 mm-dia.
pure-iron bar quenched in ice water from 850 C [47].
Chapter 7

Laboratory Experiments : Quenching Tests

The layout and operation of the industrial equipment allowed little flexibility for instru-
mentation to measure the thermal response of the steel rods during the quenching cycle.
This information is needed to estimate the heat transfer boundary condition. Thus, a
laboratory facility was designed and built to simulate the industrial operation. In this
chapter, the apparatus and experimental procedure adopted to study heat transfer in
forced convective quenching of steel bars are described.
The objectives of the laboratory experiments were twofold : 1) to characterize the
surface heat flux as a function of surface temperature, under forced convective boiling
conditions, and 2) to produce specimens for metallographic characterization and resid-
ual stress measurement (to validate the mathematical models). To accomplish the first
objective, an interstitial-free (IF) steel was chosen. The material selection was based
on the fact that, during quenching, the IF specimens transformed to ferrite only. Thus,
the effect of the heat evolved during the transformation was confined to relatively high
temperatures, facilitating the application of the inverse analysis (see Chapter 9). The
chemical composition of the IF steel is given in Table 7.1. In order to have as complete a
boiling curve as possible, the initial test temperature in these experiments was 1000 C.
Alta Steel furnished three alloyed, near-eutectoid steels, in the form of 100 mm (4 in)-dia.
bars. The chemical composition of the three steels is given in Table 7.2. One of these
steels (steel A) was quenched in tests aimed at producing data to validate the model. In
these runs, the test conditions (water temperature and austenitizing temperature) were

135
Laboratory Experiments : Quenching Tests 136

chosen to simulate the operational conditions. However, the hardenability of this steel,
combined with the smaller bar diameter used, produced through-hardened samples and
even cracked specimens (at high water velocity/low water temperature). Thus, 1045 car-
bon steel specimens, which have a lower hardenability, were also quenched. The chemical
composition of the 1045 carbon steel is given in Table 7.3.

7.1 Quenching Apparatus

For the laboratory runs it was desirable to have the maximum control possible. Thus, the
quenching apparatus was designed to maintain the sample stationary during the heating
and quenching cycles. In addition, the initial test temperature had to be uniform across
the bar diameter. With these goals in mind, the sample was heated inside a quartz tube
by induction; the quartz tube also contained the flowing water during the quench. A
600 m m long, 46 mm I.D. ( 50 mm O.D. ) quartz tube was selected based on the largest
inside diameter that would allow for a practical design. The water inlet end of the quartz
tube was connected to the rest of the system by an aluminum bronze coupling.
A schematic drawing of the test section is shown in Figure 7.1; it consists of a hollow
tapered extension (screwed to the front of the sample), a sample, an adaptor (screwed to
the back of the sample), and a holder. Once assembled, the adaptor, specimen and ex-
tension formed a single piece. The sample geometry is schematically shown in Figures 7.2
(IF steel) and 7.3 (alloyed and 1045 carbon steel). The specimens were instrumented with
two type K , Inconel sheated, 0.3 mm (0.010 in) thermocouple wires . For this purpose,
1

60 m m long thermocouple holes were drilled parallel to the axis of the test bars. In the
case of the IF specimens, one of the thermocouples was positioned at the centreline while
the other was located 1.5 mm below the surface. The subsurface thermocouple was used
to estimate the surface heat flux through the solution of the IHCP and, therefore, needed

1
The recommended range of operation for type K thermocouples is from -200 to 1260 C [206].
Laboratory Experiments : Quenching Tests 137

to be located as close to the surface as possible. For the alloyed eutectoid steels and
1045 carbon steel samples, the second thermocouple was placed at a position halfway be-
tween the surface and the centre (in order to minimize the possibility of crack formation).
The thermocouples were spring-loaded to maintain good contact with the specimen; the
spring mechanism was attached to an adaptor (Figure 7.4) that was screwed to the sam-
ple. To lessen turbulence, and insure a fully developed velocity profile at the point where
the fluid comes into contact with the hot sample, a tapered extension made of mild steel
(Figure 7.5) was attached to the water inlet end of the sample. In order to eliminate
specimen distortion at high temperatures (caused by excessive weight) a hollow extension
was used; moreover, this extension was fitted with four supporting points in the quartz
tube. The fully assembled test section was rigidly held in position by a 1 m long, 25.4
mm I.D., 31.25 mm O.D. carbon steel tube mounted on a mechanically controlled test
bed. The end of the steel tube was sealed with a rubber bung held in place by a brass
gland nut. Two slits were made in the rubber bung to allow the thermocouple leads to
pass through it.
The specimen dimensions were determined based on hydrodynamic considerations,
including entry effects. The relevant dimensionless number to insure dynamic similarity
is the Reynolds number :
Re = ^ (7.1)
t1

where v is the velocity of the fluid, L is a characteristic length, p is the fluid density and
p is the fluid viscosity. Both, p and p, should be computed at the mean temperature be-
tween the rod surface and the fluid. For annular flow the equivalent (hydraulic) diameter
is :

D, - ^ (7.2)

where A is the cross-sectional area perpendicular to the flow and P w is the wetted
Laboratory Experiments : Quenching Tests 138

perimeter. Then,

and the Reynolds number takes the form :

where D\ and D are the bar and inside tube diameters, respectively. In terms of the
2

volumetric flow rate, Q, one obtains :

A(D,-D )Q
1

with v being the dynamic viscosity (p/p).


Typical operational plant data (prototype) are given in Table 7.4. Under those con-
ditions :
#e = 4 . 1 x l 0 6
i.e., turbulent flow (7.6)

To select the sample diameter, D\, the water velocity corresponding to several com-
binations of volumetric water flow rate and bar diameter was computed with D 2 fixed
at 46 mm, and compared with the water velocity used in the industrial operation (5.2
m s ) . The results are summarized in Figure 7.6. The nominal volumetric flow rate,
_1

rated at 32 m (105 ft), of the Goulds submersible pump (model 70 L G 30) used in this
study is also shown in the figure. As can be seen in Figure 7.6, the available pump would
not yield high enough velocities for Di 25.4 mm (1 in) or 31.8 mm (1.25 in); thus a
sample diameter of 38.1 mm (1.5 in) was selected. For a water velocity of 5.2 m s _1
(the
same as in the plant), the corresponding Reynolds number for the laboratory unit was :

Re = 3 . 2 x l 0 5
i.e., turbulent flow (7.7)
Laboratory Experiments : Quenching Tests 139

A n important design criterion results from considering entry effects. To determine


whether the thermocouples were located far enough downstream to allow for fully devel-

Calculated values of the hydrodynamic entry length obtained using Eq. (7.8) are shown
in Figure 7.7. For the target velocity of 5.2 m s _1
and a Reynolds number of 3.2 x l O , a
5

value of Lhyd = 14.8(1)2 Di) = 117 mm (4.6 in) was obtained. It should be noted that
the edge in the tapered extension may cause a separation of the boundary layer [208]
and, therefore, must be considered in the design, i.e., the hydrodynamic entry length
should be measured from the edge, rather than from the tip of the tapered extension.
To avoid thermal entry effects the distance from the thermal entrance to the point of
temperature measurement needs to be greater than the thermal entry length, defined as
the distance downstream of the thermal entrance necessary for the local Nusselt number
to fall to within 5 % of its fully-developed value. The thermal entry length for Reynolds
numbers in the range 10 < Re < 10 and Prandtl numbers in the range 0 < Pr < 10
4 6 4

has been computed by Notter and Sleicher [209] by solving, numerically, the energy
equation describing the heat transfer to a fluid in a pipe, i.e., the turbulent Graetz
problem. Their results, for uniform wall heat flux, are shown in Figure 7.8. Considering
that the Prandtl number for water varies from 1 to 10, for the temperature range of
interest, the curve for Pr = 3.0 can be taken as representative of the system. Then, the
thermal entry length for the Reynolds numbers of interest is confined to

LJDe <5 (7.9)

Yielding, L th < 39.5 mm (1.55 in). The assumption of uniform wall heat flux (as op-
posed to uniform wall temperature) has been adopted to estimate L t h because thermal
Laboratory Experiments : Quenching Tests 140

entry lengths calculated in this way are slightly longer than the corresponding values for
uniform wall temperature [209] and, therefore, represent a worst case scenario.

The flow loop of the system is schematically shown in Figure 7.9. In order to handle
the large water volumes required during the experiments, the quartz tube discharged
into a 95 1 (25 gal) drum mounted horizontally on top of a 170 1 (45 gal) drum. The
large drum acted as a water reservoir and housed the submerged pump; it was mounted
on a frame to stabilize the entire apparatus. The test bed referred to above was welded
to the inside wall of the small drum. During quenching, the water in the small upper
drum drained back into the large one through a 150 mm-dia. steel tube, creating a closed
circuit. To limit the unsteady-state effects associated with the start-up of the pump, a
separate water circuit was established through which the water could be diverted back to
the reservoir (valve # 2 open, valve # 3 closed). The water flow rate into the test section
was measured with a Fisher & Porter Series B10D1465 C O P A - X industrial magnetic flow
meter. Immersion heaters were used to heat the water in the reservoir to the required
temperature.
To prevent oxidation during heating, provisions were made to allow for evacuating
the test chamber and passing H over the sample surface. To accomplish this, a gas tight
2

cap (Figure 7.10) was used to temporarily close the discharge end of the quartz tube.
The cap was made of aluminum and fitted with a stainless steel tube that served as gas
inlet. O-rings were placed in grooves machined along the surfaces of the aluminum cap
in contact with the quartz tube and the sample holder to ensure gas-tightness. Another
O-ring was placed at the back face of the seal to prevent direct contact with the quartz
tube. When quenching was initiated, the aluminum cap was pushed to the back of the
small drum by the flowing water.

The data acquisition system consisted of a Metrabyte 16-channel multiplexer ( E X P -


16/A) and a Metrabyte A / D converter and time/counter board (DAS-800). The data
acquisition software was Labtech Notebook (v 7.2.1) running on a P C - A T . The sampling
Laboratory Experiments : Quenching Tests 141

rate was 5 H z . A schematic diagram of the data acquisition configuration is shown in


2

Figure 7.11. The output signal from the magnetic flowmeter was converted from a 4 - 20
m A span to a 0 - 50 m V range by soldering a resistance to the corresponding channel in
the multiplexer. The gain in the multiplexer was set to 50.

The flowmeter was calibrated by measuring the volume of water collected in a barrel
for recorded time intervals and at the same time recording the voltage drop across a 330
0 resistance. Figure 7.12 shows the calculated volumetric flow rate (1 s ) and the output
_1

signal from the flowmeter (mA) for various valve settings, the latter given as number of
turns with respect to the fully closed valve position. The calibration curve for water flow
rate as a function of output signal from the flowmeter, in the range of water flow rates
equivalent to 2 - 8 m s , is given by :
3 _1

Q = - 3.9 +0.5867 i (7.10)

where Q is given in 1 s _1
and i is input in m A . This expression was programmed in the
data acquisition software to convert and record the output signal from the flowmeter.
Heating was carried out with a 11-turn, 100 mm I.D., 125 mm O.D., 180 mm long,
water-cooled copper coil, fabricated with a 12.7 mm O.D., 1.5 mm thick copper tubing.
The coil was powered by an Inductotherm induction furnace (model Inducto 15) operating
at approximately 10kHz. The relatively low frequency resulted in deep penetration of
the magnetic field and, therefore, more uniform heating. To reduce radiative heat losses,
a fiberglass tube was placed between the quartz tube and the coil. During preliminary
tests it was found that, in order to attain a reasonably uniform temperature across the
specimen diameter, heat losses by conduction to the tip and to the adaptor had to be

2
The time constant of the thermocouple was estimated following a procedure outlined by Hernandez-
Avila [210]. Its value was of the order of x l O s.
- 2

3
Water velocity is related to water flow rate by Q = vA, where Q is water flow rate in 1 s , v is
-1

velocity in m s , and A is area normal to the flow in m .


_1 2
Laboratory Experiments : Quenching Tests 142

minimized. To accomplish this, freshly heated saffil was placed at both ends of the
4

specimen.

Typical heating cycles obtained with this arrangement are shown in Figure 7.13 for
two initial test temperatures, 1000 and 850 C. In the first case, the power supplied
to the coil was maintained at 4 k W up to ~ 900 C and then the power was varied
(following the thermocouple readings) to reach the initial test temperature smoothly. In
the 850 test, the power was reduced after reaching ~ 500 C . Very good reproducibility
was observed in all cases. The difference between subsurface and centre temperature at
the end of the heating cycle, for the IF steel experiments, is plotted in Figure 7.14. As
can be seen, the heating cycle produced a uniform temperature field at the end of the
heating period, the temperature difference between the centre and subsurface being less
than 10 C .

7.2 Procedure

Prior to each test, the thermocouples were calibrated by immersion in boiling water
and the test section was assembled . Once the water in the reservoir was heated to the
5

desired temperature, the test section, quartz tube and induction coil were set in place
(making sure they were concentric to each other) and the aluminum cap was attached to
the discharge end of the quartz tube. To facilitate the ejection of the cap, silicon grease
was applied to the external surface of the quartz tube. The computer was then started
and the thermocouples were connected to the data acquisition system and checked for
functionality. The pump was started with the valves directing the water flow back to

4
Pure alumina fibers.
5
Due to the required number of tests and the difficulties associated with machining the specimens,
the same specimen was used for several experiments. To restore the original surface condition, previously
used specimens were clamped in a lathe and polished with emery cloth.
Laboratory Experiments : Quenching Tests 143

the reservoir. The test chamber was then evacuated with a vacuum pump before back-
filling it with hydrogen. The induction furnace was turned on and the sample heated
to the initial test temperature and held for approximately 3 min. When the desired
6

austenitizing condition was attained, the hydrogen flow into the system was stopped, the
induction furnace turned off, the thermocouple leads connected to the multiplexer and
the data acquisition started. The quench was initiated by diverting the water flow into
the test chamber. Data was recorded until the temperature at the centre of the specimen
had decreased to the water test temperature.

A total of 9 tests were conducted to study the effect of water temperature and velocity
on forced convective boiling heat transfer. The test matrix included low, medium, and
high water temperatures and velocities, as shown in Table 7.5. In all cases, the initial
test temperature was 1000 C . To check the reproducibility of the experiments, one of
the runs (Tf = 50 C , v 4.8 m s ) was repeated.
_1

Three specimens were saved for residual stress measurements and metallographic
characterization, one from each of the three types of steel used in the experiments. The
quenching conditions are shown in Table 7.6.
The effect of surface oxidation was studied by quenching clean as well as oxidized
alloyed samples (runs 20 and 32) and IF specimens (runs 10, 27 and 33). To produce a
'light' surface oxide, a clean specimen was heated to 500 C in air and furnace-cooled; the
surface was protected from further oxidation during heating in the quenching apparatus.
A 'heavy' surface oxide was produced by reusing a quenched specimen without removing
the oxide layer produced by the water quench and heating it up in the quenching appara-
tus without a protecting atmosphere. No attempt was made to characterize the surface
oxide. The quenching conditions are summarized in Table 7.7.

To investigate the effect of initial test temperature, two IF steel bars were quenched
with water flowing at 4.8 m s" at 32 C, from 1000 C (run 19) and 850 C (run 21).
1

6
During heating, the thermocouples were monitored with two digital recorders.
Laboratory Experiments : Quenching Tests 144

7.3 Metallographic Characterization

After residual stress measurements were completed on the specimens quenched under the
test conditions given in Table 7.6, the quenched bars were sectioned near midlength to
produce cylindrical specimens (~ 20 mm tall) for hardness testing and metallography.
After polishing to a 1 pm diamond finish, the samples were etched to reveal the final mi-
crostructure. Both, the 1045 carbon steel and the alloyed eutectoid steel samples where
etched with 2 % nital by rotating a soaked cotton swab on the sample surface for approx-
imately 15 seconds; the IF steel sample required a longer etching time of approximately
40 seconds. The alloyed eutectoid steel sample exhibited a fully martensitic structure
and was, therefore, used to determine the prior-austenite grain size. To measure the
prior-austenite grain size, the alloyed steel cylindrical specimen was further sectioned to
obtain a strip parallel to the radius of the sample. This section was cold mounted, pol-
ished to 1 pm diamond finish, and etched by immersing the mounted sample in a boiling
alkaline picrate solution (2g picric acid, 25 g NaOH and 100 ml water) for approximately
12 minutes, followed by a light 2 % nital etch at room temperature. Several areas were
photographed, and the revealed austenite grain boundaries drawn on tracing paper. The
prior-austenite grain size was determined by measuring the areas of the grains, according
to the Jeffries' method [211]. From the measured mean grain area, a grain diameter was
estimated by assuming a perfectly spherical grain sectioned at its equator. The analysis
was performed using a Leitz O R T H O L U X 2 microscope, a Leitz O R T H O M A T camera
system and a C ' I M A G I N G analysis system running S I M P L E imaging software.
Laboratory Experiments : Quenching Tests 145

Table 7.1: Chemical composition of the IF steel used in the forced convective quenching
experiments (in weight percent).

Element w/o
C 0.01
Mn 0.14
P 0.007
s 0.013
Si 0.01
Cu 0.02
Ni 0.03
Cr 0.03
Mo 0.001
V 0.004
Nb 0.002
Ti 0.068

Table 7.2: Chemical composition of the alloyed steels used in the forced convective
quenching experiments (in weight percent).

Steel
(Heat #)
Element A B C
(E27166) (D25404) (E27071)
C 0.69 0.68 0.65
Mn 0.88 0.83 0.82
P 0.013 0.009 0.010
s 0.020 0.013 0.020
Si 0.23 0.20 0.20
Cu 0.15 0.20 0.16
Ni 0.08 0.08 0.07
Cr 0.19 0.21 0.19
Mo 0.081 0.138 0.169
V 0.001
Nb 0.022 0.020 0.020
Laboratory Experiments : Quenching Tests 146

Table 7.3: Chemical composition of the 1045 steel used in the forced convective quenching
experiments (in weight percent).

Element w/o
C 0.46
Mn 0.78
P 0.007
s 0.021
Si 0.20
Cu 0.02
Ni 0.02
Cr 0.04
Mo 0.001
V 0.002
Nb 0.001

Table 7.4: Typical operational plant data.

Parameter Value
D l 100 mm (4 in)
D 2 200 mm (8 in)
Q 0.1262 m s- (2000 U S G P M )
3 1

V
5.2 m s- 1

T 0 815 C
32 C
V
1.29 x l O - m s- (at 315 C)
7 2 1
Laboratory Experiments : Quenching Tests 147

Table 7.5: Test matrix used for the boiling heat transfer experiments. In all cases the
initial sample test temperature was 1000 C.

Water Temperature, C Velocity, m s 1

2.8 4.8 6.9


25 RUN08 RUN09 R U N 11
50 R U N 12 RUN10 R U N 13
75 RUN16 RUN14 R U N 15

Table 7.6: Quenching conditions for specimens produced for residual stress measurements
and metallographic analysis.

Material Water Temperature Water Velocity Initial Temperature


C m s _1
c

IF 25 4.8 1000
1045 50 2.8 1000
Alloyed 75 2.8 860
Laboratory Experiments : Quenching Tests 148

CD
in
fl
a
u
CD
On o
o o
o O
O o o
o o o LO
oo
LO
oo
o
a
.2
a "fl

"2
'x
o
CD
U
o
o
oo oq oo oo oo
CD

o
CD
CD

s-l
Ti CD
OH
fl O
co o o o o CM
CO
CM
CO
LO LO LO

CD
o
Ti
fl
o fl fl fl
CJ
.2 .2
DJO
fl .2 -+^

CD Ti
^fl CJ fl fl
cj "x
-
x "x
fl O
CD
o __CD O
CD
fl fl O 6
C? rJfl
h0
bO
CD

rjD

o CM
CO
co CM
CM co
CD
id
CD CD
fa fa fa >>
o O
Laboratory Experiments : Quenching Tests
Laboratory Experiments : Quenching Tests

CO

I. "86
Laboratory Experiments : Quenching Tests

0)
CO
38.1 m m
o
CO

o
z
T


LO
CO
CO
Laboratory Experiments : Quenching Tests 153
Laboratory Experiments : Quenching Tests 154

25 10'

Prototype
O D = 25.4 m m
20 A D 31.8 m m
]

D 38.1 m m
]

46 m m 10'
15

0>

10

io l

0 10
0 6 8 10

Q, 1 s

Figure 7.6: Calculated velocity and corresponding Reynolds number as a function of


water flow rate for an annular region with D = 46 mm. The output of the submersible
2

pump used in the experiments is also shown.


Laboratory Experiments : Quenching Tests 155

Figure 7.8: Calculated thermal entry length as a function of Reynolds number for various
Prandtl numbers [209].
Laboratory Experiments : Quenching Tests 156

Vaccum
Pump

Pressure
Gauge

Electromagnetic
Flowmeter
Sample'
Holder Immersion %#2
Heater

Submerged
Pump

Figure 7.9: Flow loop of the quenching apparatus.


Laboratory Experiments : Quenching Tests 157

mui 89 o
co
o
^<
+-<
UIUI 8^7 o

uimoe o

2
Pi

13
CD

p
o
C1

CD

b)
Laboratory Experiments : Quenching Tests 158

110V(a.c.)

Power Surge Protection


a COMPUTER

Ribbon cable

EXP-16/A
7K 7K~ 7K-

4-20 mA
FLOWMETER

THERMOCOUPLE 0-20 mV
(Centre)

THERMOCOUPLE 0-20 mV
(Subsurface)

Figure 7.11: Schematic diagram of the data acquisition configuration.


Laboratory Experiments : Quenching Tests 159

Figure 7.12: Magnetic flowmeter calibration : water flow rate and current as a function
of valve setting, the latter in terms of number of turns with respect to the fully open
position.
Laboratory Experiments : Quenching Tests 160

Figure 7.13: Temperature response at the centre of the specimen during heating to (a)
1000 C and (b) 850 C.
Laboratory Experiments : Quenching Tests 161

10

8 h

6 h

4 h

2 h
0
00 o c\2 CO
o o

Figure 7.14: Difference between centre and subsurface temperature prior to the start of
the quench for the IF steel tests.
Chapter 8

Laboratory Experiments : Transformation Kinetics

To model the microstructural evolution obtained in heat treatments, the transformation


kinetics must be characterized experimentally. In this study, the kinetics of the austen-
ite decomposition in 3 alloyed, near-eutectoid steels for a range of continuous cooling
and isothermal conditions were determined using the G L E E B L E 1500 thermomechanical
simulator. The transformation kinetics of continuously cooled 1045 steel samples were
also measured. The aim of the isothermal and continuous cooling tests was to describe
the transformation in terms of the Avrami equation and to determine the transformation
start times, respectively. The prior-austenite grain size resulting from the heating cycle,
and the ferrite fraction produced by the transformation were determined metallographi-
cally. In this chapter, the details of the material and sample preparation, as well as the
procedure followed for the various tests, including the microstructural characterization,
are described.

8.1 M a t e r i a l and Sample Preparation

Alta Steel provided the alloyed near-eutectoid steel. The material was received as an-
nealed 100 mm (4 in)-dia. rods. The 1045 steel samples were machined from commercially
available 38.1 mm (1.5 in)-dia. cold-rolled bars. The transformation studies were carried
out using thin-wall tubular samples (6 mm I.D., 8 mm O.D.), shown in Figure 8.1.
To assure a homogeneous composition, free of any centreline segregation common to

162
Laboratory Experiments : Transformation Kinetics 163

rod material, rough sample blanks were cut from the 100 mm-dia. rods as shown in
Figure 8.2. The tubular cylindrical test samples were machined from these blanks and
homogenized at 900 C for 1 hour. To minimize oxidation while being homogenized, the
samples were placed in a senpak bag, which was then back-filled with argon.

8.2 Procedure

8.2.1 Continuous Cooling Tests

The progress of the austenite decomposition was monitored by measuring the central
plane diametral dilation of the cylindrical specimen with a modified Linear Variable
Differential Transformer (LVDT) C-strain gauge. The sample temperature was controlled
and monitored using a chromel-alumel thermocouple spot-welded on the exterior surface
at the same mid-axis location. Typical sampling frequencies ranged from 2.5 to 59.6 Hz
for cooling rates varying from 0.2 to 40 C s . Prior to testing, the main test chamber
_1

of the Gleeble 1500 was pumped down to 5 x l O - 4


torr and back-filled with prepurified
argon (99.998 % pure).
A total of 10 tests were conducted for each of the alloyed eutectoid steels. The test
conditions are schematically shown in Figure 8.3 and summarized in Tables 8.1 to 8.3.
In all cases, the samples were initially heated at 5 C s _1
to a holding temperature of
850 C , held for 3 minutes and continuously cooled at the test cooling rate.

Six tests were done for the 1045 carbon steel samples, following the same procedure
previously described. The test conditions are summarized in Table 8.4.

8.2.2 Isothermal Tests

The progress of the austenite decomposition under isothermal conditions was also moni-
tored by measuring the central plane diametral dilation of the cylindrical test specimen
Laboratory Experiments : Transformation Kinetics 164

as described above. Typical sampling frequencies ranged from 7 to 30 Hz for test tem-
peratures varying from 665 to 525 C.
The isothermal test conditions are schematically shown in Figure 8.4. In all cases, the
samples were heated at 5 C s _1
to a holding temperature of 850 C, held for 3 minutes,
cooled at 8 C s _1
to 750 C , rapidly cooled (at ~ 140 C s ) to the test temperature
_1

and held until completion of the transformation. The two-step cooling path was found to
be necessary to minimize errors associated with overshooting the test temperature upon
rapid cooling. A series of isothermal tests was also conducted to measure the isothermal
transformation kinetics of the austenite-to-bainite transformation in the alloyed eutectoid
steels. In this case, the samples were austenitized as described above and rapidly cooled
(He quenched) to the desired test temperature. No temperature overshoot was observed.
The test conditions for all cases are summarized in Tables 8.5 to 8.8.

8.2.3 T e m p e r a t u r e Gradient in a Gleeble Specimen

Ideally, the mid-plane (C-strain) diametral dilatometric measurements should be made


with little or no temperature gradient across the wall of the cylindrical test specimens. A
finite-element simulation of the rapid cooling of a tubular Gleeble specimen was performed
to determine if a significant temperature gradient exists across the thickness of the tubular
wall. The boundary condition at the inner surface was estimated by solving the inverse
heat conduction problem (IHCP).

From the measured temperature response obtained during test BCC05-2 ( He quench
with a cooling rate of 38 C s _ 1
), a set of temperatures was established at constant time
intervals (0.1 s), and adopted as input to the inverse heat conduction code described in
Chapter 5 (modified for a hollow cylinder). To avoid recalescence due to the austenite-
to-bainite transformation, experimental temperatures below 520 C were not included.
Considering that the heat transferred to the helium flowing inside the sample is much
higher than the heat loss due to radiation plus natural convection at the external surface,
Laboratory Experiments : Transformation Kinetics 165

the external boundary was assumed to be insulated (to simplify the solution of the IHCP).
For the calculations, the sample was subdivided into 15 nodal points using a time interval,
At, of 0.1 s; the number of future time steps adopted in the inverse analysis was r = 2.
The estimated heat-transfer boundary condition at the inner surface was adopted for
the finite-element simulation of the direct problem. For the finite-element calculations,
the cross-section of the sample wall was discretized using 10 8-node isoparametric ele-
ments (the ratio Ar/Az for a given element was kept close to 1.0), and a time interval,
A t , of 0.05 s was selected. The calculated results showed a maximum temperature dif-
ference of approximately 0.5 % across the specimen thickness, at mid-axis, when the
temperature was 700 C.
Because of the approximation of no heat flux at the outside boundary adopted for
the I H C P calculations, the actual surface heat flux at the interior boundary is expected
to be lower than its estimated value, thus reducing the estimated temperature difference
across the specimen wall even further. Also, the helium quench was a rapid cooling
condition; lower cooling rates would result in even more uniform temperature profiles in
the specimen. Thus, it was concluded that a relatively small temperature gradient within
the cylinder wall exists during the transformation kinetic measurements, resulting in an
essentially isothermal plane at the specimen mid-axis.

8.3 M i c r o s t r u c t u r a l Characterization

After each test was completed, the tubular specimen was sectioned near the thermocouple
position, cold mounted, polished to a 1 pm diamond finish and etched to reveal the
microstructural detail.

It was important to determine the prior-austenite grain size produced by the heating
cycle during measurements done with the Gleeble 1500 since it affects the transforma-
tion kinetics. To reveal the prior-austenite grain boundaries, a fully martensitic specimen
Laboratory Experiments : Transformation Kinetics 166

(alloyed eutectoid steel cooled at 58 C s ) was sectioned, cold mounted, polished to 1


_1

pm diamond, and etched by immersing the mounted sample in a boiling alkaline sodium
picrate solution (2g picric acid, 25 g NaOH and 100 ml water) for 12 min followed by
a light 2 % nital etch at room temperature. The prior-austenite grain size was deter-
mined by measuring the mean chord length of the grains. The image analysis system
consisted of a Leitz O R T H O L U X 2 microscope, a Leitz O R T H O M A T camera system and
a C ' I M A G I N G analysis system running S I M P L E imaging software.

The ferrite fraction produced by all the continuous cooling tests done using the 1045
carbon steel was measured metallographically. The ferrite fraction present in alloyed
eutectoid steel samples, continuously cooled at the slowest cooling rates, was also deter-
mined. To measure the ferrite fraction, the samples were etched with a mixture of 15 ml
2 % nital (2 % nitric acid in alcohol) and 85 ml 5 % picral (5 % picric acid in alcohol) [188];
the etchant was applied by gently swabing the sample surface for approximately 10 to 15
seconds. The ferrite fraction was determined by measuring the area occupied by ferrite
in up to 50 separate fields for each specimen, using the image analyzer described above.
Laboratory Experiments : Transformation Kinetics 167

Table 8.1: Summary of continuous cooling tests for the alloyed eutectoid steel A .

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s C s
ACC01 5 850 180 Air cooling
( C R . = 17.5 C s- ) 1

ACC02 5 850 180 1 C s- 1

ACC03 5 850 180 4 C s- 1

ACC04 5 850 180 0.25 C s" 1

ACC05 5 850 180 He quench, 1/2 turn


( C R . = 37 C s- ) 1

ACC06 5 850 180 He quench, 2 turns


( C R . = 58 C s- ) 1

ACC07 5 850 180 2.5 C s" 1

ACC09 5 850 180 8 C s- 1

ACC10 5 850 180 0.5 C s- 1

ACC11 5 850 180 He quench, 1 turn


( C R . = 44 C s- ) 1

Table 8.2: Summary of continuous cooling tests for the alloyed eutectoid steel B .

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s C s
BCC01 5 850 180 Air cooling
( C R . = 16 C s- )
1

BCC02 5 850 180 1 C s- 1

BCC03 5 850 180 4 C s- 1

BCC04 5 850 180 0.2 C s" 1

BCC05 5 850 180 He quench, 1/2 turn


( C R . = 38 C s" )
1

BCC06 5 850 180 He quench, 0.15 turn


( C R . = 28 C s" )
1

BCC07 5 850 180 2.5 C s" 1

BCC08 5 850 180 0.3 C s- 1

BCC09 5 850 180 8 C s- 1

BCC10 5 850 180 0.5 C s- 1


Laboratory Experiments : Transformation Kinetics 168

Table 8.3: Summary of continuous cooling tests for the alloyed eutectoid steel C.

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s c s
CCC01 5 850 180 Air cooling
( C R . = 17 C s- )1

CCC02 5 850 180 1 C s- 1

CCC03 5 850 180 4 C s- 1

CCC04 5 850 180 0.2 C s- 1

CCC05 5 850 180 He quench, 1/2 turn


( C R . = 40 C s" )
1

CCC06 5 850 180 He quench, 1/2 turn


( C R . = 33 C s-
1

CCC07 5 850 180 2.5 C s- 1

CCC08 5 850 180 He quench, 1/2 turn


( C R . = 28 C s" )1

CCC09 5 850 180 8 C s" 1

CCC10 5 850 180 0.5 C s- 1

Table 8.4: Summary of continuous cooling tests for the 1045 carbon steel.

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s C s
CC4501 5 850 180 1 C s" 1

CC4502 5 850 180 Air cooling


( C R . = 19 C s- ) 1

CC4503 5 850 180 He quench, 1/2 turn


( C R . = 38 C s" ) 1

CC4504 5 850 180 He quench, 1 turn


( C R . = 47 C s- ) 1

CC4505 5 850 180 He quench, 5 turns


( C R . = 104 C s" ) 1

CC4506 5 850 180 He quench, 2 1/2 turn


( C R . = 250 C s" 1
Laboratory Experiments : Transformation Kinetics 169

Table 8.5: Summary of isothermal tests for the alloyed eutectoid steel A .

