You are on page 1of 336

The University of New South Wales

Faculty of Science
School of Materials Science and Engineering

Strain Ageing Of Low Carbon Steel


Wire Rods

A Thesis in
Materials Science and Engineering
By
Sujatha Mehta

Submitted in Partial Fulfilment


of the requirements for the Degree of
DOCTOR OF PHILOSOPHY

FEBRUARY 2003
CERTIFICATE OF ORIGINALITY

I hereby declare that this submission is my own work and to the best of my

knowledge it contains no materials previously published or written by another

person, nor any material which to a substantial extent has been accepted for the

award of any degree or diploma at UNSW or any other educational institution,

except where due acknowledgement is made in the thesis. Any contribution made

to the research by others, with whom I have worked at UNSW or elsewhere, is

acknowledged in the thesis.

I also declare that the intellectual content of this thesis is the product of my own

work, except to the extent that assistance from others in the project's design and

conception or in style, presentation and linguistic expression is acknowledged.

Signed-----

11
To my parents, husband Arun and lovely daughters, Shreya and Srishti

iii
ACKNOWLEDGEMENTS

I wish to express my deepest sense of gratitude and appreciation to my supervisor Dr.

Peter Krauklis, for his continuous support and encouragement throughout my research. I

also wish to express my gratitude to Associate Professor Brian Gleeson for his

unrelenting support and guidance. I would also like to sincerely thank Mr. Barry Jessop,

formerly Manager, BHP Newcastle, for his guidance during this research.

I wish to acknowledge the support of the Australian Research Council and BHP

Newcastle and BHP Laboratories, Melbourne for their funding and support towards

supply of test material and use of hot torsion testing facilities.

I would also like to place on record my sincere gratitude to the following people:

Dr. Alan Brownrigg, formerly Manager, BHP Laboratories, Melbourne, currently at

Deakin University; Mr Brendon Perrett for his unrelenting support during hot torsion

testing, the hot torsion testing team at Deakin University; John Sharp, Inna Bolkovsky,

John Budden of the School of Materials Science and Engineering; Associate Professor

Paul Munroe and Viera Piegorova from the UNSW Electron Microscopic Unit, for their

guidance during transmission electron microscopy work.

I would also like to thank Dr. Arslan Kaya, former postdoctoral fellow at the School of

Materials Science and Engineering for his unrelenting support throughout transmission

electron microscopy.

IV
Finally, I would like to thank my family- my parents, Arun and my lovely daughters

Shreya and Srishti for their constant support and inspiration during my research.

V
ABSTRACT

A major problem encountered during metal forming operations, in particular wire

drawing is strain ageing. This phenomenon manifests itself as an increase in strength

and a decrease in ductility due mainly to the diffusion of free nitrogen and/or carbon to

dislocations. These interstitials form atmospheres around dislocations, effectively

impeding further movement. At lower temperatures, nitrogen typically plays a greater

role in strain ageing than does carbon. A common method of reducing the free

nitrogen content in steel is to add alloying elements, which form stable nitrides. This

project was therefore aimed at studying the strain ageing phenomenon in low carbon

steels as a function of temperature, time and other process variables during wire

drawing. For this purpose, hot torsion laboratory teats were adopted to simulate

conditions during industrial wire drawing operations. The main focus of this study has

been to determine the effects of strain ageing on the mechanical properties and

therefore the drawability of low carbon steels. For completeness, the behaviour of EAF

(electric arc furnace) grades have been compared with the BOS (Basic Oxygen) grades.

Four grades of steels were studied, three of which were BOS grades and one, an EAF

grade. The compositional differences were mainly based on the nitrogen content and

the boron content. These steels were subjected to hot torsion tests at temperatures of

130°C and 200°C with holding times of 30 and 600 seconds after an initial prestrain.

The effect of these test parameters on the general stress-strain curve, work hardening

vi
rate, ageing index, strain to failure and maximum stress were deteremined as a function

of material composition.

Hot torsion tests revealed occurrence of strain ageing m these materials under

conditions typical of industrial wire drawing. However, the absence of precipitate

formation under test conditions led to a conclusion that the ageing phenomenon in these

steels was due to atmosphere formation around dislocations.