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s c s
AIT01 5 850 180 Cool at 8 C s" 1
to 750 C
followed by helium quench to
625 C and hold
AIT02 5 850 180 Cool at 8 C s" 1
to 750 C
followed by helium quench to
605 C and hold
AIT03 5 850 180 Cool at 8 C s _1
to 750 C
followed by helium quench to
635 C and hold
AIT04 5 850 180 Cool at 8 C s~ l
to 750 C
followed by helium quench to
650 C and hold
AIT05 5 850 180 Cool at 8 C s _1
to 750 C
followed by helium quench to
665 C and hold
AIT06 5 850 180 Cool at 8 C s _ i
to 750 C
followed by helium quench to
575 C and hold
AIT07 5 850 180 Cool at 8 C s _1
to 750 C
followed by helium quench to
525 C and hold
Laboratory Experiments : Transformation Kinetics 170

Table 8.6: Summary of isothermal tests for the alloyed eutectoid steel A (bainite reac-
tion).

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s c s
BAI01 5 850 180 Helium quench to 350 C and
hold
BAI02 5 850 180 Helium quench to 375 C and
hold
BAI03 5 850 180 Helium quench to 400 C and
hold
BAI04 5 850 180 Helium quench to 420 C and
hold
BAI05 5 850 180 Helium quench to 440 C arid
hold
BAI06 5 850 180 Helium quench to 460 C and
hold
BAI07 5 850 180 Helium quench to 480 C and
hold
Laboratory Experiments : Transformation Kinetics 1

Table 8.7: Summary of isothermal tests for the alloyed eutectoid steel B .

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s c s
BIT01 5 850 180 Cool at 8 C s" 1
to 750 C
followed by helium quench to
630 C and hold
BIT02 5 850 180 Cool at 8 C s" 1
to 750 C
followed by helium quench to
600 C and hold
BIT03 5 850 180 Cool at 8 C s _1
to 750 C
followed by helium quench to
615 C and hold
BIT04 5 850 180 Cool at 8 C s- 1
to 750 C
followed by helium quench to
650 C and hold
Laboratory Experiments : Transformation Kinetics 1

Table 8.8: Summary of isothermal tests for the alloyed eutectoid steel C.

HOLD
Test Heating Rate Temperature Time Cooling Condition
C/s C s
CIT01 5 850 180 Cool at 8 C s" to 750 C
1

followed by helium quench to


620 C and hold
CIT02 5 850 180 Cool at 8 C s to 750 C
_1

followed by helium quench to


600 C and hold
CIT03 5 850 180 Cool at 8 C s to 750 C
_1

followed by helium quench to


635 C and hold
CIT04 5 850 180 Cool at 8 C s to 750 C
_1

followed by helium quench to


650 C and hold
CIT05 5 850 180 Cool at 8 C s- to 750 C
1

followed by helium quench to


665 C and hold
CIT06 5 850 180 Cool at 8 C s- to 750 C fol-
1

lowed by helium quench 550 C


and hold
CIT07 5 850 180 Cool at 8 C s- to 750 C
1

followed by helium quench to


500 C and hold
CIT08 5 850 180 Cool at 8 C s" to 750 C
1

followed by helium quench to


680 C and hold
Laboratory Experiments : Transformation Kinetics 173

-Hi h- 3 -H

All dimensions in mm

Figure 8.1: Geometry of the Gleeble specimen used for characterizing the phase trans-
formation kinetics.

Rough Sample Blanks

11 mm

Note : Not to scale

Figure 8.2: Location of the sample blanks taken from the 100 mm-dia. rods.
Laboratory Experiments : Transformation Kinetics

3 min

_J_

200 300 400 500

Time, s

Figure 8 . 3 : Schematic representation of continuous cooling tests

200 300 500

Time, s

Figure 8 . 4 : Schematic representation of isothermal tests.


Chapter 9

Laboratory Results and Discussion : Quenching Tests

The results of the quenching tests are presented in this chapter. The measured temper-
ature response in the quenched specimens is first discussed, followed by the estimated
surface heat flux (from the inverse analysis). Finally, the results of the microstructural
characterization of the quenched samples are presented.

9.1 Measured Temperature Response

The data acquired during the forced convective quenching experiments consisted of the
temperature response at two positions in the steel bars. Centre and subsurface temper-
ature responses were measured during quenching of IF steel specimens, while centre and
mid-radius positions were selected for the 1045 carbon and the alloyed eutectoid steels.
The quenching conditions were given in Tables 7.5 to 7.7. To assess the reproducibility
of the results, two IF steel bars were quenched in separate experiments (runs 10 and 18)
with water flowing at 4.8 m s _1
at Tj = 50 C. The temperature response at the sub-
surface and centre locations is shown in Figure 9.1. As can be seen, good reproducibility
was observed.

Visual inspection of the measured cooling curves was used to compare the cooling
power of the bath for the various quenching conditions investigated [11]. The mea-
sured temperature response at the centre, during forced convective quenching of the 38.1
mm-dia. IF steel bars in water flowing at 2.8, 4.8 and 6.9 m s _1
for 3 values of water
temperature is shown in Figures 9.2 to 9.4. The comparisons are presented based on

175
Laboratory Results and Discussion : Quenching Tests 176

temperature responses at the centre because the effects of water velocity and water tem-
perature are more evident on these curves than on the subsurface temperature responses.
When a water temperature of 75 C was used, the cooling curves showed recalescence
near the austenite-to-ferrite equilibrium temperature (910 C ) . To confirm this effect, the
temperature response during the quench was simulated numerically (see Appendix E ) .

For a given water velocity, the specimen cools faster as the water temperature de-
creases (subcooling increases); however, the temperature response for Tj 75 C is
markedly different than that for 25 and 50 C, in that a much lower rate of change of
temperature with time was observed at the early stages of the quench. This result sug-
gests that there is a threshold value that separates two distinctive types of behaviour. A
similar observation was made when a stainless steel disk was quenched from 850 C in
still water [194]; in that investigation, curves of heat-transfer coefficient vs surface tem-
perature showed different behaviour when the bath temperature was below 60 C. Note
that the difference between the curves corresponding to Tj = 25 and 50 C decreases
as the water velocity increases. On the other hand, when the results were grouped by
water temperature (Figures 9.5 to 9.7), the curves for the three water velocities showed a
remarkable similarity. As the water velocity increases, the specimen cools faster, due to
an improved liquid-solid contact. As can be seen in the figures, the temperature response
corresponding to water flowing at 4.8 m s _1
approaches that for v = 6.9 m s _1
for all
levels of water temperature investigated. This observation suggests that there is a limit
to the rate at which the specimen can be cooled, dictated by the internal resistance of
the bar.

Thermal stresses are generated in quenched parts by temperature gradients generated


during cooling. It is, therefore, important to investigate the effect of water temperature
and velocity on the measured thermal gradients as the quench progresses. The tem-
perature difference between the centre and the subsurface temperatures as a function
of subsurface temperature during forced convective quenching of IF steel specimens in
Laboratory Results and Discussion : Quenching Tests 177

water flowing at 25, 50 and 75 C for 3 water velocities is plotted in Figures 9.8 to 9.10.
Starting with an essentially uniform temperature of ~ 1000 C , the thermal gradient
across the specimen radius increases as the subsurface temperature decreases, reaching
a maximum, and then decreasing towards zero as the specimen temperature approaches
the water temperature. This behaviour is the result of a combination of two factors :
1) the dependence of the surface heat flux on surface temperature, which shows a maxi-
mum at the critical heat flux, and 2) the internal resistance to conduction heat transfer,
which causes a damping and lagging effect in the temperature response at the centre. In
all cases, the maximum temperature difference occurs at or just above 200 C; this is
an important finding, given that the martensitic transformation in the alloyed eutectoid
steel occurs at ~ 220 C. The results for Tf = 25 C are essentially identical regardless
of water velocity, while smaller thermal gradients were observed as the water velocity
decreased for Tf = 50 and 75 C , i.e., 'softer' quenching conditions.

The local cooling rate is an important parameter in determining the microstructural


evolution in a quenched part. In this work, the cooling rate was estimated by computing
the rate of change of temperature with time, using the commercial package T - S M O O T H
[212]. The resulting cooling rate at the centreline as a function of local temperature for
tests done with water flowing at 25, 50 and 75 C, for 3 water velocities is shown in
Figures 9.11 to 9.13. As the temperature decreases, the cooling rate increases, until a
maximum value is reached, and then decreases to zero. For the highest water temperature
investigated (Tf = 75 C , Figure 9.13), the cooling rate was nearly constant between 1000
and ~ 800 C before a sudden increase was observed. The maximum cooling rate occurred
at local temperatures temperatures between 500 and 700 C. The maximum cooling rate
as a function of water velocity, for the 3 levels of water temperature studied, is shown
in Figure 9.14. For a given water temperature, the magnitude of the maximum cooling
rate increases as the water velocity increases. The largest cooling rates were obtained for
the lowest water temperature (highest sub cooling). Similar calculations were performed
Laboratory Results and Discussion : Quenching Tests 178

using the data at the subsurface position. The measured maximum cooling rates as a
function of water velocity are shown in Figure 9.15 for 3 values of water temperature.
As expected, the maximum cooling rates are higher than the corresponding values at the
centreline. Also, the effect of water velocity on the maximum water flow rate was small
at the highest water temperature, while a significant influence was observed for Tj = 25
and 50 C .

In the actual plant operation, scale formation occurs during reheating and transporta-
tion of the bars before the quench commences. To study the effect of surface condition on
heat extraction, IF and alloyed eutectoid steel specimens were intentionally oxidized prior
to quenching. The measured temperature response at the centre during forced convec-
tive quenching of clean, 'lightly' oxidized and 'heavily' oxidized IF steel bars with water
flowing at 4.8 m s _ 1
at 50 C is shown in Figure 9.16. The presence of an oxide layer
increased the rate of heat extraction, but the effect was minimal. It should be pointed
out that no attempt was made to characterize the oxide layer and, therefore, references
to this layer are only qualitative. The temperature response at the centre, during forced
convective quenching of clean and 'heavily' oxidized alloyed steel bars in water flowing
at 4.8 m s _ 1
at 32 C is shown in Figure 9.17. Little difference in cooling behaviour
was observed between the two surface conditions. Differences in heat extraction have
been reported when clean and oxidized specimens were quenched in pool boiling condi-
tions [194,199]. A better adherence of the vapour film when an oxidized layer was present
was cited as the reason for the observed difference [199]. Data on the effect of an oxide
layer on heat extraction during forced convective quenching is scarce. The results of
the present investigation suggest that the enhanced water-solid contact provided by the
flowing water overcomes any influence that an oxide layer may have on heat extraction.

The effect of initial test temperature was also studied. Two IF steel bars were
quenched from 1000 and 850 C with water flowing at 4.8 m s _1
at 32 C; the tem-
perature responses at the centre are shown in Figure 9.18. Except for a short initial
Laboratory Results and Discussion : Quenching Tests 179

period, the two curves showed essentially the same behaviour. A similar result was ob-
served in boiling curves obtained during forced convective quenching of small-diameter
(0.3 or 0.5 mm) platinum wires, for a wide range of initial temperatures [213].

9.2 Estimated Surface Heat Flux

Based on the measured temperature response at the subsurface of IF steel bars, the
surface heat flux as a function of surface temperature was estimated by solving the
inverse heat conduction problem, as described in Chapter 5. Some of the data recorded
during the experiments showed fluctuations that would be amplified by the solution of
the IHCP. Therefore, the data were filtered before the inverse analysis was applied. A
11-point Savitzky-Golay filter of order 2 [214] was adopted to remove high frequency
fluctuations in the measured temperature responses. The filtering algorithm is based
on a polynomial least-square fitting inside a moving window; in general, it provides a
better smoothing than a simple moving average [214]. The raw and filtered temperature
responses during forced convective quenching of an IF steel bar in water flowing at 2.8
m s _ 1
at 25 C are shown in Figure 9.19. It should be noted that, for clarity, relatively
few raw data points were plotted. The inserts in the figure show magnified views of two
areas where sharp bends in the cooling curve were present. As can be seen, the filtering
procedure removes fluctuations in the data without losing the main characteristics of the
cooling curve.

As mentioned before, some cooling curves showed recalescence due to the heat gen-
erated during the austenite-to-ferrite transformation in the IF steel. The code adopted
to solve the IHCP was not designed to handle heat sources/sinks. Thus, the following
strategy was implemented : the temperature response in the range of the transformation
was not included in the inverse analysis; the corresponding surface heat flux was then
estimated by extrapolating the values obtained after the transformation was complete.
Laboratory Results and Discussion : Quenching Tests 180

Given that the transformation occurred at relatively high temperatures (see Appendix
E), where film boiling was present, and the recalescence effect was observed only when
water at 75 C was used, this treatment was considered to be adequate. A similar proce-
dure has been adopted by Fernandes et al. [127] to estimate the surface heat flux in the
austenite-to-pearlite transformation range during the air cooling of 4 mm-dia. carbon
eutectoid steel rods.

For the calculations, the specimens were subdivided into two regions of 1.5 and 17.55
mm, discretized by 5 and 15 nodes, respectively; this mesh was adopted based on the
subsurface thermocouple position (1.5 mm beneath the specimen surface), and the need
for a finer mesh near the surface. The actual thermocouple position was measured in few
specimens after the tests were completed, by sectioning the sample; it was found that the
thermocouple was located within 1 mm from its nominal position. The At adopted
for estimating the surface heat flux and surface temperature was 0.2 s (the same as the
experimental time step). The computer code allows the time step used in the solution
to be smaller than the experimental time step; however, no significant differences in
the estimated surface heat flux were found when computational time steps of 0.1 s and
0.05 s were adopted. In the latter case, numerical instabilities were introduced in some
runs, which then required the use of a larger number of future time steps to stabilize
the solution. A n attempt was made to simulate data recorded at higher frequency, by
interpolating values from a given temperature response; the new data was then input to
the inverse analysis code. No significant differences were observed between the results
obtained with the original data (measured every 0.2 s) and the simulated data ('measured'
every 0.1 s). The thermophysical properties of IF steel were input as explicit functions
of temperature by obtaining the best-fit curve of data reported in the literature for pure
iron and very low carbon steel [215,216] :

k = 81.86 -0.0974 T + 4.3 x 1 0 - 5


T2
(9.1)
Laboratory Results and Discussion : Quenching Tests 181

p = 7876 - 0.331 T (9.2)

Cp = a(b - T) exp[d(b - T)],


c
T < 770 C

a = 1756.8, b = 771.9,

c= -0.1506, d=-0.00047 (9.3)

Cp = a(T - b) exp[d(T - b)}


c
T > 770 C

a = 2164.6, 6 = 765.4,

c = -0.2964, d = 0.00144 (9.4)

where T is in C and k, p and C are in SI units.


p

The application of the inverse analysis is illustrated by considering the case of forced
convective quenching of an IF steel bar with water flowing at 2.8 m s _1
at 25 C. The
experimentally determined temperature response at the centre and subsurface is shown
in Figure 9.20. The corresponding estimated surface heat flux as a function of surface
temperature is plotted in Figure 9.21. The raw data was filtered using a Savitzky-Golay
filter and a value of r = 2 was adopted for the calculations. When r was set to 1 (exact
matching of the data), very large oscillations in the estimated heat flux were observed.
For comparison, the estimated surface heat flux as a function of surface temperature,
obtained with raw and filtered data, is shown in Figure 9.22. B y filtering the data, a
smoother boiling curve was obtained, while preserving the general characteristics of the
curve.

The number of future time steps adopted for the calculation of the heat flux in the
inverse solution is given by the parameter r. To investigate the effect of future time
steps and prefiltering, runs with r = 2, 4, and 6, using raw as well as filtered data
were conducted. The results are presented in Figures 9.23 and 9.24 for raw and filtered
Laboratory Results and Discussion : Quenching Tests 182

input data, respectively. In the figures, the residuals (difference between measured and
estimated subsurface temperature) are plotted as a function of measured temperature,
for 3 values of the parameter r. For clarity, the right axis was used to plot the curves
corresponding to r = 2. The scales on the left and right axis are identical, except that
one of them is shifted with respect to the origin. As can be seen, the magnitude of the
residuals increased as the value of r increased. This trend of poorer agreement between
the measured and estimated temperature responses for higher values of r, arises from the
use of additional future time steps, which tends to 'smooth' the solution of the IHCP. A
negative effect of adopting larger values of r, is that sudden changes in the heat flux, such
as peaks or valleys, are smoothed. On the other hand, larger values of r may need to be
used to stabilize the solution when particularly small time steps are adopted or very noisy
data is used [183]. There are no specific rules for choosing r, but general guidelines have
been reported [183,217]. For a plate of thickness L, heated at one boundary and insulated
at the other, values of r = 2, 3, 4, 5 and 6 were recommended for values of a/S.t j I? =
0.12, 0.07, 0.052, 0.042 and 0.036, respectively, for the case where the thermocouple was
located at the insulated boundary. In general, the closer the thermocouple is located with
respect to the active heat transfer boundary, the smaller is the required value of r [217].
By computing the thermophysical properties of IF steel at an intermediate temperature
of 500 C , and choosing L = 1.5 mm (the distance between the thermocouple and the
specimen surface), a value of aAt/L 2
= 0.76 was obtained. Low values of r can then
be adopted in the calculations; this is a direct result of having filtered the raw data,
which is equivalent to using data with small random errors. At all levels of r, prefiltering
of the data reduces the magnitude of the residuals and removes unwanted responses to
fluctuations in the experimental data (compare the residuals between 50 and 250 C for
raw and filtered data in Figures 9.23 and 9.24). The running sum of the root mean square
(RMS) of the temperatures is a good estimate of the matching between measured and
calculated temperatures. The sum of the R M S of the temperatures at the subsurface
Laboratory Results and Discussion : Quenching Tests 183

position was 1.7, 4.4 and 8.1 C for r = 2, 4, and 6, respectively when using raw data;
the corresponding values for filtered data were 0.8, 3.8 and 7.6 C. Clearly, a better
agreement between measured and estimated temperatures, especially at low values of r,
was obtained by prefiltering the measured data.
The critical heat, flux (CHF) is an important parameter in heat transfer processes
where boiling is present. Raw and filtered temperature responses from run 8 {y 2.8
m/s, Tf = 25 C) were used to assess the effect of filtering the data previous to the inverse
analysis on the estimated C H F . The effect of pre-filtering on the estimated magnitude
and temperature of occurrence of the C H F is summarized in Table 9.1. When raw data
was used, the estimated values of C H F were 5.3, 5.0, and 4.7 M W m - 2
for r = 2, 4,
and 6, respectively; the corresponding estimated surface temperatures were 445.4, 410.8
and 373.9 C . In contrast, the filtered data resulted in C H F values of 5.3, 5.0 and
4.8 M W m ~ at 411.1, 412.9, and 414.2 C for r = 2, 4, and 6, respectively. As can
2

be appreciated, the uncertainty associated with the surface temperature at which C H F


occurs is greatly reduced by pre-filtering the data, while the magnitude of the C H F is
only slightly increased (for r = 5). It can then be concluded that pre-filtering of the
measured data results in a better characterization of the critical heat flux.
Boiling curves estimated from subsurface temperature measurements recorded during
forced convective quenching of IF steel bars with water flowing at 2.8, 4.8 and 6.9 m s _1

for 3 values of water temperature are shown in Figures 9.25 to 9.27. For a given water
velocity, the surface heat flux is higher as the subcooling increases (water temperature
decreases). A t low subcoolings, a film boiling stage was evident in the boiling curves,
whereas at subcoolings of 50 and 75 C , no film boiling stage could be identified; the
boiling curves immediately reached the transition boiling stage. The boiling curves ob-
tained with subcoolings of 50 and 75 C were very similar for the intermediate (4.8 m s ) _1

and high (6.9 m s ) water velocities investigated, i.e.,.the amount of heat extracted ap-
_1

proached a limit, defined by the internal thermal resistance of the bar. In some of the
Laboratory Results and Discussion : Quenching Tests 184

boiling curves in Figures 9.25 to 9.27, a 'shoulder' can be observed. This observation has
also been reported by Ishigai et al. [73] in a study of boiling heat transfer of a water jet
impinging on a hot surface. As was the case in the present study, those researchers found
that no shoulder appeared in the transition region when the subcooling was low. From
visual inspection, they concluded that this sudden decrease of heat flux in the transition
boiling region, can be attributed to frequent and instantaneous liquid-solid contacts and
that it disapears when the solid surface is wet.

The effect of film boiling on the boiling curve can be shown by plotting the surface
heat flux as a function of time [40]. The estimated surface heat flux during forced
convective quenching of IF steel bars with water flowing at 2.8 m s _1
for the 3 water
temperatures investigated is plotted in Figure 9.28. The presence of film boiling in the
run corresponding to Tj = 75 C has displaced the start of the transition boiling stage
considerably.
Very little experimental data on forced convective boiling has been published and,
consequently, few comparisons with the results of the present study can be made. Honda
et al. [213, 218] have studied the heat transfer during forced convective boiling of thin
platinum wires (0.3 and 0.5 mm in diameter), using water and CaC^/water solutions
as the quenchant. Their results, in the form of surface heat flux as a function of wall
superheat ( A T s a t = T T ) , are shown in Figure 9.29.
s s a t When quenching in pure
water, they identified two local minimum-heat-flux points (shown as Ml and M 2 in the
figure), which correspond to the two slope changes they observed in the cooling curve.
The first minimum point was associated with the collapse of the vapour film [218]. In
contrast, when solutions of CaCl2 in water were used as a quenchant, the Ml point was
not observed in the experiments with wire of 0.5 mm-dia. They attributed these results
to an enhanced wetting of the wire surface due to deposition of CaCl2. The results
presented in Figures 9.25 to 9.27 are replotted in Figures 9.30 to 9.32 for comparison
with the boiling curves of Honda et al. [213,218]. A major difference between the results
Laboratory Results and Discussion : Quenching Tests 185

of the present study and those shown in Figure 9.29 is the absence of the point Ml of
their experiments with pure water. Instead, the boiling curves in Figures 9.30 to 9.32
resemble their results for quenching with CaCl2 solutions. The heat fluxes obtained in
this investigation are significantly lower than those reported by Honda et al. [213, 218]
due to the much larger diameter used (38.1 mm vs 0.3 - 0.5 mm).

As pointed out in Chapter 2, there are very few reported data for heat transfer
during forced convective quenching for the conditions of interest. The film boiling heat
flux as a function of surface temperature was computed using an integral method [219],
based on the solution given by Nakayama and Koyama [220]. Given that the ratio of
thickness of the boundary layer to bar radius is small, the assumption was made that
the problem could be treated as forced convective boiling (parallel flow) over a flat plate.
The results of calculations for an IF steel bar quenched with water flowing at 75 C at 3
water velocities are shown in Figure 9.33, along with the measured boiling curves. The
computed forced convective film boiling curves underestimate the magnitude of the heat
flux during the film boiling stage of the measured boiling curves. Both, the theoretical
and measured curves showed an increase in heat transfer as the water velocity increased.
Similar findings have been reported by Honda et al. [213]; they performed a numerical
analysis of film boiling on a horizontal cylinder with upward flow, and found that the
computed heat fluxes underestimated the measured values by 30 %.

The estimated critical heat flux as a function of water velocity, during forced con-
vective quenching of IF steel bars for 3 water temperatures is shown in Figure 9.34 and
summarized in Table 9.2. The critical heat flux increases, almost linearly, as the wa-
ter velocity increases; it also increases as the water temperature decreases (subcooling
increases).

The heat extraction obtained under the quench conditions investigated can also be
characterized by the average heat flux (q ), which is proportional to the area under the
Laboratory Results and Discussion : Quenching Tests 186

boiling curve :

avg
(9.5)

The average heat flux as a function of water velocity for 3 water temperatures is shown
in Figure 9.35 and summarized in Table 9.3. The average heat flux increases as the water
velocity increases and the water temperature decreases.
Once the surface heat flux and surface temperature have been estimated, it is possible
to compute the heat-transfer coefficient. Figure 9.36 shows the heat-transfer coefficient
as a function of surface temperature, during forced convective quenching of IF steel
bars with water flowing at 25 C for 3 water velocities. It should be pointed out that
the heat-transfer coefficients were computed with respect to the boiling point of water
(see Eq. (2.1)). In all cases, the heat-transfer coefficient increases monotonically as
the surface temperature decreases. The heat-transfer coefficient increases, at all levels
of surface temperature, as the water velocity increases. Even though these curves can
be easily correlated to the surface temperature, they all tend to infinity as the surface
temperature approaches the boiling point of water. Large errors in the calculated heat-
transfer coefficient can then be generated at low surface temperatures and, therefore, it
is preferable to describe the heat transfer at the active boundary in terms of surface heat
fluxes.

9.3 Metallographic Characterization

Three specimens, corresponding to each of the materials used during the quenching ex-
periments, were saved for metallographic characterization, hardness and residual stress
measurements. The quenching conditions were given in Table 7.6.

The microstructures at the centre and ~ 3 mm from the surface produced by forced
convective quenching of IF and alloyed steel bars, under the conditions given in Table 7.6,
Laboratory Results and Discussion : Quenching Tests 187

are shown in Figures 9.37 and 9.38, respectively. In both steels, the microstructure at
the centre and near the surface were very similar. The microstructure in the quenched IF
steel bar (Figure 9.37) consisted of essentially 100 % ferrite (see Figure 9.37), whereas the
alloyed eutectoid steel showed a predominantly martensitic microstructure in Figure 9.38.

A low magnification (8X) photograph of the etched macrostructure of the 1045 carbon
steel sample (Figure 9.39) clearly showed a ring of martensite (white) at the surface.
The microstructure in this steel in the martensitic band, in the transition zone, and at
the centre of the sample are shown in Figure 9.40 (a), (b) and (c), respectively. Near
the surface, the final microstructure was essentially martensitic (white) with areas of
pearlite/bainite (dark) delineating the prior-austenite grain boundaries. In the transition
zone (Figure 9.40 (b)), the amount of martensite decreases, the white areas in Figure 9.40
(a) being replaced with pearlite and bainite, the pearlite colonies outlining the original
austenite phase boundaries; at the centre (Figure 9.40 (c)), the microstructure consisted
of large patches of martensite (white) with narrow white ferrite outlining the original
austenite grains, the remaining dark equiaxed phase being pearlite.
The hardness profile across the bar diameter of the 3 steels (IF, alloyed eutectoid and
1045) is shown in Figure 9.41. The hardness of the alloyed eutectoid and the 1045 carbon
steel specimens was measured in Rockwell C units (left axis) while Vickers units (right
1

axis) were used for the IF sample. To avoid errors near the surface of the sample, hardness
measurements extended only to 17.5 mm from the center of the bar. To complete the
hardness profile, microhardness measurements were made near the surface. The alloyed
eutectoid steel specimen produced the highest level of hardness, followed by the 1045
carbon steel and the IF steel. As expected from the metallographic analysis, the hardness
across the alloyed eutectoid and the IF steel specimen was very uniform. In contrast,
the 1045 carbon steel exhibited a hardness gradient across the radius of the bar. Near
the surface, the hardness was uniform for the first 2.5 mm, and of the order of the value

1
Using a 10 kg load.
f
Laboratory Results and Discussion : Quenching Tests 188

exhibited by 100% martensite in a 1045 carbon steel [221]; it then decreased sharply (~
25 HRc units) between 2.5 and 5 mm, and remained relatively constant in the core. This
hardness profile agrees well with the microstructures shown in Figures 9.40 (a) to 9.40
(c). For comparison, a typical measured hardness profile in a 100 mm-dia., 6 m long
alloyed eutectoid steel rod, after quenching and tempering under industrial conditions,
is shown in Figure 9.42. In the figure, the hardness across the radius near both ends and
at mid-length, as well as their average, is shown.

The fully martensitic alloyed eutectoid quenched bar was used to determine the prior-
austenite grain size produced by the heating cycle. To facilitate image analysis, the
grain boundaries revealed by the boiling alkaline sodium picrate etch were drawn on
tracing paper; a typical example is shown in Figure 9.43. The prior-austenite grain size
was determined by measuring the areas of the individual grains according to the Jeffries
method [211]. Several frames were analysed and a total of 488 grain areas were measured . 2

The number of grain areas measured was considered to be statistically significant. The
measured grain area distribution is shown in Figure 9.44. The equivalent area grain
diameter (EQAD) was calculated assuming the areas in the photomicrographs represent
perfectly spherical grains sectioned at their equator :

(9.6)

where the Jeffries number, Aj in fim , is computed as :


2

where Ap is the frame area, in yum , and NQ is the number of grains measured. The
2

measured equivalent prior-austenite grain diameter was 10.1 pm, which is similar to the
grain diameter (10.8 pm) produced in the Gleeble experiments. Thus, the transformation
2
According to the Jeffries method, grains located at the border of frames are counted as half.
Laboratory Results and Discussion : Quenching Tests 189

kinetic data measured in the tubular samples can be applied to compute the microstruc-
tural evolution produced in the 38.1 mm-dia. alloyed eutectoid steel bars.
Laboratory Results and Discussion : Quenching Tests 190

Table 9.1: Estimated critical heat flux (CHF) using raw and pre-filtered data for a 38.1
mm-dia. IF steel bar quenched in water flowing at 2.8 m s at 25 C. _ I

Unfiltered Filtered
r CHF, M W m - 2
1
CHF' ^ CHF, M W m - 2
1
CHF' ^
2 5.3 445.4 5.3 411.1
4 5.0 410.8 5.0 412.9
5 4.7 373.9 4.8 414.2

Table 9.2: Estimated critical heat flux (CHF) as a function of water velocity during
forced convective quenching of 38.1 mm-dia. IF steel bars in water flowing at 3 water
temperatures.

qC H F ' - M W m 2

v, m s 1
T = 25 C T = 50 C Tf = 75 C
f f

2.8 5.0 3.7 2.6


4.8 6.4 5.8 3.8
6.9 8.1 7.1. 4.5

Table 9.3: Average heat flux as a function of water velocity during forced convective
quenching of 38.1 mm-dia. IF steel bars in water flowing at 3 water temperatures.

c , M W m-2
l a v a

v, m s 1
T = 25 C
f T = 50 C T -=
f f 75 C
2.8 2.98 1.96 1.18
4.8 3.90 3.31 1.71
6.9 4.82 4.27 2.21
Laboratory Results and Discussion : Quenching Tests 191

o RUN10

RUNIB

u
S
u
CD
a.

s IF s t e e l
-1
-

o V = 4.8 ms
H
o -
= 50 C

100

Figure 9.1: Measured temperature response at the centre and subsurface of two 38.1
mm-dia. IF steel bars quenched with water flowing at 4.8 m s at 50 C . - 1

1100
0

1000 = 75 C
o T, = 50 c
900 * T
f
= S5 c
800

700
0)
600
-p
(0
u IF steel
500
v = 2.8 m s~
p.

s 400 Centre

a> 300
H

200

100

0
20 40 60 80 100 120

Time, s

Figure 9.2: Measured temperature response at the centre of a 38.1 mm-dia. IF steel bar
quenched with water flowing at 2.8 m s for 3 values of water temperature. Note the
_1

recalescence when water at 75 C was used.


Laboratory Results and Discussion : Quenching Tests 192

1100

1000 = 75 C
T
( = 50 C
900 A 0
' T
, = 25 C

800

700

600

500

6 400 IF steel

v = 4.8 m s
300
Centre

200

100 1 O V
' V ^
OOQOOOOO
0
20 40 60 80

Time, s

Figure 9.3: Measured temperature response at the centre of a 38.1 mm-dia. IF steel bar
quenched with water flowing at 4.8 m s for 3 values of water temperature. Note the
_1

recalescence when water at 75 C was used.

1100
0
1000 v T
f = 75 C
T
f
= 50 C
900
' T
, = 25 c

800

700

600

500 IF steel

s 400
v

Centre
= 6.9 m s

<u

300

200

100
A ^ o o o o o o o o o o o o o
0
20 40 60

Time, s

Figure 9.4: Measured temperature response at the centre of a 38.1 mm-dia. IF steel bar
quenched with water flowing at 6.9 m s for 3 values of water temperature. Note the
_1

recalescence when water at 75 C was used.


Laboratory Results and Discussion : Quenching Tests 193

uoo
1000
v v= 2.8 m s

o v = 4.8 m s
900
v = 6.9 m s
800

700

600

500
0
400 h
IF steel
300

200

100

0
0 10 20 30 40 50

Time, s

Figure 9.5: Measured temperature response at the centre of a 38.1 mm-dia. IF steel bar
quenched with water flowing at 25 C for 3 values of water velocity.

1100

1000 v = 3.8 m s

v = 4.8 m s
900
v = 6.9 m s

800

700

600

CD 500
ex,
CO 400

300 IF steel
50 C
200
Centre
100
*BTttUiHp
0
20 40 60 80

Time, s

Figure 9.6: Measured temperature response at the centre of a 38.1 mm-dia. bar quenched
with water flowing at 50 C for 3 values of water velocity.
Laboratory Results and Discussion : Quenching Tests 194

1100
1000 v = 2.8 rn s
v = 4.8 m s
900 v = 6.9 m s
* O v

800
700

3 600
ctf
a) 500
s 400
H IF s t e e l
300 75 C

200 Centre

100 ***8gi>itii iii ^^


tM l!

0
20 40 60 80 100 120
Time, s

Figure 9.7: Measured temperature response at the centre of a 38.1 mm-dia. bar quenched
with water flowing at 75 C for 3 values of water velocity.

800
-1
V V = 2.8 m s
700
o V 4.8 m
s_ i
V = 6.9 m
s i
600

500
V A

O
400

300

200

100 IF steel
25 C
0 r
I i i i i I

1000 800 600 400 200


T (subsurface), C

Figure 9.8: Measured temperature difference between the centre and the subsurface as a
function of subsurface temperature during forced convective quenching of a 38.1 mm-dia.
IF steel bar in water flowing at 25 C for 3 values of water velocity.
Laboratory Results and Discussion : Quenching Tests 195

800

- l
700 V V = 2.8 m s
o V = 4.8 m s_,
A V = 6.9 m s
_ 600

V 500

400

300

200 V

100 IF s t e e l
T = 50 C
(

1000 BOO 600 400 200

T (subsurface), C

Figure 9.9: Measured temperature difference between the centre and the subsurface as a
function of subsurface temperature, during forced convective quenching of a 38.1 mm-dia.
IF steel bar in water flowing at 50 C for 3 values of water velocity.

800
-l
V = 2.8 m
700 s
o V = 4.8 m s
V = 6.9 m s 1
600

500

400

300

200

100 IF steel
T = 75 C
3 (

1000 BOO 600 400 200

T (subsurface), C

Figure 9.10: Measured temperature difference between the centre and the subsurface as a
function of subsurface temperature during forced convective quenching of a 38.1 mm-dia.
IF steel bar in water flowing at 75 C for 3 values of water velocity.
Laboratory Results and Discussion : Quenching Tests 196

-l
50 v v = 2.8 m s i

o V = 4.8 m s
-l
V = 6.9 m s

3
v
v
-50

< -100

IF s t e e l
-150 25 C
f

Centre

-200
200 400 600 800 1000
o
Temperature, C

Figure 9.11: Cooling rate at the centre as a function of local temperature during forced
convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at 25 C for 3 values
of water velocity.

-1
V = 2.8 m
h
V s
50 -l
o v = 4.8 m s
* V = 6.9 m s
-l

-50

< -100

IF s t e e l
-150 T = 50 C
(

Centre

-200
200 400 600 800 1000
o
Temperature, C

Figure 9.12: Cooling rate at the centre as a function of local temperature during forced
convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at 50 C for 3 values
of water velocity.
Laboratory Results and Discussion : Quenching Tests 197

-l
50 V V 2.8 m s
o V = 4.8 m s - l
4 V = 6.9 m s - l

-50
v t> v v v

o o >

< -100

IF steel
-150 75 C
Centre

-200
200 400 600 BOO 1000
Temperature,

Figure 9.13: Cooling rate at the centre as a function of local temperature during forced
convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at 75 C for 3 values
of water velocity.

200
O
o
A , == 25
T c
50 c

150

, = 75 c
T

100

<

50

4 6

Water velocity, m s

Figure 9.14: Maximum cooling rate at the centre as a function of water velocity dur-
ing forced convective quenching of a 38.1 mm-dia. IF steel bar for 3 values of water
temperature.
Laboratory Results and Discussion : Quenching Tests 198

50 h

0 2 4 6 8

Water v e l o c i t y , m s

Figure 9.15: Maximum cooling rate at the subsurface as a function of water velocity
during forced convective quenching of a 38.1 mm-dia. IF steel bar for 3 values of water
temperature.

1000
Clean
Light scale
Heavy scale
800

SH
600

u
1) IF Steel
400 h v = 4.8 m s
T = 50 C
f

Centre
200

80

Figure 9.16: Effect of surface condition on the temperature response at the centre during
forced convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at 4.8 m s _1

at 25 C .
Laboratory Results and Discussion : Quenching Tests 199

Clean
O Oxidized

20 40 60 80 100

Time, s

Figure 9.17: Effect of surface condition on the temperature response at the centre during
forced convective quenching of a 38.1 mm-dia. alloyed steel bar in water flowing at
4.8 m s- at 32 C .
1

Figure 9.18: Effect of initial test temperature on the temperature response at the centre
during forced convective quenching of a 38.1 mm-dia. IF steel bar in water flowing at
4.8 m s- at 32 C .
1
Laboratory Results and Discussion : Quenching Tests 200

0 10 ' 20 30 40 50

Time, s

Figure 9.19: Raw and filtered temperature response at the subsurface during forced
convective quenching of a 38.1 mm-dia. IF in water flowing at 2.8 m s at 25 C. The
_1

inserts show the performance of the filtering procedure at two selected regions of the
curve.
Laboratory Results and Discussion : Quenching Tests 201

1100

0 1
i , , , , i

0 20 40 60 80

Time, s

Figure 9.20: Experimentally determined temperature response at the subsurface and


centreline during forced convective quenching of a 38.1 mm-dia. IF steel bar with water
flowing at 2.8 m s' at 25 C.
1

IF Steel
-l
V = 2.8 m s
8 h T
f = 25 C
7 h

Figure 9.21: Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar. The raw data of Figure 9.20
were filtered before being input to the computer program.
Figure 9.22: Comparison between estimated surface heat flux as a function of surface
temperature, using raw and filtered data, during forced convective quenching of a 38.1
mm-dia. IF steel bar with water flowing at 2.8 m s at 25 C .
- 1

Figure 9.23: Effect of the parameter r on the residuals obtained during the estimation
of surface heat flux during forced convective quenching of a 38.1 mm-dia. IF steel bar
when raw data was used.
Laboratory Results and Discussion : Quenching Tests 203

50 80

40 70
r = 2
30 r = 4 60
r = 6
20 50
u u

o
o
10 40 o
"5
H 0 |- ^ 0
I 30 H
-10 1
20 ft
H _ 20 D

30 10 H

40 0

-10
50 _L
200 400 600 800 1000
-20
o
^exp' C

Figure 9.24: Effect of the parameter r on the residuals obtained during the estimation
of surface heat flux during forced convective quenching of a 38.1 mm-dia. IF steel bar
when filtered data was used.

Figure 9.25: Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar with water flowing at 2.8 m s _1

for 3 values of water temperature.


Laboratory Results and Discussion : Quenching Tests 204

0
....... T f
C25 -
T
f = 50 C -

....... o
75
c
T f =

7h

0 200 400 600 800 1000

Figure 9.26: Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar with water flowing at 4.8 m s _1

for 3 values of water temperature.

Figure 9.27: Estimated surface heat flux as a function of surface temperature, during
forced convective quenching of a 38.1 mm-dia. IF steel bar with water flowing at 6.9 m s - 1

for 3 values of water temperature.


Laboratory Results and Discussion : Quenching Tests 205

100

T = 25 C
f

T = 50 C
(

T = 75 C
10
f

v = 2.8 m s

0.01
10 20 30 40 50 60 70
Time, s

Figure 9.28: Estimated surface heat flux as a function of time, during forced convective
quenching of a 38.1 mm-dia. IF steel bar with water flowing at 2.8 m s" for 3 values of
1

water temperature.

1000

Figure 9.29: Surface heat flux as a function of wall superheat, during forced convective
quenching of thin platinum wires [218].
Laboratory Results and Discussion : Quenching Tests 206

10*

v = 2.8 m/s

10'

& 10"

io _1
....... _
25
O

c
T f

T
f = 50 c
o
f = 75
c
T

10
200 400 600 800 1000
o
AT
sat'
c

Figure 9.30: Estimated surface heat flux as a function of wall superheat, during forced
convective quenching of a 38.1 mm-dia. IF steel bar with water flowing at 2.8 m s for - 1

3 values of water temperature.

1000

Figure 9.31: Estimated surface heat flux as a function of wall superheat, during forced
convective quenching of a 38.1 mm-dia. IF steel bar with water flowing at 4.8 m s for _1

3 values of water temperature.


Laboratory Results and Discussion : Quenching Tests 207

10"

v = 6.9 m/s

10*

10

10" 1

....... T f 25 c
=
O

T
f = 50 c
o
75 c

10
200 400 600 800 1000
o
AT C

Figure 9.32: Estimated surface heat flux as a function of wall superheat, during forced
convective quenching of a 38.1 mm-dia. IF steel bar with water flowing at 6.9 m s for _1

3 values of water temperature.

Figure 9.33: Estimated and calculated (film boiling) surface heat flux as a function of
surface temperature, during forced convective quenching of a 38.1 mm-dia. IF steel bar
with water flowing at 2.8 m s for 3 values of water temperature.
_1
Laboratory Results and Discussion : Quenching Tests 208

10
o
- o T
f = 25 C
o
A T
f = 50 c
o
T
f = 75 c

Is

2 h

4 5 6

Water velocity, m s

Figure 9.34: Estimated critical heat flux (CHF) as a function of water velocity, dur-
ing forced convective quenching of a 38.1 mm-dia. IF steel bar for 3 values of water
temperature.

0
O T
f C = 25
50 o
,=
T
5 A
c
o
T, = 75 c
f

Water velocity, m s

Figure 9.35: Average heat flux as a function of water velocity, during forced convective
quenching of a 38.1 mm-dia. IF steel bar for 3 values of water temperature.
Figure 9.36: Estimated heat-transfer coefficient as a function of surface temperature,
during forced convective quenching of a 38.1 mm-dia. IF steel bar with water at 25 C
for 3 values of water velocity. A value of r = 2 was adopted for filtered data.
Laboratory Results and Discussion : Quenching Tests 210

30 pm

30 pm

00

Figure 9.37: Photomicrographs of a 38.1 mm-dia. IF steel bar quenched with water flow-
ing at 4.8 m s at 25 C , at (a) ~ 3 mm from the surface and (b) centre. Magnification :
_1

500 X .
Laboratory Results and Discussion : Quenching Tests 211

Figure 9.38: Photomicrographs of a 38.1 mm-dia. alloyed eutectoid steel bar quenched
with water flowing at 2.8 m s at 75 C, at (a) ~ 3 mm from the surface and (b) centre.
- 1

Magnification : 400 X .
Laboratory Results and Discussion : Quenching Tests 212

2 mm

Figure 9.39: Photograph of the macrostructure of the 38.1 mm-dia. 1045 carbon steel
bar quenched with water flowing at 2.8 m s at 50 C. Note the martensitic ring (white)
_1

at the surface. Magnification : 8 X .

Figure 9.40: Photomicrographs of a 38.1 mm-dia. 1045 carbon steel bar quenched with
water flowing at 2.8 m s at 50 C. (a) within the martensitic ring, (b) in the transition
_1

zone and (c) at the centre. Magnification : 800 X .


Laboratory Results and Discussion : Quenching Tests 213

Figure 9.40: Photomicrographs of a 38.1 mm-dia. 1045 carbon steel bar quenched with
water flowing at 2.8 m s at 50 C. (a) within the martensitic ring, (b) in the transition
_1

zone and (c) at the centre. Magnification : 800 X .


Laboratory Results and Discussion : Quenching Tests 214

70 900
' I l i'i l i i
1 1 1
1 1 r

60

50
o
05
33
40

0)
Pi
13 30
u
C6
w A Alloyed Steel
20 O 1045 Steel
IF S t e e l

10

i i i i i i i

-20 - 1 6 -12 -8 -4 0 4 8 12 16 20

Radial Position, mm

Figure 9.41: As-quenched hardness distribution in 38.1 mm-dia. IF, alloyed and 1045
carbon steel bars. Use the right axis to read hardness in the IF steel bar.

Figure 9.42: Hardness distribution near the ends and at mid-length in a 100 mm-dia.
alloyed eutectoid steel bar quenched and tempered under industrial conditions.
Laboratory Results and Discussion : Quenching Tests 215

Figure 9.43: Outline of prior-austenite grain boundaries in a 38.1 mm-dia. alloyed eu-
tectoid steel quenched bar, etched in a boiling alkaline sodium picrate solution. Magni-
fication : 1000 X .

140

120

100

I 80

21 60

40

20

0
2
Area, /xm

Figure 9.44: Measured prior-austenite grain area distribution in a 38.1 mm-dia. alloyed
eutectoid steel quenched bar.
Chapter 1 0

Laboratory Results and Discussion : Transformation Kinetics

The results from the transformation kinetics study are presented in this chapter. First,
the transformation kinetics measured under continuous cooling conditions are discussed,
followed by the results from the isothermal tests. Calculated phase boundaries, measured
prior-austenite grain size and measured ferrite fraction are also presented.

10.1 Phase B o u n d a r i e s

The theoretical phase boundaries for the three alloyed eutectoid steels and the 1045
carbon steel were calculated using a computer program based on the algorithm proposed
by Kirkaldy et al. [104]. The start of the austenite-to-martensite transformation, M ,s

was calculated using the empirical linear chemistry-sensitive formula given by Andrews
[116] while the start of the austenite-to-bainite transformation, B , was computed using
s

the empirical formula proposed by Kirkaldy et al. [14]. The results are summarized in
Table 10.1.

216
Laboratory Results and Discussion : Transformation Kinetics 217

10.2 Continuous C o o l i n g Tests

10.2.1 A l l o y e d E u t e c t o i d Steels

Typical measured cooling curves are shown in Figure 10.1. Note that recalescence was
observed for the intermediate cooling rates of 17.5 and 8 C s _ 1
only. The correspond-
ing dilation vs temperature curves are shown in Figure 10.2. The data from each run
were filtered by averaging every 5 points, which smoothed the curves and resulted in a
reduction of the total number of data points to be analyzed.
The analysis of the filtered data was carried out as follows :

1. Compute the change in diameter due to thermal contraction of the parent phase . 1

This was done by fitting the austenite data obtained above the transformation start
temperature by linear regression.

2. Compute the change in diameter due to thermal contraction of the product phase.
This was done by fitting the product phase data obtained below the transformation
finish temperature by linear regression.

3. Compute the difference between the experimentally measured dilation (thermal


contraction plus expansion due to transformation) and the change in diameter due
to thermal contraction only, and determine the start and end of the transformation.

These calculations are illustrated in Figures 10.3 (steps 1 and 2), 10.4 (step 3 : start of
the transformation), and 10.5 (step 3 : end of the transformation) for the 17.5 C s - 1
of
the alloyed eutectoid steel A .

dilation data measured with the Gleeble 1500 simulator is given in arbitrary length units. The
following relationship was used to transform the data to milimeters :

D = 2.88315 D,meas + D0

where D is the initial sample diameter (mm) and D is the measured dilation (arbitrary length
0
meas
]

units).
Laboratory Results and Discussion : Transformation Kinetics 218

The experimentally determined C C T diagrams for the alloyed eutectoid steels A ,


B and C are shown in Figures 10.6 to 10.8. From metallographic observations it was
determined that austenite transformed to pearlite for cooling rates up to 17 C s _1
(air
cooling); a mixture of bainite plus martensite was obtained at the higher cooling rates.
It should be noted that only one test (ACC06-1 : cooling rate = 58 C s ) produced a _1

structure composed of 100 % martensite. The experimentally determined martensite start


temperature ( M J was 218 2 C, 221.5 1.5 C and 229 2.5 C for steels A , B and
C, respectively, which compares favourably with the corresponding values (216, 221 and
234 C , respectively) obtained using the empirical formula proposed by Andrews [116].

The transformation start temperature as a function of cooling rate is shown in Fig-


ure 10.9 for the three alloyed eutectoid steels. As can be seen, the three steels exhibit
very similar behaviour for all the cooling rates examined in the study, with steel A hav-
ing the highest start temperature for any given cooling rate. From these results the
start temperature was correlated with the cooling rate, for cooling rates up to 44 C s , -1

according to the following relations :

638.2 - 74.0 C R + 9.9 C R + 7.71 C R ,


2 3
C R < 17 C s
- i
-651.8 + 2868.0 C R - 2081.6 C R + 467.6 C R , C R > 17 C s
2 3

where T t is the transformation start temperature (in C) and C R is the cooling rate
(in C s ) . Measured and predicted (using Eq. (10.1)) start temperatures are shown in
_1

Figure 10.10 (for the alloyed eutectoid steel A); excellent agreement can be observed.
To predict microstructural evolution, correlations for the C C T start time as a function
of undercooling below the appropriate asymptote are required. From the measured C C T
start time (tAVcer), the following correlation was developed :

\nt T
AVCC = 27.9 + 0.02 U l - 5.654 l n u x
(10.2)
Laboratory Results and Discussion : Transformation Kinetics 219

where u\ is the undercooling below T - Al The measured and predicted (using Eq. (10.2)),
C C T start times are shown in Figure 10.11; good agreement can be seen.

The thermal expansion coefficient of the parent and product phases was computed
from the corresponding slopes of the dilation vs temperature curves. The following rela-
tion was used :
a = 2.88315 J (10.3)

where m is the slope of the dilation vs temperature curve (steps 1 and 2 in the data
reduction procedure described above) and D is the initial sample diameter (mm). The
0

measured average values of the thermal expansion coefficient for austenite, pearlite and
bainite obtained using data from the three alloyed eutectoid steels are given in Table 10.2.

10.2.2 1045 C a r b o n Steel

A similar analysis was applied to results of the C C T tests conducted for the 1045 carbon
steel. The experimentally determined C C T diagram is shown in Figure 10.12. The
following correlation between start temperature and cooling rate was obtained from the
experimental data :

T g t = 700 - 6.35 l n C R + 6.56 l n C R - 8 . 1 0 I n C R


2 3
(10.4)

The transformation start temperature, T t as a function of cooling rate is plotted in


Figure 10.13, together with predicted values obtained using Eq. (10.4); good agreement
was obtained.

Regression equations to predict the C C T start time, t vccr, as a function of under-


A

cooling were also developed for the different microconstituents. For ferrite, the following
equation was obtained :

In t ccT-F
AV = 80.86 + 0.151 u - 20.694 In u
3 3
(10.5)
Laboratory Results and Discussion : Transformation Kinetics 220

where u$ is the undercooling below TA - The results are plotted in Figure 10.14 and, as
3

can be seen, the agreement between measured and predicted values is good.

The following regression equation for the C C T pearlite start time as a function of
undercooling below T ,Al ui, was obtained :

In t ccT-p
AV = 113.17 + 0.26 u - 30.193 In U
x l (10.6)

The measured and predicted values are shown in Figure 10.15; good agreement was
obtained.
The thermal expansion coefficient of the parent and product phases was computed
from the dilation vs temperature curves, as described above for the alloyed eutectoid
steel. The results are given in Table 10.2. For comparison, values of thermal expansion
coefficient for a Ck45 ( S A E 1045) steel reported in the literature (Ref. [222]) are also
included.

10.3 Isothermal Tests

The results of the isothermal tests obtained for the 3 alloyed eutectoid steels are presented
in this section. Typical experimental isothermal dilation vs time curves for the alloyed
eutectoid steel A are shown in Figure 10.16 for the range 525 C < T < 665 C . At
605 C and above, austenite transformed to pearlite, whereas below that temperature
the products were pearlite/bainite or bainite only. Note the enhanced transformation
rate as the test temperature decreases. Prior to the analysis, data from each run was
filtered by averaging it (moving average) every 5 points, which smoothed the curves
and resulted in a reduction of the total number of data points to be analyzed. The
experimentally determined isothermal diagrams for the three alloyed eutectoid steels are
shown in Figures 10.17 to 10.19 along with predicted 1 and 99 % isotransformation
curves. The predicted curves were computed from the correlations given by Park and
Laboratory Results and Discussion : Transformation Kinetics 221

Fletcher [223]. Their correlations were derived by fitting the results of isothermal tests
conducted on 19 eutectoid steels (base composition : 0.75 % C, 0.28 % Si). The prior
austenite grain size of all steels studied in their investigation was approximately A S T M
number 3, whereas the prior austenite grain size in the present study was 9-10 A S T M (see
section 10.4). As can be seen, the experimental results obtained in this work are consistent
with the values predicted adopting the formulae given by Park and Fletcher [223].

The Avrami equation (Eq. (2.4)) was adopted to model the kinetics of the isothermal
transformation. In order to estimate the Avrami kinetic parameters (n, b, and t v)
A

for each isothermal test, a weighted, 5-parameter, non-linear regression analysis was 2

performed on the experimental data. The 5 parameters adopted in the regression model
were : n, 6, t v,A D min (measured dilation at the start of the transformation) and D max

(measured dilation at the end of the transformation).


The estimated kinetic parameters, for the pearlitic reaction, obtained with this proce-
dure are shown in Figure 10.20. For comparison, estimated parameters for weighted and
non-weighted analyses are included. Clearly, the use of weighting functions not only re-
sults in better regressions (from visual inspection of measured and estimated D vst plots)
but also the estimated parameters show less scatter. Assuming the reaction is additive,
the parameter n should be independent of temperature. A value of n = 1.83 0.18 was
computed by averaging the estimated n values obtained with the weighted, non-linear
regression. Using this average value, the kinetic parameter b was re-estimated by re-
applying the weighted, multiple non-linear regression analysis to the data but for a fixed
value of n n. The re-estimated values of b are shown in Figure 10.21, along with a
regression line given by :

l n o = -5.4 exp(42.8/ui) (10.7)

where u\ represents undercooling below T.


Al

See Appendix F for details.


2
Laboratory Results and Discussion : Transformation Kinetics 222

Based on the 1 and 99 % transformation curves, predicted using the formulae pro-
posed by Park and Fletcher [223], the Avrami parameters for the austenite-to-pearlite
transformation were estimated for the three alloyed steels studied in the present investiga-
tion. The average value of n obtained in this fashion was 1.55 (compared with 1.83 0 . 1 8
obtained in this study). The computed values of the parameter b as a function of under-
cooling (derived from the theoretical curves) are shown in Figure 10.22, along with the
values estimated in this work. For comparison, values reported by Campbell et al. [43]
for the pearlitic transformation in a plain carbon eutectoid steel are also shown.
A series of experiments were also conducted to determine the Avrami kinetic param-
eters for the austenite-to-bainite transformation in the alloyed eutectoid steel A . The
isothermal tests were conducted in the range 350 C < T < 480 C . The kinetic pa-
rameters were estimated with the 5-parameter, nonlinear equation described above. The
resulting values of the kinetic parameter, n, are plotted in Figure 10.23. The average
value of n was h = 1.91 0.4. Using this estimate, the parameter b was re-estimated by
fixing n = n in the regression analysis. The estimated values are shown in Figure 10.24,
along with a regression line given by :

In 6 = -42.6 exp-540.2/ui (10.8)

where u\ represents undercooling below T.


Al

10.4 P r i o r - A u s t e n i t e G r a i n Size

To measure the prior-austenite grain size produced during austenitizing of the tubular
samples, a fully martensitic specimen (alloyed eutectoid steel A cooled at 58 C s ) was _1

prepared metallographically. The microstructure revealed by a boiling alkaline sodium


picrate etch is shown in Figure 10.25. To determine the prior-austenite grain size, the
mean chord length in a total of 66 grains was measured from several frames. The resulting
Laboratory Results and Discussion : Transformation Kinetics 223

mean chord length distribution is shown in Figure 10.26. The computed average chord
length is 10.8 /mi, which corresponds to a grain size of 9-10 A S T M .

10.5 Ferrite Fraction

Results of the ferrite fraction measured in the continuously cooled 1045 carbon steel
specimens as a function of cooling rate at 750 C are shown in Figure 10.27. Each point
in the plot represents the mean of no less than fifty separate fields measured on the image
analyzer, as described previously. The standard deviation of each measurement is given
as an error bar. For comparison, the values reported by Hawbolt et al. [98] for a 1025
carbon steel are also shown in the figure. The dotted line represents the extrapolation to
the equilibrium value. The ferrite fraction decreases monotonically from its equilibrium
value as the cooling rate increases. As expected, the ferrite fraction in the 1045 carbon
steel is smaller at all cooling rates.
The ferrite fraction in two tests on the alloyed eutectoid steel was determined in a
similar fashion. The tests selected corresponded to the two slowest cooling rates (0.2
and 0.3 C s ) for steel B . The results showed less than 1 % ferrite for a cooling rate of
_1

0.2 C s ; no significant trace of ferrite was found when the cooling rate was 0.3 C s .
_1 _1

Since the primary phase formed was pearlite, the alloyed steels were treated as eutectoid.
Laboratory Results and Discussion : Transformation Kinetics 224

Table 10.1: Phase boundaries calculated from the equations by Kirkaldy et al. [104], the
empirical formula given by Andrews [116], and the empirical formula given by Kirkaldy
et al. [14].

Steel
Alloyed Eutectoid A Alloyed Eutectoid B Alloyed Eutectoid C 1045 Plain Carbon
Kirkaldy et al. [104]
730 730 731 756
724.4 724 725 725
Andrews' formula [116]
M s 216 221.1 234.3 323.6
M f
1 6.1 . 19.3 108.6
Kirkaldy et al. [14]
B s
557 558 560 585

Table 10.2: Measured expansion coefficients for the alloyed eutectoid and the 1045 carbon
steel. For comparison, values reported in the literature are also given.

Phase Thermal Expansion Coefficient, C 1

Alloyed Eutectoid Steel


This Work Ref. [215]
Austenite 2.46 x l O- 5
2.05 xlO - 5

Pearlite 1.67 x l 0 ~
5

Bainite 1.63 x l O- 5

1045 Carbon Steel


This Work Ref. [222]
Austenite 2.22 x l O "5
2.1 xlO - 5

Pearlite/Bainite 1.61 x l O - 5
1.4 xlO - 5
Laboratory Results and Discussion : Transformation Kinetics 225

Figure 10.1: Experimentally determined cooling curves for the alloyed eutectoid steel B
showing the cooling rate for each test.

100 200 300 400 500 600 700 800

Time.s

Figure 10.2: Experimentally determined dilation-temperature curves obtained for con-


tinuous cooling of the alloyed eutectoid steel B showing the cooling rate for each test.
Laboratory Results and Discussion : Transformation Kinetics 226

0.40

Experimental
Due to t h e r m a l
contraction
0.35

ti
>>
>

a
0.30

0.25

0.20
300 400 500 600 700 800 900
o
Temperature, C

Figure 10.3: Continuous cooling test showing the thermal contraction of the austenite
(high temperature) and the low temperature product phase obtained for the 17.5 C s _1

cooling of the alloyed eutectoid steel B .


800

J3
a

-200
400 500 600 700 800

Temperature,

Figure 10.4: Continuous cooling test showing the procedure to determine the start of
transformation for the 17.5 C s cooling of the alloyed eutectoid steel B . The dotted
_1

lines represent 3 standard deviations.


Laboratory Results and Discussion : Transformation Kinetics 227

140

120

100 h

80

60
transf ormation
finish
40
a
20 h + 3(7
a
3 j i0^f%m

-20 500 600


100 200 300 400
o
Temperature, C

Figure 10.5: Continuous cooling test showing the procedure to determine the end of
transformation for the 17.5 C s cooling of the alloyed eutectoid steel B . The dotted
_1

lines represent 3 standard deviations.

10000

Figure 10.6: Continuous cooling diagram for the alloyed eutectoid steel A . Closed cir-
cles represent the start of diffusional transformations; open circles represent the end of
diffusional transformations.
Laboratory Results and Discussion : Transformation Kinetics 228

o' '
1 10 100 1000 10000

Time.s

Figure 10.7: Continuous cooling diagram for the alloyed eutectoid steel B . Closed cir-
cles represent the start of diffusional transformations; open circles represent the end of
diffusional transformations.

Figure 10.8: Continuous cooling diagram for the alloyed eutectoid steel C. Closed cir-
cles represent the start of diffusional transformations; open circles represent the end of
diffusional transformations.
Laboratory Results and Discussion : Transformation Kinetics 229

Cooling Rate C / m i n

10 3
10 2
10 1

800 1 1 1 1
1 | M 1 1 1 1 1 1 | 1 1 1
T
Al

700 - . A

600 - -

500 _ _

p.

a
A Steel A
o Steel B
DO Steel C

200
10' 10 1
10" 10"

C o o l i n g Rate, C / s

Figure 10.9: Transformation start temperature as a function of cooling rate for the alloyed
eutectoid steels A , B and C.

BOO

700

CD 600 \-
U
->
cS
u
<D
500

-P
400
(0
O Data
Predicted
300

200 1 1

10 2
10 1
10 10" 1

0 .
Cooling Rate, C / s

Figure 10.10: Transformation start temperature plotted as a function of cooling rate for
the alloyed eutectoid Steel A . The line plotted in the figure is based on Eq. (10.1).
Laboratory Results and Discussion : Transformation Kinetics 230

800
1 1

750 - O Measured -
Predicted

700
-

650

600 - o / -

550

o
500

450 -

400
10 u
10 1
10' 10 10*

Figure 10.11: Measured and predicted C C T transformation start times for the alloyed
eutectoid Steel A . The line plotted in the figure is based on Eq. (10.2).

1000

Figure 10.12: Continuous cooling diagram for the 1045 carbon steel. Closed circles repre-
sent the start of diffusional transformations; open circles represent the end of diffusional
transformations.
Laboratory Results and Discussion : Transformation Kinetics 231

500
1000 100 10 1 0.1
o
C o o i n g Rate, C/s

Figure 10.13: Transformation start temperature plotted as a function of cooling rate for
the 1045 carbon steel. The line plotted in the figure is based on Eq. (10.4).

Figure 10.14: Measured and predicted C C T ferrite start time for the 1045 carbon steel.
The line plotted in the figure is based on Eq. (10.5).
Laboratory Results and Discussion : Transformation Kinetics 232

740

720 h O Measured
Predicted
700

680

660

a.
640

620

600

580
0.1 10 100

Figure 10.15: Measured and predicted C C T pearlite start time for the 1045 carbon steel.
The line plotted in the figure is based on Eq. (10.6).

I , i , i i

0 100 200 300

Time, s

Figure 10.16: Experimentally determined isothermal dilation-time curves for the alloyed
eutectoid steel A .
Laboratory Results and Discussion : Transformation Kinetics 233

700

600 This work :


A 1 %
A 99 %

500 Park and Fletcher :


o 1 %

99 %

g 400

300 h

200
10 u
10 1
10* 10 J
10 4
10

Time, s

Figure 10.17: IT diagram for the alloyed eutectoid steel A . Closed circles represent the
start of diffusional transformations; open circles represent the end of diffusional transfor-
mations.

200 i i . '
10 10 1
10 2
10 3
10* 10 5

Time, s

Figure 10.18: IT diagram for the alloyed eutectoid steel B . Closed circles represent the
start of diffusional transformations; open circles represent the end of diffusional transfor-
mations.
Laboratory Results and Discussion : Transformation Kinetics 234

700 r l ui iii i i 11

h-AH O
IAH
I A H / - ^lij
600 l-^-O
This work : -
-
A 1 %
A 99 %
of 500 h IAH O Park and Fletcher : -
3
O 1 % -
cd
CD

g 400
99 %

CD

300

200
10 u
10 1
10* 10 J
10 4
10

Time, s

Figure 10.19: IT diagram for the alloyed eutectoid steel C. Closed circles represent the
start of diffusional transformations; open circles represent the end of diffusional transfor-
mations.
Laboratory Results and Discussion : Transformation Kinetics 235

2.5

2.0

1.5 e

o
Fit (no weights)
1.0 O Steel A
Steel B
A Steel C
n = f(T) Fit (weighted)
0.5
Steel A
Steel B
A Steel C
0.0
560 580 600 620 640 660 680 700
o
Temperature, C

-2 Fit (no weights)


O Steel A
Steel B
A Steel C
Fit (weighted)
Steel A
-6 Steel B
A Steel C
J2 a
a
-8

-10

A
n = f(T)
-12

-40 -60 -80 -100 -120 -140


o
Undercooling below T , C A

(b)

Figure 10.20: Kinetic parameters for the pearlitic transformation in the 3 alloyed eutec-
toid steels estimated with a 5-parameter, non-linear regression analysis (weighted and
non-weighted), (a) n; (b) In b.
Laboratory Results and Discussion : Transformation Kinetics 236

o Steel A
o Steel B
A Steel C
Best Fit

ti

-10 h

f = A exp(b/x)
-12 A = -5.4
b = 42.8

-14
60 80 100 120 140
Undercooling below T.

Figure 10.21: Kinetic parameter b, as a function of undercooling, for the pearlitic transfor-
mation in the 3 alloyed eutectoid steels estimated with a weighted, 5-parameter non-linear
regression analysis, for a constant value of n = n = 1.83 0.18. The line plotted in the
figure is based on Eq. (10.7).

-2

-4

-6

ti -8

-10

-12 This work


O Campbell et al.
-14 Park et al.

-16
60 80 100 120 140
Undercooling below T , A

Figure 10.22: Kinetic parameter b, as a function of undercooling, for the pearlitic trans-
formation in 3 alloyed eutectoid steels, as estimated in this study (solid line); derived from
the regression equations of Park et al. [223] (filled circles); and reported by Campbell
et al. [43] (for a eutectoid carbon steel).
Laboratory Results and Discussion : Transformation Kinetics 237

3.0

2.5 \-

2.0

a 1.5
n = 1.91 i 0.4

1.0

0.5
Steel A

0.0
320 360 400 440 480
o
Temperature, C

Figure 10.23: Kinetic parameter n for the bainitic transformation in the alloyed eutectoid
steel A , estimated with a 5-parameter, weighted, non-linear regression analysis.

O Steel A
-4 Predicted.