Hot torsion tests at a combination of 130°C and 200°C at delay times of 30 and 600

seconds have indicated that although the BOS high B+N grade shows a lower degree of

work hardening, the ageing index does not show an appreciable and consistent

improvement over the other grades. This result has further been confirmed by 'free

nitrogen' analyses in which only a small amount of boron has been found in the

combined form.

It is concluded that boron additions of 0.0012wt% and 0.0038wt% to steels of 0.006-

0.008 wt% N have not been effective in fixing all of the free nitrogen under the cooling

conditions of water quenching from 1100°C to 900°C and subsequent cooling to room

temperature at a cooling rate of 2°C/s.

vii
TABLE OF CONTENTS

PAGE

CERTIFICATE OF ORIGINALITY ii

ACKNOWLEDGEMENTS IV

ABSTRACT VI

TABLE OF CONTENTS Vlll

LIST OF FIGURES XI

LIST OF TABLES XXll

CHAPTER 1 INTRODUCTION

CHAPTER 2 PHYSICAL METALLURGY OF LOW CARBON


GRADE STEELS USED FOR WIRE/ROD PRODUCTS 6
2.1 STEELS COMMONLY USED FOR WIRE/ ROD PRODUCTS
AT BHP, NEWCASTLE 7
2.2 WIRE DRAWING - A GENERAL DESCRIPTION 7
2.3 PHYSICAL METALLURGY OF LOW CARBON GRADES USED FOR
WIRE/ROD PRODUCTS 9
2.3.1 Fe- Fe3C phase diagram 10
2.3.2 Transformations during continuous cooling 14
2.3.3 Effect of alloying elements 16
2.3.4 Classification of alloying elements in steel 16

CHAPTER 3 PRECIPITATION KINETICS 23


3.1 PRECIPITATION KINETICS 23
3.1.1 Diffusion 23
3.1.2 Nucleation of a solid in a liquid 31
3.1.3 Nucleation in solids 36
3.1.4 Experimental methods of monitoring precipitation kinetics 42

CHAPTER 4 STRAIN AGEING OF LOW CARBON STEELS 53


4.1 GENERAL FEATURES 54
4.2 THEORIES OF STRAIN AGEING 59
4.2.1 Atmosphere formation 62
4.2.2 Precipitation on dislocations 65
4.2.3 Effect of dislocation sources 71
4.2.4 Factors affecting strain ageing 72
4.2.5 Static Strain Ageing 73
viii
PAGE
4.2.6 Dynamic Strain Ageing 119
4.2. 7 Strain dependence of Static and Dynamic ageing 125

CHAPTER 5 RESEARCH OBJECTIVES 127

CHAPTER 6 EXPERIMENTAL METHODS 131


6.1 STEEL COMPOSITION AND PROCESSING 132
6.2 ROOM TEMPERATURE MECHANICAL PROPERTIES 132
6.3 MICROSTRUCTURAL CHARACTERISATION 132
6.3 .1 Optical microscope 132
6.3.2 Transmission electron microscope characterisation {TEM) 133
6.4 ELECTRICAL RESISTIVITY 134
6.5 HOT TENSION TESTS 134
6.6 HOT TORSION TESTS 135

CHAPTER 7 RESULTS 139


7.1 OPTICAL MICROSCOPIC EXAMINATION 140
7.2 DETERMINATION OF "FREE NITROGEN" CONTENT IN THE STEEL
GRADES 143
7.3 ROOM TEMPERATURE MECHANICAL PROPERTIES 143
7.4 TRANSMISSION ELECTRON MICROSCOPIC EXAMINATION 144
7.4 HOT TENSILE TESTS 145
7.6 HOT TORSION TESTS 146
7 .6.1 Consistency of hot torsion tests 148
7 .6.2 Hot torsion test results 148
7.6.3 Results of transmission electron microscopic examination 176

CHAPTER 8 DISCUSSION 180


8.1 OPTICAL MICROSCOPY EXAMINATION 181
8.2 ROOM TEMPERATURE MECHANICAL PROPERTIES 181
8.3 HOT TORSION PROPERTIES 182
8.3.1 Influence of temperature on the stress -strain behaviour of the
four steel compositions 182
8.3.2 Influence of delay time after prestrain on the stress - strain
behaviour of the four steel compositions 208
8.3.3 Influence of material composition 227
8.4 TRANSMISSION ELECTRON MICROSCOPIC EXAMINATION 255