-8

-10

f = A exp(b/x)
A = -42.6
-12
b = -540.2

14
240 280 320 360 400
o
U n d e r c o o l i n g below T A , C

Figure 10.24: Kinetic parameter b for the bainitic transformation in the alloyed eutectoid
steel A as a function of undercooling, estimated with a weighted, 5-parameter non-linear
regression analysis, for a constant value of n = n 1.91 0.4. The line plotted in the
figure is based on Eq. (10.8).
Laboratory Results and Discussion : Transformation Kinetics 238

15 jim.

Figure 10.25: Microstructure of etched Gleeble specimen (alloyed eutectoid steel A cooled
at 58 C s ) showing the outline of prior austenite grains. Magnification : 1000 X .
-1

35

Figure 10.26: Measured prior-austenite mean chord length distribution obtained in the
alloyed eutectoid steel A Gleeble specimen cooled at 58 C s .-1
Laboratory Results and Discussion : Transformation Kinetics 239

T 1 1 I I I I I | 1 1 1I I I I I | 1 1 1111 1 I | 1 1 1I I 1 I I

10"1
10 10 1
10 2
10 3

0 -1
Cooling Rate, C s

Figure 10.27: Ferrite fraction as a function of cooling rate for the 1045 (open circles) and
1025 (filled circles) [98] carbon steel samples. The uncertainty in the measurements is
shown as error bars.
Chapter 11

Residual Stress Measurement

The residual stress distributions in forced convective quenched IF, 1045 carbon, and
alloyed eutectoid steel bars were determined experimentally by means of neutron diffrac-
tion. The quenching conditions were given in Table 7.6. Due to the greater depth
of penetration of neutrons, complete stress profiles can be measured non-destructively.
Diffraction peak profiles were obtained with the E3 neutron diffractometer at the N R U
reactor ( A E C L - Chalk River Laboratories) in two separate campaigns. In this chapter,
the experimental technique and data reduction adopted are discussed, and the measured
strain and stress fields are presented. A discussion on the generation of stresses during
quenching is given in Chapter 12.

11.1 Experimental Procedure

11.1.1 Campaign 1

In the first set of experiments, the residual stress distributions in IF and alloyed steel
bars were measured. Neutrons of wavelength 2.412 A (1 A = 1 0 _1
nm) were obtained by
diffraction from the (113) planes of a squeezed germanium crystal at a take-off angle, 29 ,
m

of 90. This monochromator plane spacing was selected because a wavelength A > 2.34 A
and high take-off angle are required to optimize the neutron beam penetration and an-
gular resolution of diffraction peaks [224]. The mosaic spread of the monochromator was
0.2, which is well-matched to the angular resolution needed in most residual stress

240
Residual Stress Measurement 241

scanning experiments [143]. The typical scattering angle, 26, for the (110) reflection of
the material 73.

The intersection of the incident and scattered beams defines the sampling volume. The
beams were shaped with slits made of cadmium (which is opaque to thermal neutrons).
The resulting sampling volumes were 2 x 2 x 5 m m (for radial and circumferential strain
3

measurements) and 2 x 2 x 10 m m (for axial strain measurements). These sampling


3

volumes were selected to give a good spatial resolution along the radius of the bars while
maintaining a volume large enough to achieve an adequate signal intensity. The sample
orientation with respect to the incident and diffracted beams for the determination of
the axial, radial, and circumferential (hoop) strains is shown in Figure 11.1. The gauge
volume was centered on the rotational axis of the spectrometer specimen table by first
locating a 2 mm plastic pin on the rotational centre of the table and then translating the
incident slit across the neutron beam to find the maximum intensity with the scattering
angle set at 90. The diffracted-beam slit was also optimized in this way. The spatial
coordinates of the sample surface were determined by measuring the diffracted intensity
as the surface was translated through the gauge volume.
The specimens were mounted on a computer-controlled X Y translator. Strain mea-
surements were made every 1 mm from the centre along the radius of the specimen, at
mid-length, by translating the sample in the appropriate direction. The precision asso-
ciated with the radial position of the sampling volume was 50 fim. To ensure having
the gauge volume inside the specimen, the last radial position was selected to be at 17.5
mm from the centre (the specimen radius is 19.05 mm).

Neutrons were detected with a single H e detector. A diffraction peak profile with 21
3

points was obtained by scanning an angular range chosen to get a complete profile. A set
of measurements along the bar radius required 7.5 hours (for radial and circumferential
scans) and 50 h for (axial scans) to provide diffraction peaks of sufficient counting statis-
tics. The corresponding times for the IF steel were 6 and 40 hours, respectively. The
Residual Stress Measurement 242

reason for the long counting times is the attenuation of the neutron beam as it travels
inside the specimen : the attenuation increases as the sampling volume is translated
towards the centre of the specimen. This effect is particularly noticeable in the axial
measurements, due to thelonger paths involved (see Figure 11.1 (c)).

11.1.2 Campaign 2

In the second set of measurements, the residual stress distribution in forced convective
quenched 1045 carbon steel bars was determined, and measurements on alloyed steel
specimens were repeated. Data from the alloyed steel bars (Campaign 1) were difficult
to analyze, because of a near-overlap of diffraction peaks from martensite and retained
austenite. This difficulty prompted the selection of a diffraction line that did not con-
tain any contribution from the austenite retained in the material. Thus, diffraction from
(112) planes in the specimens was selected for measurements in Campaign 2. Neutrons
of wavelength 1.65 A were obtained by diffraction from the (331) planes of a squeezed
germanium crystal at a take-off angle, 29 , of 79. The mosaic spread of the monochro-
m

mator was 0.2. The typical scattering angle, 29, for the (112) reflection of the steel
was 90.
The gauge volume and sample orientation with respect to the neutron beam were the
same as described previously for the first series of tests. Strain measurements were made
along the radius of the sample, at mid-length, for radii from 0 to 18 mm.

As previously described, diffraction peaks in the alloyed steel bar were collected by
scanning a single detector through a series of scattering angles and measuring neutron
intensity at each angle setting. However, because the diffraction peaks were sufficiently
narrow, it was possible to use a multi-wire, position-sensitive detector to acquire data
more rapidly. Designed and built at Chalk River Laboratories, the detector has 32
grounded anode wires, suspended vertically, in a single chamber of H e gas. Neutrons
3

are captured by the H e atom, and a nuclear reaction generates a proton and triton with
3
Residual Stress Measurement 243

high energy. These particles generate a charge cascade that follows a high electric field
towards the nearest anode wire. The multiwire detector is set to look through a window
in a cadmium mask, which is located as close as possible to the sampling volume, to
minimize parallax effects. The array of wires in the detector spans a range of 2.7 in
scattering angle, in 32 steps. Thus, the detector collects a complete diffraction peak
without any angular motion, so high data-acquisition rates can be achieved. It should
be pointed out that this detector was used for the first time in this set of experiments.

In order to investigate possible microstructural effects (changes in plane spacing non-


related to actual stresses in the material), reference slices approximately 1-2 mm thick
(radial direction) x 4 mm 'wide' (circumferential direction) x 20 mm tall were extracted
from the steel bars. The slices were cut from a 20 mm tall cylindrical sample as shown
in Figure 11.2. The number of slices extracted, and their radial positions, were se-
lected based on the hardness distribution across the various specimens (see Figure 9.41).
Diffraction peak profiles were obtained for each individual slice in both campaigns.

11.2 Results

Typical diffraction peak profiles obtained for the IF, 1045 carbon, and alloyed eutectoid
(both campaigns 1 and 2) steel bars are shown in Figures 11.3 to 11.6. As seen in
Figure 11.5, diffraction peak profiles for the first set of strain measurements on the
alloyed eutectoid steel specimen have the appearance of a double-peak Gaussian curve.
This peak profile is the result of a combination of contributions from the (110) B C T
martensite and the (111) F C C austenite planes. The retained austenite accounts for the
shoulder at the low-angle side of the peak.

The measured diffraction peak profiles were fitted with a 5-parameter function con-
sisting of a Gaussian peak (defined by its position, width and intensity) superimposed
Residual Stress Measurement 244

on a sloping background (defined by its slope and y-intercept.). The precision in the
1

peak mean angle and width was 0.02 and 0.4, respectively; the precision in the
peak intensity, as percentage of the maximum intensity, was typically 5 %. The goodness
of fit for the regressions was estimated from the value of x obtained by comparing the
2

measured and predicted intensities [225] :

(Observed - Expected) 2

* =
2
Expected
(11.1)

The results of the curve fitting of measured diffraction peak profiles did not show variation
of x 2
with radial position, other than that associated with the maximum number of
counts; this is an indication that the measurements, and data reduction, were consistent
along the radius.
Examples of fitted profiles are shown in Figures 11.7 and 11.8 for IF and 1045 car-
bon steel bars, respectively. In the figures, the uncertainty in the measured intensities
is shown as error bars. The 5-parameter function adopted adequately described the
diffraction peaks obtained in both cases. Severe fluctuations in the measured mean an-
gle distribution obtained in the first set of measurements in the alloyed eutectoid steel
specimen, caused by the presence of retained austenite, prevented any meaningful stress
distribution from being deduced from the data. The diffraction peaks obtained for the
(112) martensite planes in the alloyed eutectoid steel in Campaign 2 were comparatively
broad and weak (see Figure 11.6) but without interference from the retained austenite,
resulting in fluctuations in the mean angle along the radial direction. To stabilize the
fitting program, an average slope and y-intercept were calculated for each strain com-
ponent after a first analysis with the fitting procedure, and the data was subsequently
reanalyzed with these fixed values of slope and y-intercept. A typical fit obtained with
this procedure is shown in Figure 11.9. The reproducibility of the technique was assessed

In some cases, the slope of the background was nearly zero. However, the fitting analysis was always
1

done with the 5-parameter function, to retain generality.


Residual Stress Measurement 245

by repeating measurements at a given radial position in the 1045 carbon steel sample. Af-
ter fitting the measured diffraction peaks, the same peak parameters (within the quoted
uncertainties) were obtained for measurements made at r = 2 and r = 3 mm.
The strain component at each radial position was computed by comparing the respec-
tive mean scattering angle (20) with a reference angle (26 ) as follows :0

sin(0)

The average of the mean scattering angles measured for a given stress component (axial,
circumferential and radial) was adopted as the reference angle for that component.
Figure 11.10 shows the variation of mean scattering angle with radial position for the
strain-free reference slices extracted from the 1045 carbon steel specimen. A decrease in
the mean scattering angle from 89.903 to RS 89.865 is observed for the two slices
extracted near the surface. Notice that this mean scattering angle profile coincides with
the microstructure obtained in the quenched bars (Figures 9.39 and 9.40). Given that
these slices are essentially strain-free, the change in mean scattering angle can only be
attributed to a microstructural effect. This variation in the mean scattering angle gives
rise to an isotropic strain of ~ 3.0 0.5 x l 0 ~ 4
m m / m m which must be added to the
strains computed with Eq. (11.2) for radial positions r > 17.05 mm (as determined from
metallography, see Figure 9.39). In contrast, the mean scattering angle was essentially
constant in all the slices extracted from the IF and the alloyed steel specimens. There-
fore, no correction of strains calculated using Eq. (11.2) was required. Interestingly, the
variation of the estimated F W H M (full width at half of maximum) with radial position
in the 1045 carbon steel bar was similar to that observed for the mean scattering angle
in the corresponding reference slices.

Once the residual strain distributions are know, the residual stresses can be computed
as follows. Assuming that the steel bars can be treated as a homogeneous elastic contin-
uum, with elastic constants ( 1 1 2 ) = 225 G N m - 2
and V(u2) = 0.276, and that the radial,
Residual Stress Measurement 246

circumferential, and axial directions correspond to the principal directions of the stress
tensor, the principal stress components may be obtained from Hooke's law, which then
takes the form :

a i i =
T^ i i +
{ l + v)(l-2 ) V '
ekkSii 1 = 3 ( 1 L 3 )

The determination of residual stress distributions via diffraction methods requires a


precise characterization of the unstressed state of the material, i.e., either the interplanar
spacing or the mean scattering angle. When the determination of these parameters
is uncertain, relative to the expected strain values, the overall equilibrium conditions
given by elasticity theory can be adopted to characterize the unstressed state of the
material [141]. Assuming axisymmetric conditions, the residual stresses normal to any
plane must balance :
'a rdrdz
zz =0 (11.4)

/ <roe drdz = 0 (11.5)

In addition, at any surface, the stress acting in the normal direction should vanish. Typ-
ically, the residual stress is known at only few radial positions and, consequently, a nu-
merical procedure needs to be applied to compute the integrals in Eqs. (11.4) and (11.5).
In this investigation, there were two factors that made the determination of the mean
scattering angle in the unstressed material uncertain : 1) the thickness of the reference
slices extracted from the quenched specimens was similar in size to that of the sampling
volume, and 2) the neutron beam attenuation observed during internal measurements
in the quenched cylindrical specimens did not occur when data was collected from the
reference slices. Thus, the following iterative procedure, based on the overall equilibrium
conditions, was adopted to calculate the residual stress distribution of each component
in the specimens :
Residual Stress Measurement 247

1. estimate 0 as the average of measured mean scattering angles;


O

2. compute the residual stress field using Eqs. (11.2) and (11.3);

3. compute the integrals in Eqs. (11.4) and (11.5);

4. if the integrals approach zero then terminate the iteration, if not, guess a new value
of #o and go back to Step 1.

The integrals in Eqs. (11.4) and (11.5) were calculated by discretizing the areas normal to
the axial and circumferential stresses as shown, schematically, in Figure 11.11. The resid-
ual stress acting on each control area was assumed to be constant within that area and
a force increment was calculated by multiplying the stress value times the corresponding
control area. The resultant force was then computed by adding the force increments.

The measured residual strain distributions along the radius of the specimen for the
IF, 1045 carbon, and alloyed eutectoid steel bars are shown in Figures 11.12 to 11.14.
Typical uncertainty in the measured values of strain was 1.3, 2.1, and 2.0 x l O - 4

mm m m - 1
for the radial, circumferential, and axial components, respectively, and is
shown in Figures 11.12 to 11.14 as an error bar.
The residual axial strain in the IF steel bar (Figure 11.12) increases from a compressive
value of - 6 x l O - 4
mm m m - 1
near the surface to a tensile value of ~ + 10 x l O - 4

mm m m - 1
at the centre. The circumferential (hoop) strain was also compressive (-
4 xlO - 4
mm m m ) near the surface and became tensile (+ 2 x l O
- 1 - 4
mm m m ) at- 1

the centre while the radial strain was tensile near the surface (+ 4 x l O - 4
mm m m ) , - 1

decreased to almost zero and was tensile (+ 3 x l O - 4


mm m m ) at the centre.
- 1

The residual axial strain distribution in the 1045 carbon steel bar (Figure 11.13)
was qualitatively similar to that observed in Figure 11.12 for the IF steel specimen;
however, the magnitudes of the compressive strain near the surface and the tensile strain
at the centre were much larger. The circumferential (hoop) strain was compressive (- 12
xlO - 4
mm m m ) near the surface, increased until reaching a maximum of + 4 x l O
- 1 - 4
Residual Stress Measurement 248

mm m m - 1
and decreased towards zero at the centre. The radial strain was tensile (+ 20
xlO - 4
mm m m ) near the surface, increased to a maximum of + 24 x l O
- 1 - 4
mm m m - 1

and decreased towards zero at the centre.


The residual strain distributions in the alloyed eutectoid steel (Figure 11.14) were
very different than those obtained for the IF and 1045 carbon steel bars. Both the
circumferential (hoop) and axial strain profiles were tensile near the surface, decreasing
monotonically to compressive values at the centre. The compressive axial strain at the
centre was larger than the circumferential component. The radial strain was slightly
tensile near the surface, showed a maximum tensile strain of ~ + 8 x l O - 4
mm m m - 1
at
mid-radius and decreased towards zero at the centre.
Strains in a given direction can be influenced by stresses acting in other directions
through the Poisson's ratio and, therefore, it is difficult to deduce mechanical behaviour
from residual strain distributions [143]. The data can be more easily interpreted based
on residual stress results. The residual stress distributions, deduced from the strains
as described above, along the radius of the IF, 1045 carbon and alloyed eutectoid steel
specimens are shown in Figures 11.15 to 11.17. The typical uncertainties in the stress
values, deduced by combining the uncertainties in the strain measurements, were 40,
50, and 50 M P a for the radial, circumferential, and axial components, respectively.
The axial residual stress in the IF specimen (Figure 11.15) was compressive (- 180
MPa) near the surface and tensile (+ 260 MPa) at the centre. The circumferential
(hoop) residual stress was also compressive at the surface and tensile at the centre. The
radial residual stress was close to zero near the surface and increased towards a tensile
value at the centre. As was the case with the measured strains, the axial residual stress
profiles obtained with the IF and 1045 carbon steel specimens were similar. The axial
residual stress distribution measured in the 1045 carbon steel bar (Figure 11.16) showed
a compressive stress (- 430 MPa) near the surface and a tensile stress (+ 500 MPa) at the
centre. The circumferential (hoop) residual stress was compressive at the surface (- 350
Residual Stress Measurement 249

MPa), increased to a maximum of + 350 M P a and decreased to a lower tensile value of


200 M P a at the centre. The radial residual stress was tensile near the surface, increased
to a maximum of approximately 4- 350 M P a and decreased to a value of + 200 M P a
at the centre. In contrast, all the components of the residual stress tensor were tensile
near the surface and decreased towards compressive values at the centre in the alloyed
eutectoid steel bar (Figure 11.17).
Residual Stress Measurement 250

specimen
motion

incident scattered
beam beam

(a)

(b)

Figure 11.1: Specimen orientations with respect to neutron beams to measure (a) radial,
(b) circumferential, and (c) axial strain components in force convective quenched steel
bars. The gauge volume is represented by the shaded area. For clarity, cross-sectional
views are shown.
Residual Stress Measurement 251

(c)

Figure 11.1: Specimen orientations with respect to neutron beams to measure (a) radial,
(b) circumferential, and (c) axial strain components in force convective quenched steel
bars. The gauge volume is represented by the shaded area. For clarity, cross-sectional
views are shown.
Residual Stress Measurement 252
Residual Stress Measurement 253

400 -
IF S t e e l

o
350 h r - 17.5 m m

V V
V
r =
r =
17.0 m m
16.0 m m
300
T r = 15.0 m m
5 D
J D r = 14.0 m m
a
3 250
o
ua
o 200
u
CD
150

100

50

0
-76 -75 -74 -73 -72 -71 -70 -69

Scattering Angle, degrees

Figure 11.3: Measured diffraction peak profiles in a 38.1 mm-dia. IF steel bar quenched
in water flowing at 4.8 m s at 25 C.
- 1

2000 1 > T l
1
i 1
i

V 1045 Steel
V
o r = 16 mm
r = 17 mm
v r = 17. 5 mm -
T r = 18 mm

T
o %
v'o
ri 1000
v
0
V*

f
-
9

500

i . i . i . i

-92 -91 -90 -89 -88

Scattering Angle, degrees

Figure 11.4: Measured diffraction peak profiles in a 38.1 mm-dia. 1045 steel bar quenched
in water flowing at 2.8 m s at 50 C.
_ 1
Residual Stress Measurement 254

"T"

BOO Alloyed Steel


o r = 17 m m
700 r = 15 m m
v r = 13 m m
600

a
3 500

u
a 400
o
3
0) 300
Z
200

100

0
-76 -75 -74 -73 -72 -71 -70 -69

Scattering Angle, degrees

Figure 11.5: Measured diffraction peak profiles in a 38.1 mm-dia. alloyed steel bar
quenched in water flowing at 2.8 m s at 75 C showing the position of the F C C austenite
_ 1

(A(in)) and B C T martensite (M(n )) peaks. Data collected during Campaign 1.


0

900
i 1
1

Alloyed Steel
800 o r = 15 m m
r = 16 m m
v r = 17 m m
700
o o * o
a
3 600
o
o
cl
o
-p 500 o. , v

3
a 8s, v

z.
400

300

200
-91 -90 -89 -88

S c a t t e r i n g Angle, degrees

Figure 11.6: Measured diffraction peak profiles in a 38.1 mm-dia. alloyed steel bar
quenched in water flowing at 2.8 m s at 75 C. Data collected during Campaign 2.
_ 1
Residual Stress Measurement 255

o l_i l i l i i i i i i i l I
-75 -74 -73 -72 -71 -70
Scattering Angle, degrees

Figure 11.7: Measured and fitted diffraction peak profile in a 38.1 mm-dia. IF steel bar
quenched in water flowing at 4.8 m s at 25 C. - 1

2000

o 1
1
1
1
1

-92 -91 -90 -89 -88

S c a t t e r i n g Angle, degrees

Figure 11.8: Measured and fitted diffraction peak profile in a 38.1 mm-dia. 1045 carbon
steel bar quenched in water flowing at 2.8 m s at 50 C. _ 1
Residual Stress Measurement 256

900

200 1
' 1 ' 1 ' 1
' 1

-92 -91 -90 -89 -88

S c a t t e r i n g Angle, degrees

Figure 11.9: Measured and fitted diffraction peak profile in a 38.1 mm-dia. alloyed steel
bar quenched in water flowing at 2.8 m s at 75 C. Data collected during Campaign
_ 1

2.

i i . | i i i i i i i i i

-89.84
o Set 1
grees

* Set 2 0
Average
-89.86 o
CD
T3
co" / A
be A
-89.88
a
<
g
CD -89.90 j _
Mean Scatt.

-89.92

-89.94 -
i i 1 i i i . 1 . . . . 1 .

0 5 10 15 20

Radial Position, m m

Figure 11.10: Mean scattering angle as a function of radial position obtained from the
strain-free reference slices extracted from the 1045 carbon steel specimen.
Residual Stress Measurement 257

Figure 11.11: Control volumes adopted for force balance calculations (Eqs. (11.4)
and (11.5)) : (a) axial component (plan view) and (b) circumferential component.
Residual Stress Measurement 258

o Radial -

Hoop -
A Axial -

03

2 4 6 6 10 12 14 16 18 20

Radial Position, mm

Figure 11.12: Measured radial, circumferential (hoop) and axial residual strains as a
function of radial position in a 38.1 mm-dia. IF steel bar quenched in water flowing at
4.8 m s at 25 C . The uncertainty in the measurements is shown as an error bar.
_1

24

20

16

12

0
G
-4

-8

-12
3 o Radial
T3 -16 Hoop
a>
03 -20 A Axial

-24
1 . 1 , 1 , 1

4 6 8 10

Radial Position, mm

Figure 11.13: Measured radial, circumferential (hoop) and axial residual strains as a
function of radial position in a 38.1 mm-dia. 1045 carbon steel bar quenched in water
flowing at 2.8 m s at 50 C. The uncertainty in the measurements is shown as an error
_1

bar.
Residual Stress Measurement 259

o Radial
Hoop
3 A Axial
S

3
tn
CO Alloyed Steel

2 4 6 8 10 12 14 16 18 20

Radial Position, m m

Figure 11.14: Measured radial, circumferential (hoop) and axial residual strains as a
function of radial position in a 38.1 mm-dia. alloyed steel bar quenched in water flowing
at 2.8 m s _1
at 75 C. Data collected during Campaign 2. The uncertainty in the
measurements is shown as an error bar.
~i iii r
1 1

500 h
IF S t e e l
o Radial
400 Hoop
A Axial

CD
s~
w
i-H
a
3
"2
cn
co

-300

4 6 8 10 12 14 16 18 20

Radial Position, mm

Figure 11.15: Measured radial, circumferential (hoop) and axial residual stresses as a
function of radial position in a 38.1 mm-dia. IF steel bar quenched in water flowing at
4.8 m s at 25 C. The uncertainty in the measurements is shown as an error bar.
_1
Residual Stress Measurement 260

600 ~i i i r
1
1 1

400
cd
CL,

200

CD

m
"3
!2 -200
tn
Hi

-400 h o Radial
Hoop
A Axial
-600

0 2 4 6 _i8_ 10 12 14 16 _i18_ 20

Radial P o s i t i o n , mm

Figure 11.16: Measured radial, circumferential (hoop) and axial residual stresses as a
function of radial position in a 38.1 mm-dia. 1045 carbon steel bar quenched in water
flowing at 2.8 m s at 50 C. The uncertainty in the measurements is shown as an error
_1

bar.

600
0 Radial
Hoop
400 A Axial
id
a.
200

0>
u
m
"3
-200
tn
cp
K
-400

Alloyed Steel
-600
_l i I i I i I i L

0 2 4 6 8 10 12 14 16 18 20

Radial P o s i t i o n , mm

Figure 11.17: Measured radial, circumferential (hoop) and axial residual stresses as a
function of radial position in a 38.1 mm-dia. alloyed steel bar quenched in water flowing
at 2.8 m s _1
at 75 C. Data collected during Campaign 2. The uncertainty in the
measurements is shown as an error bar.
Chapter 12

M a t h e m a t i c a l Analysis of Forced Convective Quenching

Results of the numerical (finite-element) simulations of microstructural evolution and


stress generation during forced convective quenching of steel bars are presented in this
chapter. As a first step in modelling the process, the applicability of the boiling curves
(surface heat flux vs surface temperature) estimated using the inverse analysis was as-
sessed by comparing predicted and measured temperature responses at the centre and
subsurface locations during the quenching of three IF steel bars. Then, the mathematical
models were applied to simulate the thermal, microstructural and mechanical responses
of selected quenched bars. As mentioned before, the thermal/mi crostructural model was
uncoupled from the mechanical module. Thus, the thermal and microstructural responses
for the whole time domain were first calculated, and the results adopted as input for the
stress model.

12.1 Verification of the Inverse Analysis

To verify the results of the inverse analysis (Chapter 5), estimated surface heat fluxes
were adopted as the active boundary condition in the finite-element code to calculate
the temperature response at the thermocouple locations. The predicted values were then
compared against measured temperatures. For this verification, three quenching condi-
tions applied to IF steel bars were considered (see Table 12.1). They fall in a diagonal of
the test matrix adopted in the forced convective quenching experiments (Chapter 7) and,
therefore, are representative of all boiling conditions found during the experimental runs.

261
Mathematical Analysis of Forced Convective Quenching 262

As was the case in the inverse analysis, the bars were assumed to behave as infinitely
long cylinders with respect to heat transfer. For the verification runs, the heat evolved
during the austenite-to-ferrite transformation was ignored. Thus, the governing equation
is given by Eq. (4.30). The computational domain represented one half of a single radial
slice (see Figure 12.1). The boundary conditions imposed were : 1) symmetry at the
centreline (Eq. (5.11)), and 2) specified heat flux at the surface (Eq. (5.12)). In addition,
surfaces perpendicular to the bar axis were considered adiabatic. A schematic repre-
sentation of the boundary conditions is shown in Figure 12.2. The initial temperature
distribution was assumed to be uniform (Eq. (5.13)). The mesh adopted for the calcu-
lations consisted of a row of 21 4-node isoparametric elements as shown, schematically,
in Figure 12.3. Note that smaller elements (in the radial direction) were used near the
surface, to improve accuracy in that region. The aspect ratio (height/width) of each
element varied from 1.33 at the centre to 4 at the surface. These values were within the
requirement that the aspect ratio of any given element should not exceed 6 [226].
During each simulation, the surface heat flux was interpolated from a table of es-
timated heat flux vs surface temperature values using a cubic spline algorithm [227].
This approach was selected because there is no single equation that can be used to fit
all regimes found in the boiling curves obtained in the experiments, and the size of the
problem and the efficiency of the interpolation algorithm did not compromise hardware
requirements. A disadvantage of this strategy is that parametric studies (see, for example,
Marjorek et al. [228]) are difficult to conduct.

The thermophysical properties of the IF steel given in Chapter 9 were adopted for
the calculations. The calculation time step, A t , was 0.2 s in all cases. A comparison
between the numerical ( F E M ) and measured temperature responses at the centreline
and subsurface positions during forced convective quenching of an IF steel bar with
water flowing at 4.8 m s _1
at 50 C is shown in Figure 12.4. The boiling curve adopted
in the simulation is shown as a broken line in Figure 12.5. During the early stages of
Mathematical Analysis of Forced Convective Quenching 263

the quench, the predicted temperature response lagged behind the measured values. It is
thought that this behaviour was associated with low values of estimated surface heat flux
at the beginning of the quench. To test this hypothesis, the boiling curve was modified
in such a way that a minimum value of surface heat flux could be attained immediately
(solid line in Figure 12.5). The results of the calculations with the modified boiling curve
are shown in Figure 12.6. As can be seen in the figure, the predictive capability of the
model improved significantly. It should be pointed out that a similar observation was
made by Wiskel [229] when simulating the temperature evolution during the start up
phase of aluminum D.C. casting. A n alternative solution to this problem would be to
acquire data at a much higher rate at the beginning of the quench and then to use smaller
time steps in the early stages of the finite-element simulation. Such an approach would
require a more flexible code to perform the inverse analysis, so that variable time steps
could be selected. Even in that start up value might be needed.

The temperature responses during forced convective quenching of IF steel bars with
water flowing at 1) 6.9 m s _1
at 25 C, and 2) 2.8 m s _1
at 75 C , were also computed
using the approach described above. The boiling curves adopted as boundary condition
are shown in Figures 9.27 and 9.25, respectively. For all three conditions simulated, no
phase transformation was considered (since only the thermal module was being tested).
The predicted and measured temperature responses at the centreline and subsurface
positions are shown in Figures 12.7 and 12.8, respectively. Good agreement between
predicted and measured temperature responses was also observed for these quenching
conditions. For the highest water temperature (75 C), the boiling curve included the
film boiling regime, which resulted in a low rate of heat extraction at temperatures
where the austenite-to-ferrite transformation occurs. Since the heat of transformation
was ignored in this set of calculations, predicted temperatures in the transformation range
were lower than the measured ones. As mentioned above, the three cases considered
covered the different boiling regimes found during the laboratory measurements. Thus,
Mathematical Analysis of Forced Convective Quenching 264

it can be concluded that the thermal module of the finite-element code, combined with
the results of the inverse analysis, correctly predicts the temperature response during
forced convective quenching.

The sensitivity of the inverse analysis to variables such as thermal conductivity, num-
ber of future time steps, and thermocouple position, was presented in Chapter 5. In the
following, the results of a sensitivity analysis of the thermal module alone, i.e., when no
transformations are included, are presented. When no phase transformations are con-
sidered, the evolution of the thermal field depends on the rate of heat extraction at the
surface and the ability of the material to transport energy. Accordingly, the variables
considered were thermal conductivity and surface heat flux. The case studied was that
of forced convective quenching of an IF steel bar with water flowing at 4.8 m s - 1
at
50 C . The effect of varying the value of the thermal conductivity by 10 % on the
thermal response at the centre and subsurface of the bar during the quench is shown in
Figure 12.9. When a higher value of thermal conductivity was adopted, a faster response
to changes in the boundary condition was predicted. The opposite effect was observed
when the thermal conductivity was lowered by 10 % with respect to its base value. Given
that these results were expected from the physics of the problem, they also serve as an
internal consistency check for the model.

To study the sensitivity of the thermal response to the magnitude of the surface heat
flux, the modified boiling curve shown in Figure 12.5 was allowed to vary by 10 % of its
base value . The calculated temperature response at the centre and subsurface positions
1

is shown in Figure 12.10. A n increase in surface heat flux resulted in faster cooling, while
the opposite effect was observed when a lower value of surface heat flux was adopted.
The shape of the temperature vs time curves was not affected, which was expected given
that the three boiling curves used in the calculations had the same characteristics.

1
Typical uncertainties quoted for boiling heat transfer correlations are in the order of 20 % [89].
Mathematical Analysis of Forced Convective Quenching 265

12.2 T e m p e r a t u r e Response and Microstructural Evolution

The results of the previous section indicate that the heat-transfer boundary condition
at the surface can be applied confidently to predict the temperature response during
forced convective quenching. It was then possible to apply this information to study the
microstructural evolution in the IF, 1045 carbon and alloyed eutectoid steel specimens
prepared for residual stress measurements and metallographic analysis (Chapter 7). To
model the temperature response and microstructural evolution during forced convective
quenching of the three steel bars, the same mesh and boundary conditions described in
the previous section, were used in the finite-element model. The governing equation must
include the heat generated during the various transformations (Eq. (4.3)), and boundary
conditions of symmetry at the centreline (Eq. (5.11)) and specified heat flux at the surface
(Eq. (5.12)). A uniform initial temperature distribution was assumed (Eq. (5.13)).

12.2.1 IF Steel

The microstructural evolution during forced convective quenching of a 38.1 mm-dia.


IF steel bar with water flowing at 4.8 m s _1
at 25 C was modelled using the finite-
element code. The thermophysical properties adopted for the calculations were given in
Chapter 9. The incremental time step, At, adopted in the simulation was 0.05 s (the total
quench time was 30 s). Due to the limited amount of available data, the kinetics of the
austenite-to-ferrite transformation were modelled via an explicit function of measured
ferrite fraction transformed as a function of temperature (see Appendix E). The latent
heat associated with the austenite-to-pearlite and the austenite-to-ferrite transformations
was estimated from the following correlations given by Campbell [188] :

A i J _ p = 70651 + 225.23 T - 0.3469 T + 6.755 x l 0 T ,


7
2 _ 5 3
T > 500 C (12.1)
Mathematical Analysis of Forced Convective Quenching 266

3.277xl0 - 10575.5 T + 11.545 T - 4.244xl0~ T , T > 780 C


6 2 3 3

AH 1-F - 2 . 9 1 7 5 x l 0 + 1.146xl0 T - 148.8 T + 0.064 T ,


7 5 2 3
720 < T < 780 C
2.216xl0 - 864.4 T + 1.979 T - 1.48xl0- T ,
5 2 3 3
T < 720 C
(12.2)
where A i 7 _ p and AH^-p are the latent heat (J k g ) of the austenite-to-pearlite and
7
- 1

austenite-to-ferrite transformations, respectively. Temperature-independent values of 92


and 82.6 kJ k g - 1
were adopted for the latent heat released by the austenite-to-bainite
and austenite-to-martensite transformations [47].
The predicted and measured thermal response at the centre and subsurface positions
2

during the quench are shown in Figures 12.11 and 12.12, respectively. Very good agree-
ment was observed at the subsurface position, except for temperatures below 150 C.
At the centre, the temperature response was underpredicted for temperatures below
~ 600 C , which correspond to measured subsurface temperatures below ~ 150 C . This
result suggests that the surface heat flux at the lower temperature end of the boiling
curve was overestimated.
The corresponding predicted and measured cooling rates at the centre and subsurface
are shown in Figures 12.13 and 12.14, respectively. As was the case with the predic-
tions of the temperature response, a better agreement was observed at the subsurface.
Another feature of interest is the evolution of the temperature gradients inside the bar
as the quench progresses, given that thermal strains arise due to non-uniform cooling.
Calculated temperature gradients at 5 different times during the quench are shown in
Figure 12.15. The surface temperature decreased very rapidly at the start of the quench.
After 8 seconds, a large temperature gradient between the surface and the centre had
been established. At later times, the temperature gradient decreased steadily to near
zero towards the end of the quench. Note that the rate of cooling was much faster at the
surface at the beginning of the quench, while the centre cooled more rapidly in the later

2
The measured temperature response shown in the plots corresponds to raw values.
Mathematical Analysis of Forced Convective Quenching 267

stages.
The predicted microstructural evolution at the centre and surface of the IF steel bar
is shown in Figures 12.16 and 12.17, respectively. As can be seen in the figures, the
austenite transformed completely to ferrite. The results of the metallographic analysis
(Chapter 7) also showed 100 % ferrite at all radial positions. Due to the slower cooling
rate at the centre, an appreciable difference in the start time for the transformation was
predicted between the surface and the centre. The surface started transforming at ~ 1 s
after the start of quenching, while the centre started to transform at 8.5 s.

12.2.2 1045 Carbon Steel

In the second example, the microstructural evolution during forced convective quenching
of a 38.1 mm-dia. 1045 carbon steel bar with water flowing at 2.8 m s _1
at 50 C
was simulated. The thermophysical properties of hypoeutectoid carbon steel [188] were
adopted for the calculations (Table 12.2). The time step, A t , adopted in the simulation
was 0.1 s (the total simulation time was 65 s). To model the transformation kinetics of
diffusional transformations, kinetic data reported by Nagasaka et al. [47] for the austenite-
to-ferrite, austenite-to-pearlite, and austenite-to-ferrite transformations were adopted.
The austenite-to-martensite transformation was modelled using the Koistinen-Marburger
equation (Eq. (2.6)). The values of heat of transformation given in Section 12.2.2 were
also adopted for this calculation.

The predicted and measured thermal response at the centre and mid-radius as the
quench progressed are shown in Figures 12.18 and 12.19, respectively. Fair agreement
between predicted and measured temperatures was observed, although the model under-
predicted the temperature response. Note that, at the centre, the recalescence associated
with the diffusional transformations was properly predicted.

The corresponding predicted and measured cooling rates at the centre and mid-radius
are shown in Figures 12.20 and 12.21, respectively. Fair agreement was observed in both
Mathematical Analysis of Forced Convective Quenching 268

cases. The decrease in cooling rate at the centre, just above 600 C, reflects the heat
generated by the diffusional transformations. Calculated temperature gradients at 5
different times during the quench are shown in Figure 12.22. Due to the internal thermal
resistance of the material, significant temperature gradients developed at early times.
Comparing Figures 12.15 and 12.22, it can be seen that the rate of change of the thermal
gradient was lower when a higher water temperature (50 C) was used for quenching.
This is a direct result of the lower rates of heat extraction at the surface obtained under
this quenching condition.

The metallographic analysis of the quenched bar showed a ring of martensite at the
surface (see Figure 9.39). In contrast, the predicted final microstructural distribution
showed a mixture of bainite and martensite at the surface. To tune the model to the
observed microstructure, the start times for the diffusional transformations were shifted
towards larger values, until a ring of martensite comparable in depth to that observed
in the quenched sample (Figure 9.39) was obtained. It was found that a shift of the
C C T 'nose' from 1.8 to 3.9 s was needed. A different austenite grain size and a slightly
different chemical composition could account for this difference. The evolution of the
microstructure at the centre and surface of the bar is given in Figures 12.23 and 12.24,
respectively. At the surface, martensite started to transform at 9.5 s, and the transfor-
mation was essentially complete after 35 s. Lower cooling rates at the centre allowed
transformation to ferrite, pearlite and bainite to start at 22, 23 and 25 s, respectively.
Martensite started to transform at 30 s. The calculated final microstructural distribution,
adopting the modified kinetics, is shown in Figure 12.25.

12.2.3 A l l o y e d E u t e c t o i d Steel

The microstructural evolution during forced convective quenching of a 38.1 mm-dia.


alloyed eutectoid steel bar with water flowing at 2.8 m s _1
at 75 C was also simulated
using the finite-element code. The thermophysical properties of plain carbon eutectoid
Mathematical Analysis of Forced Convective Quenching 269

carbon steel [188] (Table 12.3) were adopted for the calculations. The transformation
start times and J M A K (Eq. (2.4)) kinetic parameters derived in Chapter 10 were adopted
to model the kinetics of the diffusional transformations. The kinetics of the austenite-
to-martensite transformation were modelled using the Koistinen-Marburger equation,
Eq. (2.6). The heat of transformation for the reactions was estimated using the same
values adopted for the calculations of the IF and 1045 carbon steel specimens (see Section
12.2.1). A time step, At, of 1.0 s (for a total quench time of 135 s) was adopted in the
simulation.

The predicted and measured thermal responses at the centre and mid-radius during
the quench are shown in Figures 12.26 and 12.27, respectively. Good agreement was
observed. However, the recalescence associated with the austenite-to-martensite trans-
formation was overpredicted by the model.
The corresponding predicted and measured cooling rates at the centre and mid-radius
of the specimen are shown in Figures 12.28 and 12.29, respectively. Fair agreement
was observed in both cases, except in the range of 600 to 700 C, where the model
overpredicted the cooling rates. Given that the bar transformed fully to martensite, this
result indicates that the surface heat flux in this temperature range was overestimated.
The heat generated by the transformation resulted in a decrease in the cooling rate at
~ 220 C. Calculated temperature gradients at 5 different times during the quench are
shown in Figure 12.30. Since the quench was conducted with water at 75 C , the rate of
change of the temperature profile was the slowest of the three cases considered.

The predicted microstructural evolution at the centre and surface of the eutectoid al-
loyed steel bar is shown in Figures 12.31 and 12.32, respectively. At the surface, marten-
site started to transform after 25 s, and the transformation was essentially complete
at 60 s. The lower cooling rates experienced at the centre allowed transformation to
a mixture of bainite and martensite. Bainite started to transform at 40 s, followed by
martensite at 62 s. The final microstructural distribution is shown in Figure 12.33. The
Mathematical Analysis of Forced Convective Quenching 270

model predicted an essentially through-hardened material, which is consistent with the


metallographic analysis and hardness measurements.

12.2.4 Sensitivity Analysis

There are two groups of variables that affect the predictions obtained from the ther-
mal/microstructural model. They are related to 1) heat transfer from the specimen,
and 2) transformation kinetics. The first group includes such variables as rate of heat
extraction, thermophysical properties and specimen geometry. The effect of varying the
surface heat flux and thermal conductivity of the bar on the temperature response during
forced convective quenching of IF steel bars (when transformation is not considered) has
been presented above. To predict transformation kinetics, empirical equations have been
used to estimate parameters such as continuous cooling start times and kinetic constants
for the J M A K equation, Eq. (2.4). Another variable that plays an important role on
the thermal/mi crostructural predictions is the heat released during the transformation.
Campbell [188] has studied the effect of transformation kinetics-related parameters on
the predicted microstructural evolution during the austenite-to-ferrite and austenite-to-
pearlite transformations occurring during air cooling of hypo- and eutectoid carbon steel
rods. His results indicated that variations in the transformation start times, within the
expected experimental error, have an small effect on the calculated microstructural evo-
lution. A n increase in either one of the J M A K kinetic parameters (n and b) resulted
in enhanced kinetics, which in turn led to a more rapid release of heat of transforma-
tion and, consequently, higher average transformation temperatures. Similar results have
been reported by Medina [230] for the cooling of a 1005 carbon steel in a run-out table.
To avoid repetition, the sensitivity of the present model to variations of kinetic-related
variables for the austenite-to-martensite transformation only, were considered.
Mathematical Analysis of Forced Convective Quenching 271

The progress of the austenite-to-martensite transformation is a function of tempera-


ture below M and is described by the Koistinen-Marburger equation , Eq. (2.6). Thus,
s
3

the kinetic variables of interest are the transformation start temperature ( M J and the
kinetic constant (A = -0.011). Stress is known to increase the kinetics of the austenite-to
martensite reaction by increasing the transformation start temperature. Accordingly, the
effect of varying the magnitude of M was studied by simulating the temperature and mi-
s

crostructural evolution during the quench for values of M s : 1) 20 C and 2) 40 C above


its measured value under stress-free conditions. The predicted temperature response at
the centre and mid-radius are presented in Figure 12.34.

The corresponding microstructural evolution is shown in Figure 12.35. B y increas-


ing M , the reaction started at shorter times (higher temperatures). Referring to Fig-
ures 12.28 and 12.29, it can be observed that this resulted in the reaction taking place
under higher cooling rates with respect to the predicted conditions when the base value
of M was adopted. Thus, not only the reaction started earlier but also enhanced kinetics
g

at both the surface and the centre were predicted. Note that under conditions of con-
stant cooling rate, increasing M would have resulted in only a shift of the transformation
s

curves towards shorter times but without modifying their shape.


The effect of varying the magnitude of the kinetic constant in the Koistinen-Marburger
equation on the predicted temperature response and microstructural evolution was in-
vestigated by allowing this parameter to take two values that simulated enhanced kinet-
ics : 1) -0.015 and 2) -0.020. The calculated temperature response at the centre and
mid-radius, and the corresponding microstructural evolution, are shown in Figures 12.36
and 12.37, respectively. Note that the temperature axis in Figure 12.36 has been bro-
ken to improve readability. B y increasing the kinetic constant, the fraction transformed
for a given degree of supercooling, A T , increased. The martensitic transformation is
time-independent. However, the temperature at a given location is a function of the

3
E q . (2.6) is : X = 1 - exp[A(M - T)].
s
Mathematical Analysis of Forced Convective Quenching 272

local cooling rate and, therefore, there is a direct relationship between supercooling and
time. The enhanced kinetics resulted in a larger amount of heat released during the
transformation, delaying the start of the transformation at the centre. However, the rate
of transformation corresponding to the values of -0.015 and -0.020 was increased to such
extent that those reactions were completed at the centre at shorter times than the one
predicted using the base value of -0.011.

12.3 Stress Generation

The results of the previous section were adopted as input to the mechanical model to
predict transient and residual stresses during quenching of the three specimens studied.
A row of 4-node elements, like the one used in the previous section, predicts a null axial
stress at all radial positions. Thus, a new mesh was created to simulate one-half of the
entire bar , as shown in Figure 12.38. The mesh consisted of 21 elements in the radial
4

direction (with the same geometry adopted for the thermal/microstructural calculations)
and 34 elements in the axial direction. A schematic representation of the computational
domain and the thermal boundary conditions applied is shown in Figures 12.39 and 12.40,
respectively. Note that the ends of the specimen were assumed to be adiabatic. This
is a reasonable assumption, considering the very high rates of heat extraction experi-
enced at the surface. The mechanical boundary conditions are schematically shown in
Figure 12.41. At the centreline, the symmetry of the specimen dictates that the radial
displacement must be set to zero. The front end of the bar was assumed to be free to
move in the axial direction. Due to the symmetry of the problem, the only possible
rigid-motion was translation in the axial direction. The boundary condition at the back
end of the specimen reflects the fact that, as described in Chapter 7, the bar was rigidly
held in position by attaching it to a steel tube through an adaptor, restricting both axial

4
The symmetry of the problem allows the use of only one half of the specimen.
Mathematical Analysis of Forced Convective Quenching 273

and radial displacement at two points in the back end of the specimen.
The thermal/microstructural calculations were repeated with the new mesh, and the
results adopted as input to the mechanical model. It should be pointed out that the
thermal response and microstructural evolution at mid-length were identical to those
obtained with the 21 node mesh. In the following, the computed residual stresses are
compared to measured values, to assess the predictive capabilities of the model. Then,
the mechanisms of stress generation during quenching are illustrated by discussing the
evolution of transient stresses. In both cases, only the results at mid-length are presented.
Moreover, following a standard practice [231], average stresses were computed at the
geometrical centre of the individual elements.

12.3.1 Residual Stresses

12.3.1.1 I F Steel

The stress generation during forced convective quenching of a 38.1 mm-dia. IF steel
bar with water flowing at 4.8 m s at 25 C was simulated with the mechanical module
_1

of the mathematical model. The thermomechanical properties of pure iron given by


Mitter et al. [205] were adopted for the calculations, and are given in Table 12.4. The
transformation plasticity strain increment was calculated using Eq. (D.4). The At used
in the simulation was 0.05 s.

The radial, circumferential and axial components of the predicted and measured resid-
ual stress distributions are shown in Figures 12.42, 12.43 and 12.44, respectively. Good
agreement between predicted and measured residual stress distributions was observed.
Near the surface, the model overpredicted (in absolute terms) both the axial and circum-
ferential stresses. Two conditions dictated by the equilibrium equations are : 1) the radial
residual stress at the surface must be zero, and 2) at the centreline, the circumferential
and radial stresses must coincide. This conditions can be used as a consistency check for
Mathematical Analysis of Forced Convective Quenching 274

the model. As seen in Figures 12.42 and 12.43, both conditions were correctly predicted
by the model. Regarding the predicted residual stress distributions, compressive axial
and circumferential stresses were obtained at the surface while tensile stresses were com-
puted at the centre. The radial residual stress distribution was zero at the surface and
rose steadily to its maximum value at the centre.

12.3.1.2 1045 C a r b o n Steel

The model was also applied to simulate the stress generation during forced convective
quenching of a 38.1 mm-dia. 1045 carbon steel bar with water flowing at 2.8 m s a t _1

50 C . The thermomechanical properties of 1035 carbon steel given by Nagasaka et al.


[47] were adopted for the calculations, and are given in Table 12.5. The poisson ratio, u,
was assumed as constant and equal to 0.3 for all phases. The transformation plasticity
strain increment was calculated using Eq. (D.4). The At used in the simulation was 0.1
s. The microstructural evolution computed with the modified kinetics was adopted as
input for the calculations.
The radial, circumferential and axial components of the predicted and measured resid-
ual stress distributions are shown in Figures 12.45, 12.46 and 12.47, respectively. The
predicted residual stress distributions agree qualitatively with the measured ones. In
particular, very good agreement was observed regarding the circumferential component.
The radial component was underpredicted near the surface and overpredicted towards
the centre. The axial component was underpredicted in the central region. As was the
case with the previous simulation, the conditions given by the equilibrium equations were
fulfilled by the model (see Figures 12.45 and 12.46).

The residual stress distributions were similar to the ones obtained during the simula-
tion of quenching of the IF steel, i.e., the axial and circumferential stresses at the surface
were compressive while tensile stresses were predicted at the centre. The radial residual
stress distribution increased from zero at the surface, reaching its maximum value at the
Mathematical Analysis of Forced Convective Quenching 275

centre.

12.3.1.3 A l l o y e d E u t e c t o i d Steel

The residual stress distribution developed after forced convective quenching of a 38.1 mm-
dia. alloyed eutectoid steel bar with water flowing at 2.8 m s a t 75 C was predicted
_1

using the mathematical models. There is no data published on the thermomechanical


properties of this steel. Instead, the properties of an alloyed steel (3.23 pet Ni-0.83 pet
Cr) given by Nagasaka et al. [47] were adopted for the simulation. It should be noted
that Henriksen et al. [232] have pointed out that the thermomechanical properties of
steels are more sensitive to temperature and the specific phase present than to carbon
content. The transformation plasticity strain increment was calculated using Eq. (D.4).
A time step, A i , of 0.05 s was used in the calculations. The radial, circumferential
and axial components of the predicted and measured residual stress distributions are
shown in Figures 12.48, 12.49 and 12.50, respectively. The measured residual stress
distributions showed a marked difference when compared to the distributions observed in
the quenched IF and 1045 carbon steel specimens. As can be seen in the figures, the model
correctly predicted these differences, at least qualitatively. Thus, tensile circumferential
and axial stresses were predicted at the surface, while compressive values were obtained
at the centre, as a result of the through-hardening of the laboratory test sample. The
predicted radial component of the residual stress was zero at the surface, as required by
the equations of equilibrium, and became increasingly compressive towards the centre.

12.3.1.4 Sensitivity Analysis

Given that the predicted residual stress distributions are based on the thermal and mi-
crostructural evolution in the specimen, which are interdependent, and a function of the
rate of heat extraction at the surface and the kinetics of the phase transformations, it
Mathematical Analysis of Forced Convective Quenching 276

is apparent that a sensitivity analysis based on a factorial design including all variables
would result in a very large number of combinations of parameters. Instead, by ana-
lyzing the results presented above, it appears that the most significant factor affecting
the predicted residual stresses is the microstructural distribution. In particular, the mi-
crostructural evolution in the partially hardened 1045 carbon steel quenched bar is the
most likely to be miscalculated. Thus, to test the sensitivity of the stress module to
microstructural evolution, the residual stress distribution in a 1045 carbon steel bar with
the final microstructural distribution predicted based on the kinetic data reported by
Nagasaka et al. [47] was computed. The results showed that the model underpredicted
the residual stresses. This is a consequence of the smaller amount of martensite predicted
by the thermal/microstructural model when the original transformation kinetic data was
adopted. The largest volume expansion corresponds to the austenite-to-martensite trans-
formation and, consequently, when less martensite was formed, the model predicted lower
residual stresses. When the transformation behaviour was modified to account for the
microstructure obtained in the quenched sample, much better agreement between the
calculated and the measured residual stresses was obtained, as shown in Figures 12.45,
12.46 and 12.47.

12.3.2 Transient Stresses

In order to elucidate the role of thermal evolution and phase transformations on stress
generation during forced convective quenching, the mechanical response of the bars as the
quench progressed was monitored. From the neutron diffraction results, it was evident
that the stress field evolved in a similar manner in the unhardened IF and the partially
hardened 1045 carbon steel bars, while the behaviour of the through hardened alloyed
eutectoid steel was different. Thus, only the stress evolution in the 1045 carbon and the
alloyed eutectoid steel were considered.

The variation with time of the radial, circumferential and axial components of stress
Mathematical Analysis of Forced Convective Quenching 277

at the surface and centre of the alloyed eutectoid steel quenched bar are shown in Fig-
ures 12.51 and 12.52, respectively. Note that, as required by the equilibrium equations,
the radial and circumferential components at the centre were identical at all times, and
the radial component at the surface remained constant and is approximately zero. Also,
the stresses did not change appreciably towards the end of the quench, indicating that
they represent the residual stress field. At the start of the quench (Figure 12.51), the
surface cooled more rapidly than the centre (see Figure 12.30), and the accompany-
ing contraction set up tensile stresses which increased steadily from the start of the
quench up to point A . At point A (30 s), the surface transformed to martensite (see
Figure 12.32). The expansion associated with the transformation overcame the contrac-
tion due to cooling, resulting in a reversal in the stress vs time curve between points A
and B . The martensitic transformation was essentially completed at point B ; from then
on, the contraction due to cooling dominated until the end of the quench, resulting in
another reversal in the sign of the rate of change of stress. At the start of the quench,
the centre (Figure 12.52) did not cool significantly and, therefore, the contraction at the
surface was accommodated by the softer core, resulting in compressive stresses until the
martensitic transformation started (point A ) . At that point, the stresses were reversed
and, eventually, became tensile up to the time when martensite started to transform
at the centre (point C). The expansion due to the transformation caused another stress
reversal, that resulted in compressive stresses. The subsequent cooling did not modify
the stress values significantly. Note that the small amount of bainite transformed at the
centre (which started to transform at 40 s), had a minimal effect on the stress evolution.

The evolution of the radial, circumferential and axial components of stress at the
surface and centre of the 1045 carbon steel quenched bar are shown in Figures 12.54
and 12.53, respectively. At early times, the rapid cooling experienced by the surface
layers generated tensile stresses at the surface (Figure 12.53), until martensite started
to transform (point A ) . The volume change produced by the martensitic transformation
Mathematical Analysis of Forced Convective Quenching 278

was larger than the contraction caused by cooling and, therefore, the stress increment
was negative until the transformation was complete (point B in Figure 12.53). At times
greater than thermal contraction, generated a positive stress increment, until the cen-
tre started to transform. The accompanying expansion caused another stress unloading
at the surface. A t the centre, the stresses evolved along a different path. The shrinkage
experienced at the surface at the beginning of the quench produced compressive stresses
at the softer core. When the surface transformed, the stress at the centre was reversed
(point A in Figure 12.54). The positive rate of change of stress at the centre continued as
the transformation front travelled through the radius of the bar, until the centre started
to transform to ferrite (point C). The diffusional transformations (austenite-to-ferrite,-
pearlite, and -bainite) occurred between points C and D. The expansion associated with
these transformations resulted in a stress unloading at point C. Once the transformations
were completed, the rapid cooling at the centre (relatively to the surface) reversed this
trend, until the centre started to transform to martensite (point E), causing another
stress unloading. The martensitic transformation was completed at point F. From this
time onward, stresses were only generated by the relatively rapid cooling of the core.

12.4 A p p l i c a t i o n to Industrial Conditions

The thermal-microstructural module of the mathematical model was also applied to sim-
ulate the microstructural evolution in a 100 mm-dia alloyed eutectoid steel bar. The
surface heat flux was derived from measurements obtained in the laboratory facility for a
38.1 mm-dia. alloyed eutectoid steel bar quenched from 850 C with water flowing at 4.8
m s _ 1
at 32 C . The bar was treated as infinitely long; thus, the computational domain
adopted was one half of a radial slice, as in the verification calculations (Figure 12.1).
The domain was discretized with a 21 4-node element mesh. The calculational time step,
At was 0.5 s. Thermophysical and kinetic data used in the calculations for the 38.1
Mathematical Analysis of Forced Convective Quenching 279

mm-dia. alloyed eutectoid steel bar (Section 12.2.3) were adopted for the simulation.
The calculated final microstructural distribution is shown in Figure 12.55. A mixture
of pearlite and bainite was predicted in the core while 100 % martensite was predicted
at the surface. It should be pointed out that no measurements were conducted under
industrial conditions and, therefore, these results should only be considered as qualita-
tive. Additionally, there is an uncertainty associated with applying the surface heat flux
obtained in the laboratory to the industrial set up. Nonetheless, the microstructural
profile shown in Figure 12.55 is consistent with the hardness profile given in Figure 9.42.
Mathematical Analysis of Forced Convective Quenching 280

Table 12.1: Quenching conditions simulated in the verification runs of the thermal model.

Run Material Water Velocity, m s 1


Water Temperature, C
R U N 11 IF Steel 6.9 25
R U N 10 IF Steel 4.8 50
R U N 16 IF Steel 2.8 75
Mathematical Analysis of Forced Convective Quenching 281

o o \> \S \> \> V O


O o h*H h^H HH h*H H^H o o
.9 .9 o .9 .9 .9 .9 .9 .9 .9
o EN EN .9 EH E-H EH Ei EH EN EN
EN

bO
O O O O O O O
a o o o o O
o o
o o
o o o
o O
o o o o
o
o
o c- CM CO oo o
o o o
o o o o CM o o o
CD
o o o o r-- o o o
!-i o L O
o o VI VI VI oo
oo c$ VI VI VI VI VI VI t i-H
VI
11
VI
11
VI
EN EN EN
00 u
03 EN A
EN EN EN EN EN
V V V EN EN EN

VI VI VI VI VI VI EN
VI VI VI
03 a o
t
CM
r
(M
CT5
CO
o o o o o L O
o o o

o
S-.
c<3
O
I
CN
L O o
EN
o CD X
EN
0I

EN
1
1

o
o i
TI
X I1
X
o O
co
co EN 1I 03 c~
Pi CM
X +3
CO EN X C O

03 I L O I t~ CO
EN L O
11
CM EN
(-1
CO O CO co cS
CM
EN L O
o 03
CT
+ co L6 PL. i i

1

03 iI O in 11
W O + + o
o a X
r-
CM CN CM o
L O

1
+ i-H EN
.2 o
(-1
o o co EN o o
OH O
It!
+
L O L O
X CN L O

'co CO 1I
L O
CO oo | CO -1-
O
'cQ CO EN X CM EN co o t 1
O CO L O
co
>> 03 r- oo OS
L O

^3 SH co
CO
r TI co CM
CO
11
X oi
bO
L O
OH CM L O
L6
.i XT'
03
L O

o tf
CM r
+
r>
co
+
L O I

L6 EN CO L O
L O
o
oo L O
co
C O CM CO C O
CM CM o o o
co o CM
CM
L O
CO t>-
o
co
ii
L O

C O
o
L O

I 7 7
(-1
03 7 rH CO
1
i
bO bO
kg'
o

OH
O
s-.
AS
a
PU bO bO bO
>>
tf tf -si" c
Mathematical Analysis of Forced Convective Quenching 282

fl O O
CD
o o o o
.9 fl fl fl .9
.9 .9 .9 .9
E~H E^ E^ E^ E^
O
E-H E-H

bO O O O O
PI O
o o o o o o o o o o
r~ t~
tf o
o
o
o
o
o
o
o
o
o o CM
05
CO oo
r-
o o
o
CP o o o o o c- t o o
1I II o o o
fl ,I 11
CM VI VI VI r
VI VI VI
L O
oo VI
VI VI VI E~H E^ E"H t-
JH

cu E-i E*n E*n E-H


E-H
E~H
V V V
A E^
VI E-H
VI
OH
VI VI VI VI
VI r~ t~
CO
o o o o CM CM o
o o L O

E-H

E-H
X
i I
0
E-H E-H
o
iI 05 a>
1I

+
X o O X
C
N ii
E-. <D C O CD
t~ 1I

fl co X EH X C O
L O
I EH
CD O E^ I>- CO CM a3
fl ai >o co
co co r 10
o co ai CD
iI
W < o L O
fl + ,i OH

fl
X
CO C
N o
II
OS
iI
+ +
1
+
oo o
.2 o
o X C O E-H
E-H

C
O +
L O L O L O
'co co 1I
CO CM I
oo I
EH
X o CO
CP t l>- oo CO
L O
CO o
i CO co
faD
CD co co
t
II
L O L O
CM
CO L O ] X
TI
t
L O

tf CO
co CM
L O + co +
L O
CO
L O

o
oo L O
CM CO co
CM o o
co o o CM
L O
CO
o
L O

co
o
LO

I
I

u
V I
bO bO
OH
Al AS
o bO bO
l-l A!
OH
tf 1
Mathematical Analysis of Forced Convective Quenching 283

CO

fl O O O
CD o o o

O E-H E~H EH

bO O
fl oO o o
cc3 o o
tf o o O
o
o
o
o
o o
L O
o <D o 1I 1I o
M 1I co
C M
fl o
VI C M VI o
VI C O

0) E-H
CD
SH VI V
CD EH
CO PH VI E-H VI EH
VI
o o
o o
C M co

CO
_CD EH CN CN
HJ E-H EH
(H
CD <N
PH EH 1 1
1 I
O 0 0
fH x I 11
PH fl C O o 11
X X
o .2 C M
TI
11
X
L O
O
L O
O
"3 00 oq EH
fl C O

C M
C O
C O C O

w + + +
C O
o
CD I
co
fl o E-H EH
a o EH

o
o E-H
C O I 00
o 06
II
o co
L O

SH CD
L O
0 ai
CD *-< 1 o 11 11
bO
<D
tf LO
I
-tf
1
+
cn
0 O
^ 11 11
C M X C M
X X
C M L O
C M
01
O O
11 co
o C O

CM

?H
CD
PH n3 ct3
ce PH
O PH
PH
PH
3>
b
Mathematical Analysis of Forced Convective Quenching 284

GO
Pi o O O o o O o O O
Conrime

o o o o o o o o o o
.3 .5 .5 .9 .9 .9 .3 .3 .3 .3
EN EN EN EN EN EN EN EN EN EN

o o
ige

o O O o o O O O
oo
O
ci o o o o o o o o o o
tf o o o o o o o O o o o
o o o o
o o o o o o o o o
m o o o o o o o o
11 11
11 11
O
eratu

SH VI o VI
11 VI
VI VI VI
11 V
11I
VI LO V
11I VI
11 V
11I
EN V EN EN EN V
EN EN EN EN EN EN EN
VI VI
dm

VI EN VI EN
VI VI VI VI VI VI VI
o CD LO o
o o o co o o o o o o
:arli

CD LO CD
CD
CD

rtensi
steni

H^3
CD EN /EN EN
EN EN EN
EN PH EN '3
d
t OJ EN
t~ CD"
t>- oq EN

tLO "ci <D EN EN CD OJ EN
t-
oj i LO OJ oo H ^

co ii
,

CT* OJ CM PQ LO
CO LO CM CO 'in OJ CM CO oo t
<
ii
co co
^
W 'EH
OJ OJ oo SH CM
11 CM LO ]>- SH
<V
CO OJ
s-l
OJ
t-
ii

a 1 1 T1

HH
1 1 CM <H-H LO CO 1 11
1
o in
i i
o 1 in
I
ic
I;
O
m
1 CO o o CO
ci in
1
o in
i
' cn
O o 1 CO
O O I t, 1 1 I O
U
CD
CD iI T1
CM
()
iI iI <D CD o LO
i1
X X V X X V 3 o a 11
X X
7.85:

7.85:

bfj
CD oo CM cd oo co ci CD CM cc3
o t--
tf CO CO
OJ
-OJ CD CM
1>- r-
CM
CO CD OJ CO 11
11
OJ
co o
T1 .

CM CM CM ii
CM i-H

>>
Properl

, MPa

, MPa

, MPa
, MPa
MPa

MPa

MPa

MPa
MPa

MPa
MPa
MPa

b fe b b b
Mathematical Analysis of Forced Convective Quenching 285

191 mm

Figure 12.1: Computational domain adopted for the finite-element simulation of ther-
mal/microstructural evolution during forced convective quenching.

Insulated

Insulated

Figure 12.2: Schematic representation of the boundary conditions adopted for the ther-
mal/microstructural simulations.
Mathematical Analysis of Forced Convective Quenching

rjjJ
H UIUI c'o
Mathematical Analysis of Forced Convective Quenching 287
T

Time, s

Figure 12.4: Comparison between predicted and measured temperature responses at the
centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8
m s at 50 C. The boiling curve shown as a broken line in Figure 12.5 was adopted as
_1

boundary condition.

Figure 12.5: Boiling curves adopted as boundary condition to simulate forced convective
quenching of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m s at_1

50 C .
Mathematical Analysis of Forced Convective Quenching

Time, s

Figure 12.6: Comparison between predicted and measured temperature responses at the
centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8
m s at 50 C. The boiling curve shown as a solid line in Figure 12.5 was adopted as
- 1

boundary condition.