CHAPTER 9 CONCLUSIONS 256


9.1 USE OF HOT TORSION TESTING AS A SIMULATION OF
THE WIRE ORA WING PROCESS 257
9.2 INFLUENCE OF VARIOUS PARAMETERS ON THE STRAIN
AGEING BEHAVIOUR 257
9.2.1 Influence of temperature 257
9.2.2 Influence of delay time between prestrain and strain to failure 259
ix
PAGE

9.2.3 Influence of material composition 260

REFERENCES 265

APPENDIX A 275

APPENDIXB 279

APPENDIXC 296

X
List of Figures

LIST OF FIGURES

Figure 1.1 Load elongation curve for a low carbon steel strained to
point A, unloaded and then restrained immediately-curve
(a) or after ageing-curve(b).
Figure 2.1 Fe-Fe3C equilibrium phase diagram
Figure 2.2 (a) Body centred cubic structure
Figure 2.2 (b) Face centred cubic structure
Figure 2.3 Relationship between the Fe-Fe 3C and the CCT curves
0and the structural transformations resulting from various
cooling curves in steels containing a) 0.8%C b) 0.45%C
and c) 1.0%C
Figure 2.4 Effect ofsubstitutional alloying elements on the hardness of
ferrite
Figure 2.5 Effect of chromium on the ferrite region
Figure 2.6 Effect of molybdenum on austenite region
Figure 2.7 Solubility of nitrogen in different phases ofsteel

Figure 3.1 Free energy and chemical potential changes during


diffusion - (a) and (b) down hill diffusion and (c) and (d)
uphill diffusion
Figure 3.2 Interstitial positions in FCC and BCC structures
Figure 3.3 Interstitial atom in (a) equilibrium position (b) at the
position of maximum lattice distortion (c) Variation of.free
energy of the lattice as a function of the position of the
interstitial
Figure 3.4 Variation offree energy with embryo radius
Figure 3.5 Variation of r* and rmax with the amount of undercooling
Figure 3.6 Variation of L1G and N with undercooling for homogenous
and heterogenous nucleation
Figure 3.7 %Bversus %N
Figure 3.8 Change in resistance during precipitation
Figure 3.9 The peak strain as a function of the holding time at
temperature before deformation. The first plateau
indicates precipitation start time and the second, the finish
time
Figure 3.10 A precipitation time -temperature diagram for as obtained
from data in figure 2.16
Figure 3.11 Flow curves in terms of equivalent stress versus equivalent
surface strain for a low carbon steel determined at 600 'C
at nine strain rates
Figure 3.12 Flow curves in terms of equivalent stress versus equivalent
surface strain for a low carbon steel determined at 600 'C
at nine strain rates

XI
List of Figures

Figure 4.1 Load elongation curve for a low carbon steel strained to
point A, unloaded and then restrained immediately-curve
(a) and after ageing-curve(b).
Figure 4.2 The effect of ageing on the properties of mild steel temper
rolled 1.4%
Figure 4.3 Schematic diagram showing potential energy of a solute
atom as a function of the distance R from the centre of a
dislocation
Figure 4.4 Four different stages ofstatic strain ageing
Figure 4.5 Ferrite field showing equilibrium between Fe-N2 gas, Fe-
Fe,,N and Fe- Fe 1~2
Figure 4.6 Solubility of carbon and nitrogen as a function of
temperature
Figure 4.7 Internal friction curves of a dual solid solution of carbon
and nitrogen in ferrite vs temperature. Curves a and b are
the internal nitrogen and carbon peaks.
Figure 4.8 Solubility of nitrogen in equilibrium with dislocations and
precipitated nitride as a function of temperature
Figure 4.9 Diffusion coefficients of nitrogen and carbon
Figure 4.10 Strain ageing of iron -nitrogen and iron carbon alloys
strained 10% and aged for 2 hours at different
temperatures
Figure 4.11 Strain ageing of 0. 029%nitrogen-iron alloy followed by
change in the yield stress(upper curves) and solute content
offerrite as measured by internal friction
Figure 4.12 Increase in the yield stress after prestraining 3% and 8% vs
ageing time at 60 °C (Josefsson and Backstrom(5BJ ).
o-----o 0.02 C, 0.003 N; ---.0.03C0.008N; L1--
L10. 03C0. 008N
Figure 4.13 Increase in the yield stress after prestraining 3% and 8%
vs ageing time at 60 °C o-----o 0.02 C, 0.003 N; ---.0.03C
0.008N; Lt-- L10.03C 0.008N
Figure 4.14 Effect of nitrogen content on the flow stress of low carbon
steel aged at 100 °C for 20 seconds after 6% strain
Figure 4.15 Effect of nitrogen content on decrease in total elongation
due to ageing
Figure 4.16 Effects of (a) nitrogen level in solution and (b) carbon level
in solution on the change in yield stress, on prestraining
and ageing for 1hour at 100C using the following
prestrains Stephenson 10%; Enrietto 12%; Seking,
Fujishima , Butler 7.5% . Dissolved carbon and nitrogen
were determined by internal friction
Figure 4.17 Effect of level of manganese in solid solution on the rate of
strain ageing of commercial rimmed steels prestrained
10% in tension

XII
List of Figures

Figure 4.18 Stress strain curves at 2rfC and 205°C for (b) rimmed steel
which showed pronounced strain ageing and (c) aluminium
killed steel which showed no strain ageing
Figure 4.19 Influence of A/IN ratio on the solute nitrogen after
continuous annealing at 825C
Figure 4.20 Fe-B Phase diagram
Figure 4.21 Diffusion coefficient of boron (firm line) as compared to
carbon (dotted line)
Figure 4.22 Equilibrium phase diagram calculated for Fe-B-N system
Figure 4.23 Typical spherulite morphology of BN in steel
Figure 4.24 Precipitates in a furnace cooled steel from 80rfC after
heating at 130rfC for one hour and hot rolling. A/N grows
radially on BN
Figure 4.25 Schematic illustration ofprecipitation behaviour during the
manufacturing process
Figure 4.26 Type ofprecipitate formed under stee/making conditions
Figure 4.27 Influence of BIN ratio on the elongation
Figure 4.28 Stress strain curves for (a) low carbon steel 0.005%
nitrogen and (b) addition of0.06% vanadium
Figure 4.29 Precipitation temperatures of titanium compounds
Figure 4.30 Mechanical properties of rods containing titanium
additions
Figure 4.31 Effect of titanium on the ageing index of the wire
Figure 4.32 Effect of excess titanium on the tensile strength of low
carbon steel
Figure 4.33 A normal stress -strain curve (2rfC) and corresponding
curves illustrating dynamic strain ageing as a function of
temperature
Figure 4.34 Force-velocity diagram for a mobile dislocation
Figure 6.1 Typical dimensions of a hot torsion sample.
Figure 6.2 Test schedule used for hot torsion tests
Figure 7.1 Optical microstructure ofBOS low nitrogen steel
Figure 7.2 Optical microstructure ofBOS high nitrogen steel
Figure 7.3 Optical microstructure ofEAF low boron steel
Figure 7.4 Optical microstructure of BOS high boron steel
Figure 7.5 (a) TEM micrograph of a BOS grade high nitrogen steel
Figure 7.5 (b) TEM micrograph of a BOS high boron steel
Figure 7.6 Consistency in hot tension tests
Figure 7.7 A typical true stress-strain curve derived from the hot
torsion data
Figure 7.8 Consistency of hot torsion results on EAF steel containing
low boron and high nitrogen, when tested at 200C and a
delay time of 30 seconds after prestrain.
Figure 7. 9 (a) Error band calculated for the stress values
Figure 7.9 (b) Error band calculated for the strain values.