0 5 10 15 20 25 30

Time, s

Figure 12.7: Comparison between predicted and measured temperature responses at the
centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with water flowing at 6.9
m s- at 25 C.
1
Mathematical Analysis of Forced Convective Quenching 289

Time, s

Figure 12.8: Comparison between predicted and measured temperature responses at the
centre and subsurface of a 38.1 mm-dia. IF steel bar quenched with water flowing at 2.8
m s" at 75 C.
1

Time, s

Figure 12.9: Effect of varying the value of thermal conductivity by 10 % adopted in


the finite-element simulation on the temperature response at the centre and subsurface
of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m s at 50 C .
_ 1
Mathematical Analysis of Forced Convective Quenching 290

Time, s

Figure 12.10: Effect of varying the surface heat flux by 10 % adopted in the fi-
nite-element simulation on the temperature response at the centre and subsurface of a
38.1 mm-dia. IF steel bar quenched with water-flowing at 4.8 m s at 50 C .
_1

0 10 20 30 40

Time, s

Figure 12.11: Comparison between predicted and measured temperature responses at the
centre of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m s at 25 C.
- 1
Mathematical Analysis of Forced Convective Quenching 291

1000
O Measured
FEM

800 h

u
3 600
cd
cu

400
IF Steel

Subsurface
200

40

Figure 12.12: Comparison between predicted and measured temperature responses at the
subsurface of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m s at _ 1

25 C .

40

20 Measured
FEM
0

-20

o -40
co"
+J
cd
-60

-80
a
"3 -100
o
o
-120

-140 IF Steel
Centre
-160

200 400 600 800 1000


o
Temperature, C

Figure 12.13: Comparison between predicted and measured cooling rates at the centre
of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m s at 25 C. _1
Mathematical Analysis of Forced Convective Quenching 292

20 h

-180 \-

0 200 400 600 800 1000


o
Temperature, C

Figure 12.14: Comparison between predicted and measured cooling rates at the sub-
surface of a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m s at _1

25 C.
1 1 1
1 1

t = 0 s
1000

t = 8s
-

3 600 -
12 s

t = 16 s
200

t = 29 s *
1 , 1 , 1 ,
0
0 5 10 15

Radial Position, mm

Figure 12.15: Calculated temperature gradients at five different times during forced
convective quenching of a 38.1 mm-dia. IF steel with water flowing at 4.8 m s at _ 1

25 C .
Mathematical Analysis of Forced Convective Quenching 293

1.0

Ferrite
Pearlite
0.8 Bainite
T3
Martensite

0.6
d

a 0.4
o

o
6-
0.2 IF Steel
Centre

0.0

10 20 30

Time, s

Figure 12.16: Calculated microstructural evolution at the centre during forced convective
quenching of a 38.1 mm-dia. IF steel bar with water flowing at 4.8 m s at 25 C . _ 1

1.0

Ferrite
Pearlite
0.8 Bainite
CP Martensite

0.6
d
u
a 0.4
o
'
o
c
S-,

0.2 IF S t e e l
Surface

0.0
_i_

6 8 10 12 14

Time, s

Figure 12.17: Calculated microstructural evolution at the surface during forced convective
quenching of a 38.1 mm-dia. IF steel bar with water flowing at 4.8 m s at 25 C . _ 1
Mathematical Analysis of Forced Convective Quenching 294

1000
o Measured
FEM
800

600
a)
u
CD
ft

400

200

Figure 12.18: Comparison between predicted and measured temperature responses at the
centre of a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at 2.8 m s
_1

at 50 C.

~i ' r
1000
o Measured
FEM

800

0
u
3 600
CC
u
CB
CM

S 400
C
O
E-i

200 1045 Steel


MidRadius
<>-<"> O O P o

50 60 70

Figure 12.19: Comparison between predicted and measured temperature responses at


mid-radius of a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at 2.8
m s - at 50 C.
1
Mathematical Analysis of Forced Convective Quenching 295

40

SO Measured
FEM

-20

-40 h
0
O-'
o
o -60

1045 Steel
Centre
-80 h

-100
200 400 600 800 1000
o
Temperature, C

Figure 12.20: Comparison between predicted and measured cooling rates at the centre
of a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at 2.8 m s at _ 1

50 C .
40

20

-20

-40
a
a
O -60

1045 Steel
-80 Mid-Radius

-100
200 400 600 800 1000
o
Temperature, C

Figure 12.21: Comparison between predicted and measured cooling rates at mid-radius
of a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at 2.8 m s at - 1

50 C .
Mathematical Analysis of Forced Convective Quenching 296

i
1 1
1

t = 0 s
1000

t = 10 s

800

600 -

CO 20 s

s
P.
30 s
CO 400 -
E-

200 -

t = 65 s """--...^

1 , 1 , 1 ,

5 10 15

Radial Position, mm

Figure 12.22: Calculated temperature gradients at five different times during forced
convective quenching of a 38.1 mm-dia. 1045 carbon steel bar with water flowing at 2.8
ms- at50C.
1

1.0
1045 Steel
Ferrite
Centre
Pearlite
0.8 h Bainite
Martensite

0.6

d
o 0.4
o
CO

0.2

0.0
10 20 30 40 50 60 70

Time, s

Figure 12.23: Calculated (with modified kinetics) microstructural evolution at the centre
during forced convective quenching of a 38.1 mm-dia. 1045 carbon steel bar with water
flowing at 2.8 m s" at 50 C.
1
Mathematical Analysis of Forced Convective Quenching 297

1.0

Ferrite
0.8
Pearlite
Bainite
Martensite
0.6
a
a
u
H
a 0.4 h
o
rH
o

0.2 h
1045 Steel
Surface

0.0

10 20 30 40 50 60 70

Time, s

Figure 12.24: Calculated (with modified kinetics) microstructural evolution at the surface
during forced convective quenching of a 38.1 mm-dia. 1045 carbon steel bar with water
flowing at 4.8 m s at 50 C.
- 1

Radial Position, mm

Figure 12.25: Calculated (with modified kinetics) final microstructural distribution in a


38.1 mm-dia. 1045 carbon steel bar quenched with water flowing at 4.8 m s at 50 C. _1

Compare with Figure 9.39.


Mathematical Analysis of Forced Convective Quenching 298

Figure 12.26: Comparison between predicted and measured temperature responses at the
centre of a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water flowing at 2.8
m s- at 75 C.
1

800
O Measured
FEM

600

400
n
CO
H

200

_1_
20 40 60 80 100 120 140

Time, s

Figure 12.27: Comparison between predicted and measured temperature responses at the
subsurface of a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water flowing at
2.8 m s- at 75 C .
1
Mathematical Analysis of Forced Convective Quenching 299

40

Measured
20 FEM

cd
K
bo
a -20 h
"
3
o
o

-40 Alloyed Steel


Centre

-60
200 400 600 800
o
Temperature, C

Figure 12.28: Comparison between predicted and measured cooling rates at the centre
of a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water flowing at 2.8 m s _1

at 75 C .
40

Measured
20 \- FEM

cd

be
el -20

-40
Alloyed Steel
Mid-Radius

-60
200 400 600 800
o
Temperature, C

Figure 12.29: Comparison between predicted and measured cooling rates at the subsur-
face of a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water flowing at 2.8 m
s" at 75 C.
1
Mathematical Analysis of Forced Convective Quenching 300

t = 0 s

800

t = 25 s

600

... t = 35 s
400

t = 65 s
200

t = 135 s

5 10 15

Radial Position, m m

Figure 12.30: Calculated temperature gradients at five different times during forced
convective quenching of a 38.1 mm-dia. alloyed eutectoid steel bar with water flowing at
2.8 m s- at 75 C .
1

1.0 h
Alloyed Steel
Ferrite
Centre
Pearlite
0.8 Bainite
Martensite

% 0.6
a

0.4

0.2

0.0
_l_
20 40 60 80 100 120 140

Time, s

Figure 12.31: Calculated microstructural evolution at the centre during forced convective
quenching of a 38.1 mm-dia. alloyed eutectoid steel bar with water flowing at 2.8 m s _1

at 75 C.
Mathematical Analysis of Forced Convective Quenching 301

1.0 r-
Alloyed Steel
Surface

0.8 h
CD

B 0.6 h
Ferrite
cd

H
1H Pearlite
Bainite
a 0.4
o Martensite
a
cd
u
0.2

0.0
_i_
20 40 60 80 100 120 140

Time, s

Figure 12.32: Calculated microstructural evolution at the surface during forced convective
quenching of a 38.1 mm-dia. alloyed eutectoid steel bar with water flowing at 2.8 m s _1

at 75 C.

5 10 15

Radial Position, mm

Figure 12.33: Calculated final microstructural distribution in a 38.1 mm-dia. alloyed


eutectoid steel bar quenched with water flowing at 2.8 m s at 75 C. Compare with _1

Figure 9.38.
Mathematical Analysis of Forced Convective Quenching 302

Figure 12.34: Effect of varying the value of M adopted in the finite-element simulation
s

on the temperature response at the centre and subsurface of a 38.1 mm-dia. alloyed
eutectoid steel bar quenched with water flowing at 2.8 m s at 75 C .
- 1

Figure 12.35: Effect of varying the value of M adopted in the finite-element simulation
s

on the microstructural evolution at the centre and subsurface of a 38.1 mm-dia. alloyed
eutectoid steel bar quenched with water flowing at 2.8 m s at 75 C .
_1
Mathematical Analysis of Forced Convective Quenching 303

900

A = - 0.011
BOO A = - 0.015
A = - 0.020
700

300
Centre
CU

s
cu 200 MidRadius
E-i

100
Alloyed Steel

20 40 60 80 100 120 140

Time, s

Figure 12.36: Effect of varying the value of the kinetic constant in the Koisti-
nen-Marburger equation adopted in the finite-element simulation on the temperature
response at the centre and subsurface of a 38.1 mm-dia. alloyed eutectoid steel bar
quenched with water flowing at 2.8 m s at 75 C ._ 1

1.0 Alloyed Steel

XI 0.8
a>
6
u
o 0.6 h
<**
CD
d
C8

.2 0.4
Surface

0.2 A = - 0.011
A = - 0.015
A = - 0.020
0.0
i

100 120 140

Time, s

Figure 12.37: Effect of varying the value of the kinetic constant in the Koisti-
nen-Marburger equation adopted in the finite-element simulation on the microstructural
evolution at the centre and subsurface of a 38.1 mm-dia. alloyed eutectoid steel bar
quenched with water flowing at 2.8 m s at 75 C ._ 1
Mathematical Analysis of Forced Convective Quenching 304

19.05 mm

Figure 12.38: Finite-element mesh adopted to simulate the thermal/microstructural evo-


lution and stress generation during quenching.
Mathematical Analysis of Forced Convective Quenching 305

T
38.1 mm

JL
191 mm H

Figure 12.39: Computational domain adopted for the finite-element simulation of stress
generation during forced convective quenching.

Figure 12.40: Schematic representation of the boundary conditions applied for the ther-
mal/mi crostructural simulations.
Mathematical Analysis of Forced Convective Quenching 306

Figure 12.41: Schematic representation of the boundary conditions applied for the stress
simulations.

Figure 12.42: Comparison between predicted and measured residual stress distribution
(radial component) in a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m
s- at 25 C .
1
Mathematical Analysis of Forced Convective Quenching 307

Figure 12.43: Comparison between predicted and measured residual stress distribution
(circumferential (hoop) component) in a 38.1 mm-dia. IF steel bar quenched with water
flowing at 4.8 m s at 25 C.
_1

Figure 12.44: Comparison between predicted and measured residual stress distribution
(axial component) in a 38.1 mm-dia. IF steel bar quenched with water flowing at 4.8 m
s- at 25 C .
1
Mathematical Analysis of Forced Convective Quenching 308

T 1
i 1
i 1
i 1
r
600
1045 Steel

400
cd
a.
S 200 f

-200

Radial Stress
-400 o Measured
FEM

-600

2 4 6 8 10 12 14 16 18 20

Radial Position, mm

Figure 12.45: Comparison between predicted and measured residual stress distribution
(radial component) in a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing
at 2.8 m s at 50 C . Modified kinetics.
_1

600
1045 Steel

400
td
PH

200

td
3
xi
-200
CO
Ct>
PH Hoop Stress
-400 Measured
FEM

-600

0 2 4 6 8 10 12 14 16 18 20

Radial Position, mm

Figure 12.46: Comparison between predicted and measured residual stress distribution
(circumferential (hoop) component) in a 38.1 mm-dia. 1045 carbon steel bar quenched
with water flowing at 2.8 m s at 50 C. Modified kinetics.
_1
Mathematical Analysis of Forced Convective Quenching 309

600 h 1045 Steel


^ A A

400
CM

200

CO

-200
in

Axial Stress
-400
A Measured
FEM
-600 h

0 2 4 6 8 10 12 14 16 18 20

Radial Position, mm

Figure 12.47: Comparison between predicted and measured residual stress distribution
(axial component) in a 38.1 mm-dia. 1045 carbon steel bar quenched with water flowing
at 2.8 m s" at 50 C. Modified kinetics.
1

1 1
i 1
r

600
Radial Stress
o Measured
FEM
400 h
a,
200

P
-200
CD

-400
A l l o y e d Steel
-600
j i i i i i i i i i i_
0 2 4 6 8 10 12 14 16 18 20

Radial Position, mm

Figure 12.48: Comparison between predicted and measured residual stress distribution
(radial component) in a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water
flowing at 2.8 m s at 75 C.
_1
Mathematical Analysis of Forced Convective Quenching 310

600 h
Hoop Stress
Measured
400 FEM
tS
CU

200

td

a -200
cn
cu


-400

Alloyed Steel
-600

_L_
6 8 10 12 14 16 18 20

Radial Position, mm

Figure 12.49: Comparison between predicted and measured residual stress distribu-
tion (circumferential (hoop) component) in a 38.1 mm-dia. alloyed eutectoid steel bar
quenched with water flowing at 2.8 m s at 75 C. - 1

600
Axial Stress
A Measured
400 FEM
tS
cu
200

cS

-200
CU

-400
Alloyed Steel

-600 h

4 6 8 10 12 14 16 18 20

Radial Position, mm

Figure 12.50: Comparison between predicted and measured residual stress distribution
(axial component) in a 38.1 mm-dia. alloyed eutectoid steel bar quenched with water
flowing at 2.8 m s" at 75 C.
1
I

Mathematical Analysis of Forced Convective Quenching 311

1000

800
CT
600 r
CT
400
z

200 CT.

CM 0

-200

-400

-600

-800

-1000
A l l o y e d Steel
-1200 Surface

-1400 _l_ _l_ _l_


20 40 60 80 100 120 140

Time, s

Figure 12.51: Evolution of radial, circumferential (hoop) and axial stresses at the surface
during forced convective quenching of a 38.1 mm-dia. alloyed eutectoid steel bar.

140

Figure 12.52: Evolution of radial, circumferential (hoop) and axial stresses at the centre
during forced convective quenching of a 38.1 mm-dia. alloyed eutectoid steel bar.
Mathematical Analysis of Forced Convective Quenching 312

400

300 a
r
200 CT
z
100
0" =
0
CM

-100

0) -200
u
-300

-400

-500
1045 Steel
-600 Surface

-700
10 20 30 40 50 60 70

Time, s

Figure 12.53: Evolution of radial, circumferential (hoop) and axial stresses at the surface
during forced convective quenching of a 38.1 mm-dia. 1045 carbon steel bar.

Figure 12.54: Evolution of radial, circumferential (hoop) and axial stresses at the centre
during forced convective quenching of a 38.1 mm-dia. 1045 carbon steel bar.
Mathematical Analysis of Forced Convective Quenching 313

T ' I 1
I 1
I 1
I ' V J _ 1
I 1
L ^ J I 1

Radial Position, mm

Figure 12.55: Calculated final microstructural distribution in a 100 mm-dia. alloyed


eutectoid steel bar quenched from 850 C with water flowing at 4.8 m s at 32 C . _1
Chapter 13

Summary and Conclusions

The research programme was undertaken to study, under laboratory conditions, an in-
dustrial operation aimed at producing grinding rods of improved abrasion resistance and
toughness. The motivation to manufacture a product of these characteristics, lies on
the fact that grinding constitutes the greatest operating cost in mineral processing [2].
The proprietary heat treatment process developed at AltaSteel is based on forced con-
vective quenching of steel rods, and produces a material with a martensitic shell and a
bainitic/pearlitic core. In order to optimize the process, as well as to assist in future
developments, it is desirable to generate a predictive tool that would obviate the need
for expensive and time-consuming trial-and-error operations.
The work presented involved a comprehensive study of heat transfer and stress gen-
eration during forced convective quenching of steel bars and was conducted under the
framework of microstructural engineering. Mathematical models of heat transfer, phase
transformations, and elasto-plastic stress were applied to predict the thermal response,
microstructural evolution and stress generation during the quench. Once debugged and
tested, the mathematical models were validated by comparing calculated results with
values measured in the laboratory. The effect of internal stresses on transformation ki-
netics was not considered. Thus, the models could be effectively uncoupled, i.e., the
results of the thermal/microstructural evolution model were adopted as input to the
stress generation model.

The problem at hand is highly nonlinear and, therefore, a numerical solution needed

314
Summary and Conclusions 315

to be implemented. In structural analysis, the finite-element method ( F E M ) has become


a de facto standard (due to its ability to handle components of complex geometry) and
was adopted for the analysis. Based on geometric considerations, an axisymmetric for-
mulation was deemed adequate. A finite-element program developed to simulate thermal
stresses in fused-cast monofrax-s refractories [182] was used as a basis for the tran-
sient thermal/microstructural model, while a computer program developed for the time-
independent, elasto-plastic analysis of stress evolution in water spray-quenching of steel
bars [47] was adopted to model internal stress generation. In the thermal/microstructural
model, the additivity principle was applied to compute the amount of fraction trans-
formed under continuous cooling conditions. Diffusional transformations (austenite-to-
ferrite,-pearlite, and -bainite) were described using the J M A K equation. The martensitic
transformation was modelled using the Koistinen and Marburger equation. The mechani-
cal model includes thermal- and transformation-related strains. In particular, strains due
to volume changes associated with cooling and phase transformations, as well as those
arising from transformation plasticity, variation of the elastic constants with tempera-
ture, and variation of the flow stress with temperature were included. The mechanical
properties were obtained from the literature.

A critical component of the modelling exercise is the characterization of the heat


transfer boundary condition. Due to a lack of data in the literature, for the conditions
of interest, the surface heat flux, as a function of surface temperature, was estimated
from the measured temperature response of instrumented samples through the solution
of the inverse heat conduction problem (IHCP). A n existing computer program [183]
that implements the sequential function specification technique, was modified for this
purpose.

The layout and operation of the industrial equipment allowed little flexibility for in-
strumentation to measure the thermal response of the steel rods during the quenching
cycle. This information is needed to estimate the heat transfer boundary condition.
Summary and Conclusions 316

Thus, a laboratory facility was designed and built to simulate the industrial operation.
The objectives of the laboratory experiments were twofold : 1) to characterize the surface
heat flux as a function of surface temperature, under forced convective boiling conditions,
and 2) to produce specimens for metallographic characterization and residual stress mea-
surement (to validate the mathematical models). A test matrix consisting of three water
velocities (2.8, 4.8 and 6.9 m/s) and three water temperatures (25, 50 and 75 C) was
adopted to obtain the boiling curves using IF steel as test material. Additional tests were
conducted using alloyed eutectoid and 1045 carbon steels. The effect of an oxide layer
was also studied.

The kinetics of the austenite decomposition in 3 alloyed, near-eutectoid steels for a


range of continuous cooling and isothermal conditions were determined using the G L E E -
B L E 1500 thermomechanical simulator. The transformation kinetics of continuously
cooled IF and 1045 carbon steel samples were also measured. The transformation studies
were carried out using thin-wall tubular samples (6 mm I.D., 8 mm O.D.). The aim of
the isothermal and continuous cooling tests was to describe the transformation in terms
of the Avrami equation and to determine the transformation start times, respectively.
The results of the heat transfer measurements showed that, for a given water velocity,
the surface heat flux is higher as the water temperature decreases. The total amount
of heat extracted is, therefore, higher as the water temperature decreases (subcooling
increases). The boiling curves obtained for water flowing at 25 and 50 C showed sim-
ilar features. In these cases, no film boiling stage could be identified, i.e., the boiling
curves immediately reached the transition stage. The similarity in the boiling curves was
particularly noticeable for the intermediate (4.8 m s ) and high (6.9 m s ) water ve-
_1 -1

locities investigated. In contrast, the boiling curves corresponding to quenching in water


at 75 C did show a film boiling stage. Thermal responses in clean, 'lightly' and 'heavily'
oxidized specimens were not significantly different.
Summary and Conclusions 317

To produce specimens for residual stress measurements and metallographic charac-


terization, bars of IF, 1045 carbon and alloyed eutectoid steel were quenched in the lab-
oratory facility. The residual stress distributions in selected IF, 1045 carbon and alloyed
eutectoid steel quenched bars were measured using neutron diffraction. The variation of
measured axial and circumferential residual stresses with radial position, in the IF and
1045 carbon steel quenched bars, were similar. In both cases, they were compressive at
the surface and tensile at the centre, while the radial component was always tensile. On
the other hand, the three components obtained in the alloyed eutectoid steel bar were
compressive at the centre. Metallographic analysis of the IF steel bar showed that the
austenite had completely transformed to ferrite. The alloyed eutectoid steel exhibited a
fully martensitic structure, while the 1045 carbon steel showed an outer ring of martensite
and a core consisting of a mixture of diffusional and martensitic products. The measured
hardness distributions were consistent with the observed microstructures.
Comparisons between measured and model-predicted thermal responses, final mi-
crostructure and residual stress distributions have been made. It was found that the
position of the 'nose' of the continuous cooling diagram, when a mixture of diffusional
and martensitic products was present, has a significant influence on the predicted final
microstructure distribution and, therefore, on the predicted residual stress distribution.
The difference in the measured residual stress distributions obtained in the alloyed eu-
tectoid steel specimen, when compared with those found in the IF and 1045 carbon steel
quenched bars, has been explained based on the sequence of transformations that took
place during the quench. The models were also applied to obtain a qualitative description
of the microstructural evolution under industrial conditions.

Given the strong effect of water temperature on boiling heat transfer, it is recom-
mended that the water temperature be monitored and controlled tightly in the industrial
operation. The difference between inlet and exit water temperatures should be kept to
a minimum in order to prevent significant differences in mechanical properties along the
Summary and Conclusions 318

length of the bars. Possible modifications to the process can be envisioned, and are
now described. A n optimum cooling rate that minimizes transient internal stresses, and
consequently distortion, while producing the desired final microstructural distribution
consists of slow cooling at high temperatures, to minimize thermal stresses, rapid cooling
at intermediate temperatures to achieve the targeted microstructural changes, and slow
cooling at lower temperatures. At the present time, the inlet water temperature and ve-
locity are kept constant throughout the quench; one action that could be taken would be
to vary water temperature and/or velocity to modify the cooling path. Another option
would involve the use of oil or aqueous polymeric solutions as quench medium; however,
issues of recyclability and cost would need to be considered. The quench time controls
the maximum temperature attained during self-tempering. Currently, the water flow is
stopped completely after a given quench time; alternatively, the water flow rate could
be reduced at a controlled rate to modify the self-tempering conditions. Progressive
induction hardening, where the workpiece is moved at a constant speed through a coil
and cooling ring, has been used to produce parts with a hard surface with compressive
stresses and a tough core with tensile stresses. The process is based on the interaction
betweeen eddy currents generated by the magnetic field induced by the coil, and the
workpiece. The heat area is confined to few millimeters below the surface. This thin
layer is subsequently quenched by the cooling ring. For highly alloyed steels, air could
be used as the quench medium.

Future work should include temperature measurements under operational conditions


in order to devise a strategy to scale the laboratory results to the plant operation. A
comprehensive experimental programme to evaluate the mechanical properties of the
alloyed eutectoid steel, as a function of temperature and phase composition, should be
undertaken. The mechanical model may be modified to include kinematic and, eventually,
mixed strain hardening. Research aimed at clarifying the role of transformation plasticity
and the effect of internal stresses on phase transformation kinetics need to be done.
Bibliography

C. K i m . Determination of Residual Stresses in Austenite and Martensite in Case-


Hardened Steels by the sin xp Method. Adv. X-ray Anal., 25:343 - 353, 1981.
2

B . A . Wills. Mineral Processing Technology, page 13. Pergamon Press, Oxford, 4th
edition, 1988.

A . K . Gangopadhyay and J.J. Moore. Assessment of Wear Mechanisms in Grinding


Media. Minerals and Metallurgical Processing, 2(3):145 - 151, 1985.

S.G. Malghan. Methods to Reduce Steel Wear in Grinding Mills. Mining Engi-
neering, 34(6):684 - 690, June 1982.

C. A . Rowland, Jr. and D . M . Kjos. Rod and Ball Mills. In A . L . Mular and R . B .
Bhappo, editors, Mineral Processing Plant Design, pages 239 - 278. Society of
Mining Engineers of A I M E , New York, 1980.

R . W . Pugh and S.L. M a . Heat Treated Grinding Rod. In 25th Annual Meeting
of the Canadian Mineral Procesors. The Canadian Institute of Mining, Metallurgy
and Petroleum, 1993. Paper No. 9.

T. Inoue and K . Tanaka. A n Elastic-Plastic Stress Analysis of Quenching When


Considering a Transformation. Int. J. Mech. Sci., 17:361 - 367, 1975.

J.L. Myers. Fundamentals of Experimental Design. Allyn and Bacon, Inc., Boston,
2nd edition, 1972.

R. B . Frank and R. K . Mahidhara. Effect of Heat Treatment on Mechanical Prop-


erties and Microstructure of Alloy 901. In S. Reichman, D. N . Duhl, G . Maurer,
S. Antolovich, and C. Lund, editors, Superalloys 1988, pages 23 - 32. The Metal-
lurgical Society, 1988.

T. E . Lim. Optimizing Heat Treatment with Factorial Design. The Journal of The
Minerals, Metals and Materials Society, pages 52 - 53, March 1989.

G . E . Totten, M . E . Dakins, and R . W . Heins. Cooling Curve Analysis of Synthetic


Quenchants - A Historical Perspective. J. Heat Treating, 6(2):87 - 95, 1988.

Robert M . Brick, Alan W . Pense, and Robert B . Gordon. Structure and Properties
of Engineering Materials, page 282. Mc Graw-Hill, 4th edition, 1977.

319
Bibliography 320

[13] M . A . Grossman. Hardenability Calculated from Chemical Composition. Trans.


TMS-AIME, 150:227 - 259, 1942.

[14] J.S. Kirkaldy and D. Venugopalan. Prediction of Microstructure and Hardenability


in Low Alloy Steels. In A . R . Marder and J.I. Goldstein, editors, Phase Transfor-
mations in Ferrous Alloys, pages 125 - 148. The Metallurgical Society of A I M E ,
1983.

[15] D. V . Doane. A Critical Review of Hardenability Predictors. In D. V . Doane and


J. S. Kirkaldy, editors, Hardenability Concepts with Applicatons to Steel, pages 351
- 396. The Metallurgical Society, 1978.

[16] A . L. Boegehold. Hardenability Control for Alloy Steel Parts. Metal Progress,
53(5):697 - 709, 1948.

[17] J . J . Lakin. Wolfson Heat Treatment Data Sheets - A New Presentation of Hard-
enability Information. Heat Treat, of Metals, 2(1):7 - 10, 1975.

[18] J . S. Kirkaldy and S. E . Feldman. Optimization of Steel Hardenability Control. J.


Heat Treat, 7:57 - 64, 1989.

[19] M . Umemoto, N . Komatsubara, and I. Tamura. Prediction of Hardenability Effects


from Isothermal Transformation Kinetics. J. Heat Treating, pages 57 - 64, June
1980.

[20] G . T. Eldis. A Critical Review of Data Sources for Isothermal Transformation and
Continuous Cooling Transformation Diagrams. In D. V . Doane and J . S . Kirkaldy,
editors, Hardenability Concepts with Applicatons to Steel, pages 126 - 148. The
Metallurgical Society, 1978.

[21] A . Rose and W . Strassburg. Anwendung des Zeit-Temperatur-Umwandlungs-


Schaubildes fur Kontinuierliche Abkuhlung auf Fragen der Warmebehandlung.
Arch. Eissen., 24:505 - 514, 1953.

[22] W . W . Cias. Phase transformation kinetics and hardenability of medium-carbon


alloy steels. Technical report, Climax Molybdenum Co., 1977.

[23] American Society for Metals. ASM Handbook, volume 4, pages 20 - 32. A S M
International, Warrendale, P A , 10th edition, 1990.

[24] Ph. Maynier, J . Dollet, and P. Bastien. Prediction of Microstructure V i a Empirical


Formulae Based on C C T Diagrams. In D. V . Doane and J . S. Kirkaldy, editors,
Hardenability Concepts with Applicatons to Steel, pages 163 - 178. The Metallur-
gical Society, 1978.
Bibliography 321

[25] American Society for Testing and Materials. Standard Test Method for Quenching
Time of Heat-Treating Fluids (Magnetic Quenchometer Method), pages 310 - 316.
A S T M , Philadelphia, P A , 1981.

[26] C . E . Bates and G . E . Totten. Quantifying Quench-Oil Cooling Characteristics.


Advanced Materials and Processes, 139(3):25 - 28, 1991.

[27] I. Tamura, K . Fukuhara, and S. Asada. Comparative Study of Cooling Curves with
JIS Silver Specimens and Alloy 600 Specimens in Relation to Additive Effectivenes.
J. of Materials Engineering and Performance, 3(3):367 - 370, 1994.

[28] J . Bodin and S. Segerberg. Measurement and Evaluation of the Quenching Power
of Quenching Media for Hardening. In G . E . Totten, editor, 1st Int. Conf. on
Quenching and Distortion Control, pages 1 - 12.. A S M International, 1992.

[29] J . W . Evancho and J.T. Staley. Kinetics of Precipitation in Aluminum Alloys Dur-
ing Continuous Cooling. Metallurgical Transactions, 5:43 - 47, 1974.

[30] J.T. Staley. Quench Factor Analysis of Aluminium Alloys. Materials Science and
Technology, 3:923 - 935, 1987.

[31] E . Scheil. Anlaufzeit der Austenitumwandlung. Arch. Eissen., pages 565 - 567,
June 1935.

[32] C. E . Bates. Selecting Quenchants to Maximize Tensile Properties and Minimize


Distortion in Aluminum Parts. J. Heat Treating, 5(1):27 - 40, 1987.

[33] C . E . Bates and G . E . Totten. Application of Quench Factor Analysis to Predict


Hardness Under Laboratory and Production Conditions. In G.E. Totten, editor, 1st
Int. Conf. on Quenching and Distortion Control, pages 33 - 39. A S M International,
1992.

[34] T. Reti, I. Czinege, M . Reger, and J . Takacs. A n Indirect Method for Quantitative
Characterization of the Quenching Power of Quenching Media. In G . E . Totten,
editor, 1st Int. Conf. on Quenching and Distortion Control, pages 301 - 309. A S M
International, 1992.

[35] K . - E . Theling. C C T Diagrams with Natural Cooling. Scandinavian J. of Metal-


lurgy, 7(6):252 - 263, 1978.

[36] N . Shimizu and I. Tamura. Effect of Discontinuous Change in Cooling Rate during
Continuous Cooling on Pearlitic Transformation Behavior of Steel. Trans. ISIJ,
17(8):469 - 476, 1977.
Bibliography 322

[37] N . Shimizu and I. Tamura. A n Examination of the Relation between Quench-


hardening Behavior of Steel and Cooling Curve in Oil. Trans. ISIJ, 18(7):445 -
450, 1978.

[38] S. M . Gupta. Measurement of Quench Heat Transfer Coefficients and Their Use in
Heat Treatment Design. Master's thesis, University of British Columbia, 1977.

[39] B . Liscic and T. Filetin. Computer-Aided Evaluation of Quenching Intensity and


Prediction of Hardness Distribution. J. Heat Treating, 5(2):115 - 124, 1988.

[40] B . Liscic, S. Svaic, and T. Filetin. Workshop Designed System for Quenching Inten-
sity Evaluation and Calculation of Heat Transfer Data. In G . E . Totten, editor, 1st
Int. Conf. on Quenching and Distortion Control, pages 17 - 26. A S M International,
1992.

[41] H . M . Tensi and A Stich. Possibilities and Limits to Predict the Quench Harden-
ing of Steel. In G . E . Totten, editor, 1st Int: Conf. on Quenching and Distortion
Control, pages 27 - 32. A S M International, 1992.

[42] P.C. Campbell, E . B . Hawbolt, and J . K . Brimacombe. Microstructural Engineering


Applied to the Controlled Cooling of Steel Wire Rod : Part I. Experimental Design
and Heat Transfer. Metall. Trans A, 22A:2769 - 2778, 1991.

[43] 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. Metall. Trans A, 22A:2779 -
2790, 1991.

[44] P.C. Campbell, E . B . Hawbolt, and J . K . Brimacombe. Microstructural Engineering


Applied to the Controlled Cooling of Steel Wire Rod : Part III. Mathematical
Model - Formulation and Predictions. Metall. Trans A, 22A:2769 - 2778, 1991.

[45] R. A . Wallis, P. R. Bhoval, N . M . Bhathena, and Raymond E . L. Modeling the


Heat Treatment of Superalloy Forgings. The Journal of The Minerals, Metals and
Materials Society, pages 35 - 37, Feb. 1989.

[46] D. Persampieri, A . San Roman, and P.D. Hilton. Process Modeling for Improved
Heat Treating. Advanced Materials and Processes, 139(3): 19 - 23, 1991.

[47] Y . Nagasaka, J . K . Brimacombe, E . B . Hawbolt, I.V. Samarasekera, B . Hernandez-


Morales, and S.E. Chidiac. Mathematical Model of Phase Transformations and
Elastoplastic Stress in the Water Spray Quenching of Steel Bars. Met. Trans. A,
24A:795 - 808, 1993.
Bibliography 323

[48] T. Inoue, D - Y . Ju, and K . Arimoto. Metallo-Thermo-Mechanical Simulation of


Quenching Process - Theory and Implementation of Computer Code 'Hearts'. In
G.E. Totten, editor, 1st Int. Conf. on Quenching and Distortion Control, pages 205
- 212. A S M International, 1992.

[49] Sponzilli Keith, C. J., Sharma J . T., V . K . , and G . H . Walter. International


Harvester's C H A T System for Selecting Optimum Compositions for Heat Treated
Steels. In D. V . Doane and J. S. Kirkaldy, editors, Hardenability Concepts with
Applicatons to Steel, pages 493 - 517. The Metallurgical Society, 1978.

[50] Ph. Maynier, B . Jungmann, and J . Dollet. Creusot-Loire System for the Prediction
of the Mechanical Properties of Low Alloy Steel Products. In D. V . Doane and J . S.
Kirkaldy, editors, Hardenability Concepts with Applicatons to Steel, pages 518 - 545.
The Metallurgical Society, 1978.

[51] S. E . Feldman. The Minitech Computerized Alloy Steel Information System. In


D. V . Doane and J . S. Kirkaldy, editors, Hardenability Concepts with Applicatons
to Steel, pages 546 - 567. The Metallurgical Society, 1978.

[52] M . Gergely and T. Reti. Application of a Computerized Information System for


the Selection of Steels and Their Heat Treatment Technologies. J. Heat Treat.,
5:125 - 140, 1988.

[53] American Society for Metals. ASM Handbook, volume 4, pages 638 - 656. A S M
International, Warrendale, P A , 10th edition, 1990.

[54] J . Bodin and S. Segerberg. Benchmark Testing of Computer Programs for De-
termination of Hardening Peformance. In G . E . Totten, editor, 1st Int. Conf. on
Quenching and Distortion Control, pages 133 - 139. A S M International, 1992.

[55] M . Schwalm and H . M . Tensi. Surface Reactions during Immersion Quenching of


Metallic Parts. In D . B . Spalding and N . H . Afgan, editors, Heat and Mass Transfer
in Metallurgical Systems, pages 565 - 572. Hemisphere Publishing Corporation,
1981.

[56] M . Mitsutsuka and K . Fukuda. The Transition Boiling and Characteristic Tem-
perature in Cooling Curve During Water Quenching of Heated Metal. Trans. ISIJ,
16(1):46 - 50, 1976.

[57] J.C. Chevrier, F . Moreaux, and G . Beck. L'Effusivit'e et la R'esistance Thermique


des Zones Superficielles du Solide Determinent le Processus de Vaporisation du
Liquide en Regime de Trempe. Int. J. Heat Mass Transfer, 15:1631 - 1645, 1972.
Bibliography 324

C . E . Bates, G . E . Totten, and R.L. Brennan. Quenching of Steel, volume 4, pages


67 - 120. A S M International, 10th edition, 1991.

T . E . Diller. Advances in Heat Flux Measurements. In J.P. Hartnett and T . F .


Irvine Jr., editors, Advances in Heat Transfer, volume 23, pages 279 - 368. Aca-
demic Press, San Diego, 1993.

M . Bamberger and B . Prinz. Determination of Heat Transfer Coefficients During


Water Cooling of Metals. Materials Science and Technology, 2:410 - 415, April
1986.

V . K . Dhir. Nucleate and Transition Boiling Heat Transfer under Pool and External
Flow Conditions. In G . Hetsroni, editor, Heat Transfer 1990, volume 1, pages 129
- 155, New York, 1990. Hemisphere Publishing Comp.

H . Auracher. Transition Boiling. In G . Hetsroni, editor, Heat Transfer 1990,


volume 1, pages 69 -90, New York, 1990. Hemisphere Publishing Comp.

A . Sakurai. Film Boiling Heat Transfer. In G . Hetsroni, editor, Heat Transfer


1990, volume 1, pages 129 - 155, New York, 1990. Hemisphere Publishing Comp.

S. Nukiyama. The Maximum and Minimum Values of Heat Transmitted from


Metal to Boiling Water Under Atmospheric Pressure. Int. J. Heat Mass Transfer,
9:1419 - 1435, 1966.

W . M . Rohsenow. A Method of Correlating Heat Transfer Data to Surface Boiling


Liquids. Trans. ASME, 74:969 - 976, 1952.

J. G . Collier. Heat Transfer in the Postburnout Region and During Quenching and
Reflooding. In Handbook of Multiphase Systems, pages 6-142 - 6-188. Hemisphere
Publishing Corp., Washington, 1982.

G.F. Hewitt. Burnout. In Handbook of Multiphase Systems, pages 6-66 - 6-142.


Hemisphere Publishing Corp., Washington, 1982.

A . E . Bergles and W . G . Thompson, Jr. The Relationship of Quench Data to Steady-


State Pool Boiling Data. Int. J. Heat Mass Transfer, 13:55 - 68, 1970.

Peyayopanakul W . and J.W. Westwater. Evaluation of the Unsteady-State Quench-


ing Method for Determining Boiling Curves. Int. J. of Heat and Mass Transfer,
21:1437 - 1445, 1978.

X . C . Huang, P. Weber, and G. Bartsch. Comparison of Transient and Steady-State


Boiling Curves for Forced Upflow of Water in a Circular Tube at Medium Pressure.
Int. Comm. Heat Mass Transfer, 20:383 - 392, 1993.
Bibliography 325

[71] S.C. Cheng, W . W . L . Ng, and K . T . Heng. Measurements of Boiling Curves of


Subcooled Water Under Forced Convective Conditions. Int. J. Heat Mass Transfer,
21:1385 - 1392, 1978.

[72] H.S. Ragheb, S.C. Cheng, and D . C . Groeneveld. Observations in Transition Boil-
ing of Subcooled Water under Forced Convective Conditions. Int. J. Heat Mass
Transfer, 24(7):1127 - 1137, 1981.

[73] S. Ishigai, S. Nakanishi, and T. Ochi. Boiling Heat Transfer for a Plane Water
Jet Impinging on a Hot Surface. In Sixth International Heat Transfer Conference,
volume 1, pages 445 - 450, 1978.

[74] D . H . Wolf, F.P. Incropera, and R. Viskanta. Jet Impingement Boiling. In J.P.
Hartnett and T . F . Irvine Jr., editors, Advances in Heat Transfer, volume 23, pages
1 - 132. Academic Press, San Diego, 1993.

[75] I. Mudawar and W.S. Valentine. Determination of the Local Quench Curve for
Spray-Cooled Metallic Surfaces. J. Heat Treat., 7:107 - 121, 1989.

[76] M . S . Sohal. Critical Heat Flux in Flow Boiling of Helium I. In J.P. Hartnett and
T . F . Irvine Jr., editors, Advances in Heat Transfer, volume 17, pages 319 - 340.
Academic Press, Orlando, 1985.

[77] T. Tanaka. Overview of Accelerated Cooling of Steel Plates. In G . E . Ruddle


and A . F . Crawley, editors, Accelerated Cooling of Rolled Steel, pages 187 - 208.
The Metallurgical Society of the Canadian Institute of Mining, Metallurgy and
Petroleum, 1987.

[78] I. V . Samarasekera and J . K . Brimacombe. Thermal and Mechanical Behaviour of


Continuous-Casting Billet Moulds. Ironmaking and Steelmaking, 9(1):1 - 15, 1982.

[79] B . K . H . Sun. High Temperature Heat Transfer Application to Nuclear Power Safety
Reflooding and Core Unrecovery Phenomena. In Mizushina, T. and Yang, W.-J.,
editor, Heat Transfer in Energy Problems, pages 167 - 171. United States-Japan
Cooperative Science Program, Hemisphere Publishing Corporation, 1983.

[80] R . B . Duffey and D . T . C . Porthouse. The Physics of Rewetting in Water Reactor


Emergency Core Cooling. Nuclear Engineering and Design, 25:379 - 394, 1973.

[81] M . Belhadj, T. Aldemir, and R . N . Christensen. Determining Wall Superheat under


Fully Developed Nucleate Boiling in Plate-Type Research Reactor Cores with Low-
Velocity Upward Flows. Nuclear Technology, 95:95 - 102, 1991.

[82] G . Yadigaroglu. The Reflooding Phase of the L O C A in P W R s . Part I : Core Heat


Transfer and Fluid Flow. Nuclear Safety, 19(1):20 - 36, 1978.
Bibliography 326

E. Elias and G . Yadigaroglu. The Reflooding Phase of the L O C A in P W R s . Part


II : Rewetting and Liquid Entrainment. Nuclear Safety, 19(2):160 - 175, 1978.

R. Nelson and C. Unal. A Phenomenological Model for the Thermal Hydraulics


of Convective Boiling during the Quenching of Hot Rod Bundles Part I : Thermal
Hydraulic Model. Nuclear Engineering and Design, 136:277 - 298, 1992.

M . Ishii and G . DeJarlais. Flow Visualization Study of Inverted Annular Flow of


Post Dryout Heat Transfer Region. Nuclear Engineering and Design, 99:187 - 199,
1987.

D. P Incropera, F.P. and. De Witt. Fundamentals of Heat and Mass Transfer,


chapter 10. Wesley & Sons, New York, third edition, 1990.

W . M . Rohsenow. General Boiling. In G . Hestroni, editor, Handbook of Multiphase


Systems, pages 6.5 - 6.26. Hemisphere Publishing Corp., Washington, D . C . , 1982.

C.L. Vandervort, A . E . Bergles, and M . K . Jensen. Heat Transfer Mechanisms in


Very High Heat Flux Subcooled Boiling. In R.D. Boyd and S.G. Kandlikar, editors,
Fundamentals of Subcooled Flow Boiling, pages 1 - 9 . A S M E , 1992.

G.P. Celata, M . Cumo, and A . Mariani. Assesement of Correlations and Models


for the Prediction of C H F in Water Subcooled Flow Boiling. Int. J. of Heat and
Mass Transfer, 37(2):237 - 255, 1994.

W . A . Johnson and R . F . Mehl. Reaction Kinetics in Processes of Nucleation and


Growth. Trans. AIME, 135:416 - 442, 1939.

M . Avrami. Kinetics of Phase Change I. J. Chem. Phys., 7:1103 - 1112, 1939.

M . Avrami. Kinetics of Phase Change II. J. Chem. Phys., 8:212 - 224, 1940.

M . Avrami. Kinetics of Phase Change III. J. Chem. Phys., 9:177 - 184, 1941.

A . N . Kolmogorov. Izv. Akad. nauk USSR Ser. Matemat., 1:355, 1937.

J. W . Cahn. Transformation Kinetics During Continuous Cooling. Acta Met.,


4:572 - 575, 1956.

P. K . Agarwal and J . K . Brimacombe. Mathematical Model of Heat Flow and


Austenite-Pearlite Transformation in Eutectoid Carbon Steel Rods for Wire. Met.
Trans. B, 12B:121 - 133, 1981.

E. B . Hawbolt, B . Chau, and J . K . Brimacombe. Kinetics of Austenite-Pearlite


Transformations in Eutectoid Carbon Steel. Met. Trans. A, 14A:1803 - 1815, 1983.
Bibliography 327

[98] E . B . Hawbolt, B . Chau, and J . K . Brimacombe. Kinetics of Austenite-Ferrite and


Austenite-Pearlite Transformations in 1025 Carbon Steel. Met. Trans. A, 16A:565
- 578, 1985.

[99] D.P. Koistinen and R . E . Marburger. A General Equation Prescribing the Extent
of the Austenite-Martensite Transformation in Pure Iron-Carbon Alloys and Plain
Carbon Steels. Acta Metallurgica, 7:59 - 60, 1959.

[100] C. Zener. Kinetics of Decomposition of Austenite. Trans. AIME, 167:550 - 595,


1946.

[101] K . C . Russell. Grain Boundary Nucleation Kinetics. Acta Metall, 17:1123 - 1131,
1969.

[102] Y . Sakamoto, M . Saeki, M . Nishida, T. Tamaka, and Y . Ito. Report, Kawasaki


Steel Co., 1981.

[103] J.S. Kirkaldy, G.O. Pazionis, and S.E. Feldman. A n Accurate Predictor for the
Jominy Hardenability of Low-Alloy Hypoeutectoid Steels. In Heat Treatment '76,
pages 169 - 200. The Metallurgical Society, 1976.

[104] J . S. Kirkaldy, B . A . Thomson, and E . A . Baganis. Prediction of Multicomponent


Equilibrium and Transformation Diagrams for Low Alloy Steels. In D. V . Doane
and J . S. Kirkaldy, editors, Hardenability Concepts with Applicatons to Steel, pages
82 - 125. The Metallurgical Society, 1978.

[105] T. T. Pham. Mathematical Modelling of the Onset of Transformation from Austen-


ite to Pearlite under Non-Continuous Cooling Conditions. Master's thesis, Univer-
sity of British Columbia, 1993.

[106] K . Mukunthan. Evolution of Microstructure and Texture during Continuous An-


nealing of Cold Rolled, Ti-Stabilized Interstitial-Free Steel. PhD thesis, University
of British Columbia, 1994.

[107] P.T. Moore. Anisothermal Decomposition of Austenite in a Medium-Alloy Steel.


J. of The Iron and Steel Institute, 7:305 - 311, 1954.

[108] J . K . Iyer, J . K . Brimacombe, and E . B . Hawbolt. Prediction of the structure


and mechanical properties of control-cooled eutectoid steel rods. In Mechanical
Working and Steel Processing XXII, pages 47 - 58. ISS, 1985.

[109] P. Campbell, J . K . Brimacombe, and E. B . Hawbolt. Application of Microstructural


Engineering to the Controlled Cooling of Steel Wire Rod. In G . E . Ruddle and A . F .
Crawley, editors, Accelerated Cooling of Rolled Steel, pages 309 - 330. C I M , 1987.
Bibliography 328

J. W . Cahn. On the Kinetics of the Pearlite Reaction. Trans. AIME, pages 140 -
144, January 1957.

J. W . Christian. The Theory of Transformation in Metals and Alloys. Pergamon


Press, 1975.

M . B . Kuban, R. Jayaraman, E . B . Hawbolt, and J . K . Brimacombe. A n Ass-


esement of the Additivity Principle in Predicting Continuous Cooling Austenite-
to-Pearlite Transformation Kinetics Using Isothermal Transformation Data. Met.
Trans. A, 17A:1493 - 1503, 1986.

C. Verdi and A . Visintin. A Mathematical Model of the Austenite-Pearlite Trans-


formation in Plain Carbon Steel Based on the Scheil's Additivity Rule. Acta Met-
ali, 35(11):2711 - 2717, 1987.

R. Kamat, E . B . Hawbolt, L. C. Brown, and J. K . Brimacombe. Diffusion Mod-


elling of Pro-Eutectoid Ferrite Growth to Examine the Principle of Additivity. In
Diffusion Analysis and Applications, pages 1 - 2 . The Metallurgical Society, 1988.

D. J. Riehm. Kinetics of the Pearlite to Austenite Reversion Transformation. Mas-


ter's thesis, University of British Columbia, 1990.

K . W . Andrews. Empirical Formulae for the Calculation of Some Transformation


Temperatures. J. of the Iron and Steel Institute, pages 721 - 727, July 1965.

J . S . Kirkaldy. Prediction of Alloy Hardenability from Thermodynamic and Kinetic


Data. Met. Trans., 4:2327 - 2333, 1973.

K . Hashiguchi, J.S. Kirkaldy, T. Fukuzumi, and V . Pavaskar. Prediction of the


Equilibrium, Paraequilibrium and No-Partition Local Equilibrium Phase Diagrams
for Multicomponent Fe-C Base Alloys. Computer Coupling of Phase Diagrams and
Thermochemistry (CALPHAD), 8(2):173 - 186, 1984.

W . Steven and A . G . Haynes. The temperature of Formation of Martensite and


Bainite in Low-Alloy Steels. JISI, 183:349 - 359, 1956.

G. B . Olson and M . Cohen. Kinetics of Strain-Induced Martensitic Nucleation.


Met. Trans. A, 6:791 - 795, April 1975.

L. Remy. Kinetics of Strain-Induced F C C - H C P Martensitic Transformation. Met.


Trans. A, 8 A:253 - 258, Feb. 1977.

S. S. Hecker, M . G. Stout, K . P. Staudhammer, and J . L. Smith. Effects of Strain


State and Strain Rate on Deformation-Induced Transformation in 304 Stainless
Steel : Part I. Magnetic Measurements and Mechanical Behaviour. Met. Trans. A,
13 A:619 - 626, April 1982.
Bibliography 329

M . Umemoto, H . Ohtsuka, and I. Tamura. Transformation to Pearlite from Work-


Hardened Austenite. Trans. ISIJ, 23(9):775 - 784, Feb. 1983.

V . A . Dubrov. High-Temperature Metallographic Analysis of the Influence of Stress


on the Bainitic Transformation Kinetics. Fiz. Metal. Metalloved., 28(2):309 - 314,
1969.

M . Violle. Influence d'une Contrainte Uniaxiale de Compression sur la Cinetique de


la Transformation (3 a d'Alliages Uranium-Chrome et sur leur Etat de Cristalli-
sation en Phase a. Mem. Sci. Rev. Met., 65(5):415 - 424, 1968.

E. Gautier. Transformations Perlitique et Martensitique sous Contrainte de Trac-


tion dans les Aciers. PhD thesis, L'Institut National Polytechnique de Lorraine,
1985.

F. M . B . Fernandes, S. Denis, and A . Simon. Mathematical Model Coupling Phase


Trasnformation and Temperature Evolution During Quenching of Steels. Mater.
Sci. Technoi, 1(10):838 - 844, October 1985.

M . Andre, E . Gautier, A . Simon, and G . Beck. Influence of Quenching Stresses on


the Hardenability of Steel. Materials Science and Engineering, 55:211 - 217, 1982.

S. Denis, E . Gautier, S. Sjostrom, and A . Simon. Influence of Stresses on the


Kinetics of Pearlitic Transformation during Continuous Cooling. Acta Metall.,
35(7):1621 - 1632, 1987.

American Welding Society. Welding Handbook, volume 1, pages 222 - 278. Amer-
ican Welding Society, Warrendale, P A , 7th edition, 1976.

L C . Noyan and J . B . Cohen. Residual Stress. Measurement by Diffraction and In-


terpretation. Springer-Verlag, New York, 1st edition, 1986.

C. K i m . Fracture of Case-Hardened Steels and Residual Stress Effects. In Case-


Hardened Steels : Microstructural and Residual Stress Effects, pages 59 - 87. A I M E ,
1984.

Measurements Group. Measurement of Residual Stresses by the Hole-Drilling


Strain-Gage Method. Technical Report TN-503-2, Measurements Group, 1986.

G. S. Schajer. Measurement of Non-Uniform Residual Stresses Using the Hole-


Drilling Method. Part I - Stress Calculation Procedures. Trans, of the ASME,
110:338 - 343, October 1988.
Bibliography 330

[135] G.S. Schajer. Measurement of Non-Uniform Residual Stresses Using the Hole-
Drilling Method. Part II - Practical Application of the Integral Method. Trans, of
the ASME, 110:344 - 349, October 1988.

[136] D. Rosenthal and J.T. Norton. A Method of Measuring Triaxial Residual Stress in
Plates. Welding Journal Research Supp., pages 295s - 307s, 1945.

[137] American Society for Testing and Materials. Standard Test Method for Determin-
ing Residual Stresses by the Hole-Drilling Strain-Gage Method, pages 713 - 760.
A S T M , Philadelphia, P A , 1989.

[138] P.F. Willemse, B . P. Naughton, and C. A . Verbraak. X-Ray Residual Stress


Meauserement on Cold-Drawn Steel Wire. Materials Science and Engineering,
pages 25 - 37, 1982.

[139] C A . Peck. Practical Aspects of Residual Stress Measurement by X-Ray Diffraction.


In W . B . Young, editor, Residual Stress in Design, Process and Materials Selection,
pages 7 - 9 . A S M International, 1987.

[140] A . I . Johnston, T.R. Finlayson, and J.R. Griffiths. Triaxial Stress Measurement
Using Neutron Diffraction. Materials Science Forum, 27 - 28:465 - 470, 1988.

[141] H . J . Prask and C.S. Choi. Residual Stress Measurements in Armament-System


Components by Means of Neutron Diffraction. In W . B . Young, editor, Residual
Stress in Design, Process and Materials Selection, pages 21 - 26. A S M Interna-
tional, 1987.

[142] T . R . Lambrineas, T.R. Finlayson, J.R. Griffiths, C . J . Howard, and T . F . Smith.


Neutron Diffraction Residual Stress Measurements on a Thin Steel Plate. NDT
International, XXX(5):285 - 290, 1987.

[143] J.H. Root, T . M . Holden, J . Schroder, C R . Hubbard, S. Spooner, T . A . Dodson, and


S.A. David. Residual Stress Mapping in Multipass Ferritic Steel Weld. Materials
Science and Technology, 9:754 - 759, 1993.

[144] M . E . Todaro, M . A Doxbeck, and G.P. Capsimalis. Residual Stress in Quenched


Steel Cylinders. In W . B . Young, editor, Residual Stress in Design, Process and
Materials Selection, pages 59 - 62. A S M International, 1987.

[145] K . Tiito. Use of Barkhausen Effect in Testing for Residual Stresses and Material
Defects. In W . B . Young, editor, Residual Stress in Design, Process and Materials
Selection, pages 27 - 36. A S M International, 1987.

[146] C O . Ruud. A Review of Nondestructive Methods for Residual Stress Measurement.


J. of Metals, 33(7):35 - 40, 1981.
Bibliography 331

[147] American Society for Testing and Materials. Standard Test Method for Mercurous
Nitrate Test for Copper and Copper Alloys, pages 274 - 275. A S T M , Philadelphia,
PA, 1989.

[148] H . Fujio, T. Aida, and Y . Masumoto. Distortions and Residual Stresses of Gears
Caused by Hardening. Bull, of JSME, 20(146):1051 - 1058, 1977. -

[149] F . G . Rammerstorfer, D.F. Fischer, W . Mitter, K . J . Bathe, and M . D . Snyder. On


Thermo-Elastic-Plastic Analysis of Heat-Treatment Including Creep and Phase
Changes. Computers and Structures, 13:771 - 779, 1981.

[150] A . Kumar Singh and D. Mazumdar. Comparison of Several Numerical Methods


for Thermal Fields during Phase Transformation of Plain Carbon Steels. ISIJ
International, 31(12):1441 - 1444, 1991.

[151] S. Denis, D. Farias, and A . Simon. Mathematical Model Coupling Phase Trans-
formationns and Temperature Evolutions in Steels. ISIJ International, 32(3):316
- 325, 1992.

[152] T. Inoue, T. Yamaguchi, and Wang Z. Stresses and Phase Transformations Ocur-
ring in Quenching of Carburized Steel Gear Wheel. Mater. Sci. Technoi, pages
899 - 903, October 1985.

[153] S. Sjostrom. Interactions and Constitutive Models for Calculating Quench Stresses
in Steel. Mater. Sci. Technoi, 1(10):823 - 829, October 1985.

[154] S. Denis, A . Simon, and G. Beck. Analysis of the Thermomechanical Behaviour


of Steel During Martensitic Quenching and Calculation of Internal Stresses. In
E. Macherauch and V . Hauk, editors, Eigenspannungen-Entstehung-Berechnung
Messung-Bewertung, pages 211 - 238. Deutsche Gesellschaft fur Metalkunde, 1983.

[155] J.T. Pindera. Comments on Modeling Plastic Deformation of Low Carbon Steel. In
A.S. Krausz, J.I. Dickson, J-P. A . Ammarigeon, and W . Wallace, editors, Constitu-
tive Laws of Plastic Deformation and Fracture, pages 279 - 284. Kluwer Academic
Publishers, Dordrecth, The Netherlands, 1990.

[156] J . M . Alexander. On Problems of Plastic Flow of Metals. In A . Sawczuk and


G. Bianchi, editors, Plasticity Today - Modelling, Methods and Applications, pages
683 - 714. Elsevier Applied Science Publishers, London, 1985.

[157] S.S. Hecker. Experimental Studies of Yield Phenomena in Biaxially Loaded Metals.
In J . A . Stricklinand K . J . Saczalski, editors, Constitutive Equations in Viscoplastic-
ity : Computational and Engineering Aspects, pages 1 - 33. The American Society
of Mechanical Engineers, 1976.
Bibliography 332

[158] W . Johnson and P.B. Mellor. Engineering Plasticity, page 87. Ellis Horwood
Limited, Chichester, England, 1st edition, 1983.

[159] W . F Chen and D . J . Han. Plasticity for Structural Engineers, page 311. Springer-
Verlag, New York, 1st edition, 1987.

[160] B.Jr. Hunsaker, W . E . Haisler, and J . A . Stricklin. On the Use of Two Hardening
Rules of Plasticity in Incremental and Pseudo Force Analysis. In J . A . Stricklin and
K . J . Saczalski, editors, Constitutive Equations in Viscoplasticity : Computational
and Engineering Aspects, pages 139 - 170. The American Society of Mechanical
Engineers, 1976.

[161] J . H . Weiner and J.V. Huddleston. Transient and Residual Stresses in Heat-Treated
Cylinders. J. of Applied Mechanics, 26(3):31 - 39, 1959.

[162] H . G . Landau. The Elastic-Plastic Plate under Cycles of Moving Dilations and
Small Applied Load. J. of the Mechanics and Physics of Solids, 11(2):97 - 117,
1963.

[163] T. Inoue and B . Raniecki. Determination of Thermal-Hardening Stress in Steels


by Use of Thermoplasticity Theory. J. Mech. Phys. Solids, 26:182 - 212, 1978.

[164] S. Denis, E . Gautier, A . Simon, and G. Beck. Stress-Phase Transformation Inter-


actions - Basic Principles, Modelling, and Calculation of Internal Stresses. Mater.
Sci. Technoi, 1(10):805 - 814, October 1985.

[165] B . Buchmayr and J.S. Kirkaldy. Modeling of the Temperature Field, Transfor-
mation Behaviour, Hardness and Mechanical Response of Low Alloy Steels during
Cooling from the Austenite Region. J. Heat Treat, 8(2): 127 - 136, 1990.

[166] A . M . Habraken. Coupled Thermo-Mechanical Analysis with Microstructural Com-


putation of Steel Pieces. In E . G . Thompson, R.D. Wood, and A . Zienkiewicz, O.C.
andd Samuelsson, editors, NUMIFORM 89, pages 165 - 170, 1989.

[167] S. Denis, S. Sjostrom, and A. Simon. Coupled Temperature, Stress, Phase Transfor-
mation Calculation Model Numerical Illustration of the Internal Stresses Evolution
during Cooling of a Eutectoid Carbon Steel Cylinder. Met. Trans. A, 18A:1203 -
1212, 1987.

[168] T. Inoue and Zhi-Gang Wang. Finite Element Analysis of Coupled Thermoinelas-
tic Problem with Phase Transformation. In J.F.T. Pittman, R . D . Wood, J . M .
Alexander, and O.C. Zienkiewicz, editors, Numerical Methods in Industrial Form-
ing Processes, pages 391 - 400, Swansea, U . K . , 1982. Pineridge Press Ltd.
Bibliography 333

[169] K . F . Wang, S. Chandrasekar, and H . T . Y . Yang. A n Efficient 2D Finite Element


Procedure for the Quenching Analysis With Phase Change. Trans, of the ASME,
115:124 - 138, Feb. 1993.

[170] G . W . Greenwood and R . W . Johnson. The Deformation of Metals under Small


Stresses during Phase Transformations. Proc. Roy. Soc, 283A:403 - 422, 1965.

[171] C.L. Magee and H.W. Paxton. The Microplastic Response of Partially Transformed
Fe-31Ni. Trans, of The Metallurgical Society of AIME, 242:1741 - 1749, August
1968.

[172] F . Abrassart. Influence des Transformations Martensitique sur les Propietes


Mechaniques des Alliages du Systeme Fe-Ni-Cr-C. PhD thesis, University of Nancy,
1972.

[173] Y . Desalos and F . Guinsberg. Report no. 902, IRSID, Germain-en-Laye, 1982.

[174] J . Giusti. Contraintes et Deformations Residuelles D'Origine Thermique : Appli-


cation au Soldage et a la Tempre des Aciers. PhD thesis, University of Paris, V I ,
1981.

[175] J . B . Leblond. Framatome internal report no. tm/c dc/80.066, Societe Framatome,
Saint-Marcel, 1980.

[176] J . B . Leblond, G. Mottet, J . Devaux, and J.-C. Devaux. Mathematical Models of


Anisothermal Phase Transformations in Steels, and Predicted Plastic Behaviour.
Mater. Sci. Technoi, 1(10):815 - 822, October 1985.

[177] B . G. Thomas, I. V . Samarasekera, and J . K . Brimacombe. Mathematical Model


of the Thermal Processing of Steel Ingots : Part II. Stress Model. Met. Trans. B,
18B(1):131 - 147, 1987.

[178] O.C. Zienkiewicz and L C . Cormeau. Visco-Plasticity and Creep in Elastic Solids
- A Unified Numerical Solution Approach. Int. J. for Numerical Methods in Engi-
neering, 8:821 - 845, 1974.

[179] H . G . Fjcer and A . Mo. A L S P E N - A Mathematical Model for Thermal Stresses in


Direct Chilled Casting of Aluminum Billets. Met. Trans. B, 21B:1049 - 1061, Dec.
1990.

[180] Y . Ebisu, K . Sekine, and M . Hayama. Analysis of Thermal and Residual Stresses
of a Low Alloy Cast Steel Ingot by the Use of Viscoplastic Constitutive Equations
Considering Phase Transformation. Testsu-to-hagane, 78(6) :50 - 57, 1992.
Bibliography 334

D . R . J . Owen and E . Hinton. Finite Elements in Plasticity : Theory and Practice,


chapter 8. Pineridge Press Ltd., Swansea, U . K . , 1st edition, 1980.

S. L. Cockroft. Thermal Stress Analysis of Fused-Cast Monofrax-S Refractories.


PhD thesis, University of British Columbia, 1990.

J. V . Beck. User's Manual for C O N T A - Program for Calculating Surface Heat


Fluxes From Transient Temperatures Inside Solids. Report sand83-7134, Sandia
National Laboratory, Dec. 1983.

M . R . Spiegel. Mathematical Handbook of Formulas and Tables, page 113. McGraw-


Hill Book Co., New York, 1st edition, 1968.

F . L . Stasa. Applied Finite Element Analysis for Engineers, pages 514 - 515. Holt,
Rinehart and Winston, New York, 1985.

T. Dupont, G . Fairweather, and J.P. Johnson. Three-Level Galerkin Methods for


Parabolic Equations. SIAM J. Numerical Analysis, 11(2):392 - 410, 1974.

R . L . Burden and J.D. Faires. Numerical Analysis, pages 409 - 413. Prindle, Weber
& Schmidt Publishers, Boston, 3rd edition, 1985.

P. C. Campbell. Application of Microstructural Engineering to the Controlled Cool-


ing of Steel Wire Rod. PhD thesis, University of British Columbia, 1989.

F. Kreith and Black. W . Z . Basic Heat Transfer, page 138. Harper & Row, New
York, 1980.

M . Necati Oziik. Heat Conduction, pages 101-102. John Wiley & Sons, New York,
1980.

J . V . Beck, B . Blackwell, and C R . St. Clair, Jr. Inverse Heat Conduction : III
Posed Problems, page 36. Wiley-Interscience, New York, 1985.

J . V . Beck, B . Litkouhi, and C R . St. Clair Jr. Efficient Solution of the Nonlinear
Inverse Heat Conduction Problem. Numerical Heat Transfer, 5:275-286, 1982.

M . Necati Oziik. Heat Conduction, page 586. John Wiley & Sons, New York, 2nd
edition, 1993.

B . Hernandez-Morales, S.M. Gupta, J . K . Brimacombe, and E . B . Hawbolt. Deter-


mination of Quench Heat-Transfer Coefficients Using Inverse Techniques. In G . E .
Totten, editor, 1st Int. Conf. on Quenching and Distortion Control, pages 155 -
164. A S M International, 1992.
Bibliography 335

C A . Muojekwu, I.V. Samarasekera, and J . K . Brimacombe. Heat Transfer and


Microstructure during the Early Stages of Metal Solidification. Metallurgical and
Materials Transactions B, 26B:361 - 382, 1995.

K . Ho and R . D . Pehlke. Metal-Mold Interfacial Heat Transfer. Metallurgical Trans-


actions B, 16B:585 - 594, 1985.

B . Hernandez-Morales, J . K . Brimacombe, and E . B . Hawbolt. Application of In-


verse Techniques to Determine Heat-Transfer Coefficients in Heat-Treating Oper-
ations. J. of Materials Engineering and Performance, 1(6):763 - 771, December
1992.

H.S. Carslaw and J . C Jaeger. Conduction of Heat in Solids, page 203. Oxford
University Press, 1946.

N . Lambert and M . Economopoulos. Measurement of the Heat-Transfer Coeffi-


cients in Metallurgical Processes. JISI, 208(10):917 - 928, 1970.

J.F. Besseling and E . van der Giessen. Mathematical Modelling of Inelastic Defor-
mation, page 175. Chapman & Hall, London, 1st edition, 1994.