Xlll
List of Figures

Figure 7.10 Stress-strain curves of BOS low N steel as a function of test


temperature tested after a delay time of 30 seconds after
prestrain.
Figure 7.11 Stress-strain curves of BOS high N tested as a function of
temperature tested after a delay time of 30 seconds after
prestrain.
Figure 7.12 Stress-strain curves of EAF low B steel as a function of
temperature tested after a delay time of 30 seconds after
prestrain.
Figure 7.13 Stress-strain curves of BOS high N+B steel as a function of
temperature tested after a delay time of 30 seconds after
prestrain.
Figure 7.14 Stress-strain curves of BOS low N steel as function of
temperature tested after a delay time of 600 seconds after
prestrain.
Figure 7.15 Stress-strain curves of EAF low B steel as a function of
temperature tested after a delay of 600 seconds after
prestrain
Figure 7.16 Stress-strain curves of BOS high N steel as a function of
temperature tested after a delay time of 600 seconds after
prestrain
Figure 7.17 Stress-strain .curves of BOS high N+B steel as a.function of
temperature tested after a delay of 600 seconds after
prestrain.
Figure 7.18 Stress-strain curves of BOS low N steel as a function of
delay time after prestrain tested at l 30°C.
Figure 7.19 Stress-strain curves of BOS high N steel as a function of
delay time tested at J30°C
Figure 7.20 Stress-strain curves of EAF low B steel as a function of
delay time tested at l 30°C.
Figure 7.21 Stress-strain curves of BOS high N+B steel as a function of
delay time after prestrain tested at J3a°C
Figure 7.22 Stress-strain curves of BOS low N steel as a function of
delay time after prestrain tested at 200°C
Figure 7.23 Stress-strain curves of BOS high N steel as a function of
delay time after prestrain tested at 200°C
Figure 7.24 Stress-strain curves of EAF low B steel as a function of
delay time after prestrain tested at 200°C
Figure 7.25 Stress-strain curves of BOS high N+B steel as a function of
delay time after prestrain tested at 200°C
Figure 7.26 Stress-strain curves of the four material compositions
tested at 130 °c and 30 seconds delay time after prestrain
Figure 7.27 Stress-strain curves of the four material compositions
tested at l 30°C and a delay time of 600 seconds after
prestrain

XIV
List of Figures

Figure 7.28 Stress-strain curves of the four material compositions


tested at 200°C and a delay time of 30 seconds after
prestrain.
Figure 7.29 Stress-strain curves offour material compositions tested at
200°C and a delay time of 600 seconds after prestrain.
Figure 7.30 Transmission electron micrograph showing a very high
density of dislocations and a matrix devoid ofprecipitates
in BOS low nitrogen grade ofsteel tested at 20rf C after a
delay of 600 seconds.
Figure 7.31 Transmission electron micrograph showing a very high
density of dislocations and a matrix devoid ofprecipitates
in BOS high nitrogen grade of steel tested at 20rf C after a
delay of 600 seconds.
Figure 7.32 Transmission electron micrograph showing a very high
density of dislocations and a matrix devoid ofprecipitates
in EAF low boron grade ofsteel tested at J3rfC after a
delay of 600 seconds.
Figure 7.33 Transmission electron micrograph showing a very high
density of dislocations and a matrix devoid ofprecipitates
in BOS high nitrogen+ boron grade of steel tested at J3()°C
after a delay of 600 seconds.
Figure 7.34 Transmissio11; electron micrograph showing a very high
density of dislocations and a matrix devoid ofprecipitates
in BOS high nitrogen+ boron grade ofsteel tested at 200°C
after a delay of 30 seconds.
Figure 7.35 Transmission electron micrograph showing a very high
density of dislocations and a matrix devoid ofprecipitates
in BOS low nitrogen grade ofsteel tested at 20a°C after a
delay of 600 seconds.
Figure 8.1 Comparison of work hardening exponents in the prestrain
region as a function of temperature tested at 30 seconds
delay after prestrain
Figure 8.2 Comparison of work hardening exponents obtained in the
prestrain region as a function of temperature after 600
seconds delay after prestrain
Figure 8.3 Stress strain behaviour as a function of temperature BOS
Low N tested at 30 seconds delay after prestrain
Figure 8.4 Stress-strain. behaviour as a function of temperature of
BOS High N steel tested at 30 seconds delay after prestrain
Figure 8.5 Stress-strain behaviour as a function of temperature of
EAF low B steel tested at 30 seconds delay after prestrain
Figure 8.6 Stress-strain behaviour as a function of temperature of
BOS high N+ B steel tested at 30 seconds delay after
prestrain