Y . Yamada, N . Yoshimura, and T. Sakurai. Plastic Stress-Strain Matrix and Its


Application for the Solution of Elastic-Plastic Problems by the Finite Element
Method. Int. J. Mech. Sci., 10:343 - 354, 1968.

D.J. Johns. Thermal Stress Analyses, pages 35 - 38. Pergamon Press, Oxford, 1st
edition, 1965.

S.P. Timoshenko and J.N. Goodier. Theory of Elasticity, pages 443 - 445. McGraw-
Hill, New York, 3rd edition, 1970.

H . Buhler, H . Buchholz, and E . H . Schulz. Archiv. Eisenhiittenwes., 5:413, 1932.

W . Mitter, F . G . Rammerstofer, 0 . Grundler, and G . Wiedner. Discrepancies Be-


tween Calculated ans Measured Residual Stressses in Quenched Pure Iron Cylinder.
Mater. Sci. TechnoL, 1(10):793 - 797, October 1985.

O M E G A . O M E G A Complete Temperature Measurement Handbook and Encyclo-


pedia. Volume 28, 1992.

A.S. Foust, L . A . Wenzel, C W . Clump, L. Maus, and L . B . Andersen. Principles of


Unit Operations, page 162. John Wiley & Sons, New York, 1st edition, 1960.

V . L . Streeter and E . B . Wylie. Fluid Mechanics, pages 253 - 257. M c Graw-Hill


Book Co., New York, 8th edition, 1985.
Bibliography 336

[209] R . H . Notter and C A . Sleicher. A Solution of the Turbulent Graetz Problem


III. Fully Developed and Entry Region Heat Transfer Rates. Chemical Engineering
Science, 27:2073 - 2093, 1972.

[210] V . H . Hernandez-Avila. Heat Transfer Model of the Hot Rolling Runout Table-
Cooling and Coil Cooling of Steel. Master's thesis, University of British Columbia,
1994.

[211] R . C . Gifkins. Optical Microscopy of Metals, pages 178 - 186. American Elsevier
Publishing Company, New York, 1st edition, 1970.

[212] Co. Trucomp. T S M O O T H . Optimal Smoothing and Differentiation of data. The-


oretical and users manual, D Y N A C O M P , Inc, June 1993.

[213] H . Honda, H . Takamatsu, and H . Yamashiro. Heat-Transfer Characteristics during


Rapid Quenching of a Thin Wire in Water. Heat Transfer - Japanese Research,
21(8):773 - 791, 1992.

[214] W . H . Press, S.A. Teukolsky, W . T . Vetterling, and B.P. Flannery. Numerical


Recipes in FORTRAN, page 442. Cambridge University Press, Cambridge M A ,
2nd edition, 1992.

[215] British Iron and Steel Research Association. Physical Constants of Some Com-
mercial Steels at Elevated Temperatures, pages 3 - 14. Butterworth's Scientific
Publications, Guilford, Surrey, U . K . , 1953.

[216] American Society for Metals. ASM Handbook, volume 1, pages 145 - 149. A S M
International, Warrendale, P A , 9th edition, 1978.

[217] J . V . Beck and A . M . Osman. Analysis of Quenching and Heat Treating Processes
Using Inverse Heat Transfer Method. In G . E . Totten, editor, 1st Int. Conf. on
Quenching and Distortion Control, pages 147 - 153. A S M International, 1992.

[218] H . Honda, H . Takamatsu, and H . Yamashiro. Heat Transfer and Liquid Solid
Conatact during the Rapid Quenching of Thin Wires in Water and CaCLj /Water
Solution. In G . F . Hewitt, editor, Heat Transfer 1994, volume 5, pages 201 - 211.
Institution of Chemical Engineers, 1994.

[219] V . H . Hernandez-Avila. Private Communication.

[220] A . Nakayama and H . Koyama. Integral Treatment of Subcooled Forced Convection


F i l m Boiling on a Flat Plate. Warme und Stoffiibertragung, 20:121 - 126, 1986.

[221] American Society for Testing and Materials. The Hardness of Martensite, pages
310 - 316. A S T M , Philadelphia, P A , 1986.
Bibliography 337

[222] R. Schroder. Influences on Development of Thermal and Residual Stresses in


Quenched Steel Cylinders of Different Dimensions. Mater. Sci. TechnoL, 1(10):745
- 764, October 1985.

[223] Y . J . Park and F . B . Fletcher. Effects of Manganese, Chromium, and Molybde-


num on the Isothermal Transformation of Austenite in Eutectoid Steels. J. Heat
Treating, 4(3):247 - 252, June 1986.

[224] T . C . Hsu, F . Marsiglio, J . H . Root, and T . M . Holden. Effects of Multiple Scatter-


ing and Wavelength-Dependent Attenuation on Strain Measurements by Neutron
Scattering. J. of Neutron Research, 3(1):27 - 39, 1995.

[225] B . Carnahan, H . A . Luther, and J.O. Wilkes. Applied Numerical Methods, pages
560 - 561. John Wiley & Sons, New York, 1st edition, 1969.

[226] A . J . Fletcher and C. Lewis. Effect of Free Edge on Thermal Stresses in Quenched
Steel Plates. Mater. Sci. TechnoL, 1(10):780 - 785, October 1985.

[227] W . H . Press, S.A. Teukolsky, W . T . Vetterling, and B.P. Flannery. Numerical


Recipes in FORTRAN, pages 107-110. Cambridge University Press, Cambridge
M A , 2nd edition, 1992.

[228] A . Majorek, B . Scholtes, H . Muller, and E . Macherauch. The Influence of Heat


Transfer on the Development of Stresses, Residual Stresses and Distortions in
Martensitically Hardened S A E 1045 and S A E 4140. In G . E . Totten, editor, 1st
Int. Conf. on Quenching and Distortion Control, pages 171 - 179. A S M Interna-
tional, 1992.

[229] J . B . Wiskel. Thermal Analysis of the Startup Phase for D.C. Casting. PhD thesis,
University of British Columbia, 1995.

[230] F . Medina. Thermal and Microstructural Evolution of a Hot Strip on a Runout


Table. Master's thesis, University of British Columbia, 1992.

[231] O.C. Zienkiewicz and R . L . Taylor. The Finite Element Method, volume 1, pages
80-81. McGraw-Hill, London, 4th edition, 1989.

[232] M . Henriksen, D . B . Larson, and C . J . Van Tyne. Modeling Distortion and Residual
Stress in Carburized Steels. In G . E . Totten, editor, 1st Int. Conf. on Quenching
and Distortion Control, pages 213 - 212. A S M International, 1992.

[233] F . L . Stasa. Applied Finite Element Analysis for Engineers, page 423. Holt, Rine-
hart and Winston, New York, 1985.

[234] J . C . Russ. Practical Stereology, pages 21 - 22. Plenum Press, New York, 1986.
Bibliography 338

[235] J . Neter, W . Wasserman, and M . H . Kutner. Applied Linear Statistical Models,


page 142. Richard D. Irwin, Inc., Homewood, IL, 3rd edition, 1990.

[236] J . Neter, W . Wasserman, and M . H . Kutner. Applied Linear Statistical Models,


page 32. Richard D. Irwin, Inc., Homewood, IL, 3rd edition, 1990.

[237] D . W . Marquadt. A n Algorithm for Least Squares Estimation of Non-Linear Pa-


rameters. J. of the Society of Industrial and Applied Mathematics, 11:431 - 441,
1963.

[238] N . R . Draper. Applied Regression Analysis, pages 141 - 175. John Wiley and Sons,
New York, 2nd edition, 1981.

[239] J . Neter, W . Wasserman, and M . H . Kutner. Applied Linear Statistical Models,


page 419. Richard D. Irwin, Inc., Homewood, IL, 3rd edition, 1990.

[240] F . Boratto. Private Communication.


Appendix A

T h e r m o - M i c r o s t r u c t u r a l M o d e l : Finite Element Equations

The governing equation was found to be

1 d /, dT\ Id /, dT\ . 8T
(A.l)
r y r { k r
f r +
r ^ { k r
d z - + q = P P
^

which can be recast in an integral form by adopting the Galerkin method [233]

IA (kr) (kr dV = 0 (A.2)


r dr \ dr dz \ dz ,

which allows the approximation to be obtained element by element.


Considering
dV = 27rrdrd2; (A.3)

one obtains :

2TT / [ i V ] ^ (kr?P\ drdz + 2TT / [N] ^- (kr^f-) rdrd;


T T

JA e dr \ dr J JA dz \ dz J e

r r dT
+ 2TT / [N} q r drdz - 2n / [N] p C -drdz
T T
Pl =0 (A.4)
JAe JA e dt

Applying the Green-Gauss theorem to the terms that involve second-order derivatives

t dT r d dT
2TT / [N] kr^-n dC + 2ir /
T
^-[N} kr^-drdz
r
T

Jc e dr JAe dr dr
r dT r d dT
+2n / [Nfkrn dC + 2TT / \N} krdrdz z
T

Jc dz e JAe dz dz

339
Thermo-microstructural Model : Finite Element Equations 340

r / BT
+2TT / [N] q rdrdz - 2TT
T
/ [N) p C drdz
T
p =0 (A.5)
J A e JA e Bt

where n and n are director cosines of the normal to the surface.


r z

For a boundary condition of convection and/or specified heat flux, a heat balance at

the boundary is expressed as :

qn = qv- q C s (A.6)

where q cv is the heat transported by convection, q is an specified heat flux, and q , the s n

heat transported by conduction, is given by :

BT BT
qn = -k-^-n r - kn z ( A . 7 )
Br Bz

Using Newton's law of cooling :

q cv = -h(T a - T) (A.8)

Finally, by combining Eqs. (A.5) to (A.8), one obtains, at the element level : 1

[cy{ y + [Knay = uy
a (A.9)

where

[K] e
= [K ] rr
e
+ [K ] zx
e
+ [K ] cv
e
(A.10)

o r = { / j + UcvY e
+ uyq

1
The dimensions of the terms involved are : [C] = E L - 0 e 2 _ 1
, [K] = E L " ^
e - 1
T" 1

{/} = E L 2 T
e
r p - l
- 2
Thermo-microstructural Model : Finite Element Equations 341

and

\cy = 2TT f [N] p T


C r[N]drdz
p (A.ll)
JA E

[K ]
rr
e
= 2^ [ ^[N] k T
r[N]drdz (A.12)
JA' Or or

[K ] e
= 2TT [ ^-\N] k T
r^-[N]drdz (A.13)
JA oz oz
zz c

[K ]
cv
e
= 2TT / [N]h r drdz (A.14)
JC E

{UY = 2TT / [N]q r drdz (A.15)


JA E

{/J e
= 2TT / q r dC
s (A.16)

2TT [ hT r f dC (A.17)

A.l Isoparametric Elements and Numerical Integration

Isoparametric elements can be adopted to define the shape functions required in the

equations above. In this numerical integration (the most common being Gauss
quadrature) is required. In the following, the isoparametric representation as well as

expressions for the numerical integration of each term in Eqs. (A.9) and (A. 10) are
given.

Elemental Capacitance Matrix

Isoparametric representation :

[C] = 2TT *[Hc](t,vW*V


e
(A-18)

where

[Hc]((,v) = [N) p C r[N]\det J\


T
p (A.19)
Thermo-microstructural Model : Finite Element Equations

Numerical Integration :

[cr = 2KY.Y.^AHc]{^m)
3

Elemental Conductance Matrix (due to conduction)

Isoparametric representation :

f+l r+l
K]rr
e
= 2irJ i j ^ [H }{(,n)d(dn
rr

and
[K \ zz
e
= 2irj_^ [H]{t, )dtdri
V

where
[H m,v)-^ [N} kr-^[N}\detJ\
rr r
T

[H ]((,n)
zz = -^[N} kr-?-[N]\detJ\
T

Oz oz

Numerical Integration :

[K Y = rr 2nY J2 i A rr}(^,Vj)
/
U> U H

[K ] zz
e
= 2*Y.Y,^AH ]{^m) zz

Elemental Conductance Matrix (due to convection)

Isoparametric representation :

[K } cv
e
= 2n\H }(t,n)\ dti
cv 111
Thermo-microstructural Model : Finite Element Equations

where
[H }(t,r ) = [N} hr[N}^J?
cv ]
T
1 +J :

or
r+l

where

[H u](M
e = [N] hr[N]y/j?
T
1 + J?2

Numerical Integration :
[K ] =cv
e
27rJ2^[H }(Ci,l) cv

or

Elemental Load Vector (due to heat generation)

Isoparametric representation :

r + i /+!

where
{gM,v) = [N] qr\detJ\
T

Numerical Integration :

UiY = 2TT J2 i }{9g}(^


u u
Vi)
Thermo-microstructural Model : Finite Element Equations

Elemental Load Vector (due to convection)

Isoparametric representation :

r+i

where

{9cM,ri) = {N} h rTfy/jl T 2+ J:


22

or
e fr+i

{fcvY =
2n
J 1 i9cv}(l,n)dr]

where {9cv}(<i,n) = [N] hrTf \ j + T


J\

Numerical Integration :

or
{fcvY = 27r
I] i{^}(l,?7i)
w

Elemental Load Vector (due to heat flux)

Isoparametric representation :

{/
" /_r ^ '
}e=27r { }(e i)de

where
Thermo-microstructural Model : Finite Element Equations

or
r+i
{f } q
e
= 2TT {g }(l,n)dr,
q

where

{g }(l,v)
q = [N} q r^jI+j[
T
s 2

Numerical Integration :
{/J e
= 27r]C^J(6,i)

or
{fcvY = 2 7 r ^ o ; j {flfJ(l,7/ ) i
Appendix B

Thermo-Microstructural Model : Semi-Analytical Solution

A heat balance for a solid, long cylinder with a distributed heat source and exposed to
a fluid of heat transfer coefficient h and temperature Tf, assuming Newtonian behaviour
includes the following terms :

Heat evolved = qV

Heat flow in = 0

Heat flow out = -hA(T - T) f

.dT
Heat accumulated = pC V- p
~dt

where A and V are the area and volume, respectively, and q = Q (T). 0

Assembling these terms results in :

qV= hA(T -T)


f + pC V^- p (B.l)

A finite decrease in temperature, AT, during the time interval At can then be calcu-

lated from
rt+M = rT+AT d T

Jt pC V
P JT JiATf + qV-hAT 1
' '

Assuming that q ~ const during the time interval A t , the integral on the R.H.S. of

Eq. (B.2) results in :

rT+AT d T = i [6 + fl(r + Ar)]

JT hATj + qV -hAT a (b+aT) {


' ;

346
Thermo-Microstructural Transfer Model : Semi-Analytical Solution 347

where a = hA and b = hATf + qV


Also,

I
f*+ A i
dt 1
[(t + At) - t] (B.4)
* pC V p pC V p

Finally :
(6 + aT)exp(^f)-6
AT = ;B.5)

It should be noted that a = const while b = f(T). For the present case b is always
evaluated at the previous temperature.
Appendix C

Stress M o d e l : Finite Element Equations

The state of stress of a body can be computed by solving the equilibrium equations [203],
which relate the external forces per unit volume acting on the body, {&}, and the internal
force :
| ^ | + {6} = 0 onH (C.l)

A weak form of the equilibrium equations can be obtained by invoking the principle
of virtual displacements , as follows. Given that the set of equations, Eq. ( C . l ) , has to
1

be zero at each point of the domain, f i , it follows that

/o ({} + ) <">*<> = <


> ("J

where {v} T
is a set of arbitrary weighting functions.

For stress analysis, one chooses the arbitrary weighting function vector, {v} , T
to be
equal to a virtual displacement, {Su} :

{v}T
= {8u} T
= {8u, 8v, Sw} (C.3)

then

Jn

1
The Principle of Virtual Displacements is stated as : if the work done by the external forces on a
structural system is equal to the increase in strain energy of the system for any set of admissible virtual
displacements, then the system is in equilibrium.

348
Stress Model : Finite Element Equations 349

Integrating each term by parts (using Green's theorem) and rearranging

- {Int - Int ) + Int = 0


x 2 3 (C.5)

where, for a general three-dimensional continuum,

Inh = ^ ^ ^ ( ^ ) + --- + r ^ ^ ( < 5 u ) + ^ ( ^ ) ^ + - - - dV (C.6)

Int2 = J (5u b + 8v by + 8w b ) dV
x z

Int3 [5u(a n
x x + r yn
x y + Sw b ) + 5v(- ) + 5w(- )] dr
z

Recalling that the strains are related to the displacements by (for small strains)

{e} = [L]{u} (C.7)

then, the virtual strains are given by

{<5e} = [L]{8u} (C.8)

and

Inh = J {Se} {a} dV


v
T
(C.9)

Also, by defining the vector of traction forces acting per unit area as
{t} {a n x x +r n xy y -\ } (C.10)

then
InU J {5u} {t} dr
r
T
(c.ii)
Stress Model : Finite Element Equations 350

Therefore, the weak form of the equilibrium equations is expressed in matrix form as
2

J {Se} {a}dV-
v
T
J {Su} {b}dV
v
T
- J^{5u} {t}dT = 0
T
(C.12)

Note that this equation is valid for both linear and non-linear stress analysis.
Assuming the integral in Eq. (C.12) can be evaluated, one seeks the expressions for
the stiffness matrix and the load vector (at the element level) such that

[K] {aye
= {R} e
(C.13)

To this end, the displacements are approximated by :

{u} {u} = = [N]{a} e


(CU)
1

where [N] is the shape function matrix and {a} is the nodal displacement vector.
e

Also, the strains and nodal displacements are related by

{e} = [L]{u} = [L][N]{aY . (C.15)

Let us define the strain-nodal displacement matrix by

[B] = [L][N] (C.16)

then
{e} = [B]{ay (C.17)

It receives this name because the equilibrium equations contain second derivatives of the displace-
2

ments, whereas Eq. (C.12) contains only first derivatives.


Stress Model : Finite Element Equations 351

Given that [B] does not contain nodal displacements,

{5e} = [B]{Sa} e
(C.18)

and

{8e} T
= ({SaYf [Bf (C.19)

where {5a} is the vector of the virtual nodal displacements. Similarly,


e

{Su} = [N]{Sa} e
(C.20)

and

{8u} T
= ({SaYf [N] T
(C.21)

Substituting Eqs. (C.19) and (C.21) in Eq. (C.12) and rearranging,

({8ay) T
{ [B] {a} dV -
T
[Nf{b} dV - [N) {t} dA^ = 0
T
(C.22)

Since {8a} is nonzero , one obtains


e 3

/ [B] {a} dV = I [N) {b} dV + /


T T
[N] {t} dA
T
(C.23)
JV e JV e JAe

v ' s :
'

internal force loads

A general constitutive equation can now be invoked :


{a} = [D] ({e} - {e}) + {a } 0 (C.24)

For an elastic problem, the material property matrix, [D], is taken as [D], whereas for

3
Under the principle of virtual displacements {Sa} is arbitrary because it represents arbitrary virtual
e

displacements and it is nonzero.


Stress Model : Finite Element Equations 352

elastoplastic solutions, [D ] is adopted.


ep

Introducing Eq. (C.17) in Eq. (C.24) and substituting in Eq. (C.23) :

[J [B] [D][B]dv}{aY
v
T
= J [B] [D}{e}dV
v
T
(C.25)

+ / [N} {b}.dV
T

+ I [N] {t]
T
dA

- / [B] {a} T
dV

The last equation can be recast in matrix form as

[K] {a} e e
= {R} e
(C.26)

where

[tf] e
= / [B] [D][B]dVT
(C.27)

{RY = {R o}
e
e
+ {R }b
e
+ {R } t
e
- {R o} a
e
(C.28)

For the problem in hand we may ignore loads due to body forces, tractions, and initial
stresses. Then,
{RY = {R oY = / [B} [D}{e } dV

T 0
(C.29)
JVe

C.l Isoparametric Elements and N u m e r i c a l Integration

Isoparametric elements can be adopted to define the shape functions required in the
equations above. In this numerical integration (the most common being Gauss
quadrature) is required. In the following, the isoparametric representation as well as
expressions for the numerical integration of the previous equations, is given.
Stress Model : Finite Element Equations

Stiffness M a t r i x

[K] e
= 2TT / [B] [D][B]r dr dz T

JA e

w 2nr j [B] [D][B]drdz


T

J A e

Isoparametric representation :

where

[H ]({, ) K V = [B} [D][B}f\detJ\


T

Numerical Integration :

E l e m e n t a l L o a d Vector

{R} e
= 2TT / [B] [D]{e}r T
dr dz
JA e

27rf / [B] [D}{e}dr T


dz
JA e

Isoparametric representation :

{R oY = 2n*{Sc}(t,
V) d dr?

where
{g }({,n)
s = [B}[D}{e}f\detJ\
Stress Model : Finite Element Equations

Numerical Integration :'

i j

Elemental Internal Force Vector

{F}e
= 2TT / [B} {a}rdrdz
Tm+1

w 27rf / [B] {cr}drdz


Tm+1

JA e

Isoparametric representation :

{FY = 2 * / _ * y _ * V } ( > 7 ) ^dr?

where
{9F}(t,r,) = [B] {<r}\detJ\
Tm+1

Numerical Integration :

{Fy = 2nY,i:^{9Fm,V>)
Appendix D

Stress M o d e l : Nonmechanical Strains

In the following, explicit equations for calculating the nonmechanical strains (Eq. (6.2))
in the elastoplastic stresss model are given [47]. Since these expressions are applied in
the finite-element formulation, they are presented in incremental form.

Thermal Strain Increment

Ae th
= Ar + v>,(r-r )Ax 0 f c hi p.i)

where

a mix = J2<Xk(T)X k (D.2)


k

Transformation Strain Increment

As Y,(3 AX k k Sii (D.3)


k
L

Transformation Plasticity Strain Increment

A
4 = I **(2(1 - X ))a' AX k tJ k (D.4)
1
k

where
A'2-4 = 4 . 1 8 x l O - M P a
5 _ 1
(ferrite, pearlite, bainite) (D.5)

355
Stress Model : Nonmechanical Strains

and

K - 4 = 5.08xlO- MPa
2
5 _1
(martensite)

Strain Increment due to = Dijki(T,Xk)


AE
Ae AE
n

E n

where

E \y. =
m ^2 Ek{T) Xk

Strain Increment due to a = o- (T, Xk)


y y

Ae i:j = -a -cr-j-Acr^mix

where

and
0 if elastic
1 if plastic
Appendix E

Recalescence During Quenching of IF Steel Bars

Some of the measured temperature responses during forced convective quenching of 38.1
mm-dia. IF steel bars showed recalescence at high temperatures. To determine whether
this effect was due to the heat evolved during the austenite-to-ferrite transformation, the
temperature response during the quench, including the heat evolved due to the transfor-
mation, was simulated numerically. As a first step, the start and finish transformation
temperatures under continuous cooling conditions in tubular IF steel samples were de-
termined following the procedures outlined in Chapter 7. The cooling rates considered
ranged from 2 to 24.5 C s , which correspond to the cooling rates observed near the
_1

equilibrium austenite-to-ferrite transformation temperature (910 C) during quenching


in water flowing at 75 C. This water temperature was selected because it represents
an extreme case. The results are shown in Figure E . l . The transformation start tem-
perature varied from 892 to 873 C for cooling rates of 2 and 24.5 C s , respectively.
_1

The corresponding transformation finish temperature varied from 850 to 835 C. To


characterize the progress of the ferrite transformation, the fraction transformed was de-
termined from the dilatometric measurements, and correlated to the undercooling below
the transformation start temperature. For example, the fraction transformed obtained
with a cooling rate of 24.5 C s _1
is given by :

(E.l)

To model the evolution of the thermal field in the 38.1 mm-dia. bar during forced

357
Recalescence During Quenching of IF Steel Bars 358

convective quenching, the response at the subsurface was adopted as the heat-transfer
boundary condition. To simplify the problem, the fraction transformed was assumed to
follow E q . ( E . l ) , irrespective of the local cooling rate. The thermophysical properties of
IF steel given in Chapter 8 were adopted for the calculations.' The calculation domain
was discretized with 10 4-node isoparametric elements and a time increment, At, of 0.2 s
was adopted. The predicted temperature response at the centre and mid-radius of an IF
steel bar quenched in water flowing at 2.8 m s _1
at 75 C is shown in Figure E.2. The
broken lines in the figure delimit the temperature range where the transformation takes
place. The combination of the low surface heat fluxes obtained at high temperatures
(due to the presence of film boiling) and the heat generated by the austenite-to-ferrite
transformation resulted in recalescence at the centre. This effect was less pronounced at
mid-radius, due to the higher cooling rates experienced at that position. The simplified
mathematical model correctly predicted the measured form of the temperature response
at the centre (Figure 9.2); thus, it can be concluded that the recalescence observed in
the measured cooling curves was due to the transformation to ferrite.
Recalescence During Quenching of IF Steel Bars 359

1000

Figure E . l : Start (closed circles) and finish (open circles) temperatures during continuous
cooling of tubular IF steel specimens.

1000
start
o
900

u 800
o

co end
u
p 700 O
ns o
u
CO O 1
(X
600 o
g
co

500 Centre
O MidRadius

400

300
20 40 60

Time, s

Figure E.2: Predicted temperature response at the centre and mid-radius of a 38.1
mm-dia. IF steel bar quenched with water flowing at 2.8 m s at 75 C. The start _1

and end of the transformation are shown as broken lines.


Appendix F

Isothermal Tests : D a t a Reduction

F.l Linear Regression

A linear regression analysis of isothermal kinetic data can be performed after transforming
the independent and dependent variables in the Avrami equation as follows :

(F.l)

Then, from a plot of l n [ l n ( l / l X(t)] against ln(t tAv) the parameters n and In 6 can
be estimated by the slope and the intercept of the regressed line, respectively. Given that
only two parameters can be estimated with a linear regression, the start of the isothermal
transformation tAv has to be estimated by an iterative procedure, where guessed values
of tAv are adopted to estimate n and In b until the optimum value (which corresponds to
the maximum in a plot of correlation coefficient against IAV) is found.

A fundamental problem with this approach lies in the fact that transforming both
independent as well as dependent variables results in a change in the shape of the dis-
tribution of the error terms and also leads to variable error term variances [234,235]. In
this regard it should be pointed out that one of the statistical assumptions underlying
linear and non-linear regression analysis is that the error terms are assumed to have con-
stant variance and, therefore, the responses have also the same constant variance [236].
To illustrate this effect, exact values of fraction transformed were computed using the
Avrami equation with n = 2, In b = 11 and i ^ y = 0.5 s, and constant error bars in

360
Isothermal Tests : Data Reduction 361

the response, equivalent to 0.05 units, were added as shown in Figure F . l (a). When
the data is linearized according to Eq. ( F . l ) the error bars are no longer uniform for all
levels of the independent variable (see Figure F . l (b)) and, therefore, the assumption of
constant variance in the errors is no longer valid. From that figure it is also clear that
this effect is more pronounced at very low and very high values of fraction transformed.

F.2 N o n - L i n e a r Regression

Alternatively, a multiple non-linear regression analysis can be applied to the fraction


transformed vs time data; the Marquadt's algorithm [237] is commonly adopted for this
task. It should be pointed out that all three parameters, n,b, and t v, A can now be
estimated simultaneously. A multiple non-linear regression was performed on the same
exact data of Figure F . l (a), using the commercial package SigmaPlot v 5.0, and is shown
graphically in Figure F.2. Eventhough the general trend is correctly predicted, there are
large differences (up to almost 10 %) between observed and predicted values. In order
to establish the reason for this variation, an examination of the residuals (the difference
between observed and predicted values [238]) was performed. When the residuals are
plotted against the predicted values, a 'horizontal band' indicates no abnormality [238];
on the other hand, unsatisfactory residuals behaviour is characterized by departures from
this 'horizontal band'. A plot of residuals against predicted values for the regression on
the exact data is shown in Figure F.3; clearly, a departure from normal behaviour is
observed.

It was thought that the reason for the irregularity in the residuals behaviour was re-
lated to a non uniform variance of the data and, therefore, a weighted multiple non-linear
regression was attempted. The weighted least squares criterion as applied to multiple
Isothermal Tests : Data Reduction 362

regression is [239] :

Qv, = J2 i(Yi
w
- A> - PiXi,i - - / 3 p _ i ^ , P - i ) 2
(F.2)
i=l
where u>; is the weight associated with observation YJ. In general, an observation Y{
that has an small variance receives more weight than another observation that exhibits
a large variance; this is intuitively reasonable. When the variances are unknown, as was
the case, they can be estimated assuming them being proportional to either the inverse
of the observations or the inverse of the square of the observations. Figure F.4 (a) shows
a comparison of the results obtained for regressions performed with these methods as
well as the non-linear regression without weights. The use of weights in the regression
improved the results, with the non-linear regression with weights equal to the inverse
of the square of the observations resulting in an almost perfect fit. The corresponding
residuals for these regressions are shown in Figure F.4 (b); by assigning variable weights,
a dramatic improvement on the residuals behaviour has been accomplished.
Given that the Marquadt-Levenberg algorithm is iterative in nature, one must specify
initial parameter values for a non-linear regression analysis. Obviously, it is important
to ensure that the curve fitter converges for a reasonable range of initial parameters; this
result shall serve as a proof of the robustness of the algorithm. For the set of exact data
points studied, several regression analyses were carried out for the following ranges of
initial parameters : 1.4 < n < 2.6, -17 < In 6 < -5, tAV = 0.05 s. The results, summarized
in Table F . l , show a very good degree of robustness for the algorithm as applied to this
problem.

When a multiple non-linear regression analysis was applied to actual experimental


data, it was found that a more detailed description for the weights was necessary. The
exact form the the weighting function depends on the data being analyzed but a general
Isothermal Tests : Data Reduction 363

form is given by

1//CO 2
y < 0.05

0.05 < f(U) < 0.5


w.
io//(<0 (F.3)
10/(/(*0 - 0.5) 0.5 < /(t,-) < 0.95

-/CO) 2
/(<0 > 0.95

Note the symmetry in the weighting function, that reflects the almost symmetrical be-
haviour of f(t{).
As an example, the results of a non-linear regression analysis applied to experimental
data are shown in Figure F.5. A good fit was obtained by using weights in the non-linear
regresssion.
A weak point in the data analysis (using either linear or non-linear regression) is
the lack of a selection criterion for D and D . This problem can be overcome by
mm max * J

including these values as parameters in the multiple non-linear regression analysis [240] :

D{t) = A[l - exp(-b(t - t ) )} + B


AV
n
(F.4)

where A = D max Dmm and B = D .mm In this way, the values of A and B are optimized
through the non-linear regression; initial values are easy to estimate from inspection. As
before, the use of weighting functions was found necessary for most cases. A n example
of an application of this analysis to real data is shown in Figure F.6.
Isothermal Tests : Data Reduction 364

Table F . l : Initial values for n and In b used to test the robustness of the weighted
non-linear regression analysis (a yj denotes excellent performance). The initial value for
tAv was 0.05 s for all the cases.

n In b
-5 -7 -9 -13 -15 -17
1.4 V v 7
V X X

1.6 V V V V X X

1.8 V V V V X X

2.2 V V V V V X

2.4 V V V V
2.6 V V V V V V
Isothermal Tests : Data Reduction 365

~i 1 1 r

1.0
10
t 0.8

X 0 - 6 h

X
ea
I 0.4

0.2

o\t(t.)\ = a = 0.05
0.0 h 0 _j i_

0 100 200 300 400 500 600

Time, s

(a)

(b)

Figure F . l : (a) Exact data corresponding to n = 2, In 6 = 11 and t v = 0.5 s. The errorA

bars correspond to 0.05 units and are equal at all levels of t. (b) Exact data linearized
using Eq. (F.l).
Isothermal Tests : Data Reduction 366

o.i

0.0

d
3
i3
co
OS
-0.1

-0.2
0.0 0.2 0.4 0.6 0.8 1.0

f(t) (estimated)

Figure F.3: Residuals plot against estimated values corresponding to the non-linear re-
gression of Figure F.2.
Isothermal Tests : Data Reduction 367

1.0 h

O.B

f 0.6

x
I 0.4 O Data Points
w = 1.0
w = f(t)
0.2 w = f(t)

0.0

100 200 300 400 500 600

Time, s

(a)

0.100

0.075

0.050

0.025

"3 0.000
HJ

on
d> -0.025
CM
-0.050

-0.075

-0.100

-0.125
0.0 0.2 0.4 0.6 0.8 1.0

f(t) (estimated)

(b)

Figure F.4: Comparison of weighted and non weighted multiple non-linear regression
applied to exact data : (a) estimated and observed values, (b) residuals plot against
estimated values.
Isothermal Tests : Data Reduction 368

0 100 200 300 400 500

Time, s

Figure F.5: Multiple (3 parameter) weighted non-linear regression applied to experi-


mental data : estimated and observed values.

o Data
Fitted
d
d ) 0 OQOOO<~lOOOQ

5 0.035
d
CD

cS
->
<-l
Q
u 0.030
<A

d"
o
-rH
U
CO
Test : CIT03-1
i\
FH
T = 635 C
o

0.025
100 200 300 400 500

Time, s

Figure F.6: Multiple (5 parameter) weighted non-linear regression applied to experi-


mental data : estimated and observed values.

Você também pode gostar