xv
List of Figures

Figure 8.7 Stress-strain behaviour as a function of temperature of


BOS Low N steel tested at 600 seconds delay after
prestrain
Figure 8.8 Stress-strain behaviour as a function of temperature of
BOS high N steel tested at 600 seconds delay after
prestrain
Figure 8.9 Stress-strain behaviour as a function of temperature of
EAF Low B steel tested at 600 seconds delay after prestrain
Figure 8.10 Stress-strain behaviour as a function of temperature of
BOS high N+ B steel tested at 600 seconds delay after
prestrain
Figure 8.11 Comparison of maximum stress values obtained in the
prestrain region as a function of temperature
Figure 8.12 Comparison of maximum stress values in the strain to
failure region as a function of temperature after a delay of
30 seconds after prestrain
Figure 8.13 Comparison of maximum stress values as a function of
temperature after a delay of 600 seconds after prestrain
Figure 8.14 Comparison of total strains to failure as a function of
temperature tested at 30 seconds delay after prestrain
Figure 8.15 Comparison of total strains to failure as a function of
temperature tested at 600 seconds delay after prestrain
Figure 8.16 Comparison of ageing indices as a function of temperature
at 600 seconds delay after prestrain
Figure 8.17 Comparison of work hardening exponents in the prestrain
region as a function ofdelay time after prestrain at 13V°C
Figure 8.18 Comparison of work hardening exponents in the prestrain
region as a function of delay time after prestrain at 20V°C
Figure 8.19 Stress-strain behaviour of BOS low N as a function of delay
time after prestrain tested at J3V°C
Figure 8.20 Stress-strain behaviour of BOS high N as a function of
delay time after prestrain tested at 13V°C
Figure 8.21 Stress-strain behaviour of EAF low B steel as a function of
delay time tested at 13V°C
Figure 8.22 Stress-strain behaviour of BOS High N+ B steel as a
junction of delay time after prestrain tested at 13V°C
Figure 8.23 Stress-strain behaviour of BOS Low N as a function of
delay time after prestrain tested at 20V°C
Figure 8.24 Stress-strain· behaviour of BOS High N as a function of
delay time after prestrain tested at 20V°C
Figure 8.25 Stress-strain behaviour of EAF low B steel as a function of
delay time after prestrain tested at 20V°C
Figure 8.26 Stress-strain behaviour of BOS high N+B steel as a
function of delay time after prestrain tested at 20V°C

XVI
List of Figures

Figure 8.27 Comparison of maximum stress values in the strain to


failure region as a function of delay time after prestrain
tested at 13D°C
Figure 8.28 Comparison of maximum stress values in the strain to
failure region as a function of delay time after prestrain
tested at 20D°C
Figure 8.29 Comparison of values of total strain to failure as a function
of delay time after prestrain tested at 13D°C
Figure 8.30 Comparison of values of total strain to failure as a function
of delay time after prestrain tested at 20D°C
Figure 8.31 Comparison of ageing indices as a function of delay time
after prestrain tested at 13D°C
Figure 8.32 Comparison of ageing indices as a function of delay time
after prestrain tested at 20D°C
Figure 8.33 Stress-strain behaviour of the four different steel
compositions tested at 13D°C and 30 seconds delay after
prestrain
Figure 8.34 Stress-strain behaviour of the four steel compositions tested
at 13D°C and 600 seconds delay after prestrain
Figure 8.36 Stress strain behaviour of the four steel compositions tested
at 20D°C and 30 seconds delay after prestrain
Figure 8.37 Stress strain behaviour of the four steel compositions tested
at 20D°C and 600 seconds after prestrain
Figure 8.38 Variation of work hardening exponents (prestrain region)
and interstitial content for the four different steel
compositions tested at 13D°C and 30 seconds
Figure 8.39 Variation of work hardening exponents (prestrain) and
interstitial content for the different steel compositions
tested at 13D°C and 600 seconds
Figure 8.40 Variation of work hardening exponents (prestrain) and
interstitial content of four different steel compositions
tested at 20D°C and 30 seconds
Figure 8.41 Variation of work hardening exponents (prestrain) and
interstitial content of the four steel compositions tested at
200°C 600 seconds
Figure 8.42 Maximum stress at constant strain (1. 08) for the four steels
at 13D°C and 30 seconds
Figure 8.43 Maximum stress at constant strain (1. 08) for the four steels
at 13D°C and 600 seconds
Figure 8.44 Maximum stress at constant strain (1. 08) for the four steels
at 20D°C and 30 seconds
Figure 8.45 Maximum stress at constant strain (1. 08) for the four steels
at 20D°C and 30 seconds
Figure 8.46 Variation of interstitial content and ageing index for the
four steel compositions tested at 13D°C and 30 seconds

xvu
List of Figures

Figure 8.47 Variation of interstitial content and ageing index for the
four steel compositions tested at I 30°C and 600 seconds
Figure 8.48 Variation of interstitial content and ageing index for the
four steel compositions tested at 20V°C and 30 seconds
Figure 8.49 Variation in the interstitial content and ageing index for the
four steel compositions tested at 20V°C and 600 seconds.
Figure A-I Hot tensile curves of EAF Low B steel tested at I 3V°C after
a delay of 600 seconds.
Figure A-2 Hot tensile curves of EAF Low B steel tested at I 3V°C after
a delay of 30 minutes.
Figure A-3 Hot tensile curves of BOS high N steel tested at J3(fC after
a delay of 600 seconds.
Figure B-1 Stress-strain curve of BOS Low N tested at 130°C for a
delay time of 30 seconds after prestrain.
Figure B-2 Stress-strain curve of BOS Low N steel tested at I 30°C
after a delay time of 600 seconds after prestrain
Figure B-3 Stress-strain curve for a BOS Low N steel tested at 200°C
and 30 seconds after prestrain.
Figure B-4 Stress-strain curve for a BOS Low N steel tested at 200°C
and 30 seconds after prestrain
Figure B-5 Stress-strain curve of a BOS Low N steel tested at 200°C
after a delay time of 600 seconds after prestrain..
Figure B-6 Stress-strain curve of a BOS High N steel tested at I 30°C
after a time delay of 30 seconds after prestrain
Figure B-7 Stress-strain curve of a BOS high N steel tested at J30°C
after a time delay of 600 seconds after prestrain.
Figure B-8 Stress-strain curve of a BOS high N steel tested at 200°C
after a delay time of 30 seconds after prestrain
Figure B-9 Stress-strain curve of a BOS high N steel tested at 200°C
after 600 seconds delay after prestrain.
Figure B-10 Stress-strain curve of an EAF low B steel tested at 130°C
after a time delay of 30 seconds after prestrain.
Figure B-11 Stress-strain curve of an EAF low B steel tested at J30°C
after a time delay of 600 seconds after prestrain
Figure B-12 Stress-strain curve of an EAF low B steel tested at 200°C
after a delay time of 30 seconds after prestrain
Figure B-13 Stress-strain curve of an EAF low N+B steel tested at
200°C after a time delay of 600 seconds after prestrain.
Figure B-14 Stress-strain curve of a BOS high N+B steel tested at
I 30°C after a delay time of 30 seconds after prestrain
Figure B-15 Stress-strain curve of a BOS high N+B steel tested at
130°C after a delay of 600 seconds after prestrain.
Figure B-16 Stress-strain curve of a BOS high N+B steel tested at
200°C after a delay of 30 seconds after prestrain.
Figure B-17 Stress-strain curve of a BOS high N+B steel tested at
200°C after a delay of 600 seconds after prestrain

xvm
List of Figures

Figure C-1 Work hardening exponent in the prestrain region of the


stress strain curve of BOS low N tested at 130°C and 30
seconds delay time after prestrain.
Figure C-2 Work hardening exponent in the prestrain region of the
stress strain curve of BOS high N steel tested at J30°C
after 30 seconds delay after prestrain.
Figure C-3 Work hardening exponent in the prestrain region of strain
curve of EAF low B tested at 130°C and 30 seconds delay
time after prestrain
Figure C-4 Work hardening exponent in the prestrain region of the
stress strain curve of BOS high N+ B tested at 130°C and
30 seconds delay time after prestrain.
Figure C-5 Work hardening exponent in the prestrain reton of the
stress strain curve of BOS low N tested at 130 C and 600
seconds delay time after prestrain.
Figure C-6 Work hardening exponent in the prestrain region of the
stress strain curve of BOS high N tested at 130OC and a
time delay of 600 seconds after prestrain
Figure C-7 Work hardening exponent in the prestrain region of the
stress strain curve of EAF low B tested at 130°C after a
time delay of 600 seconds after prestrain
Figure C-8 Work hardening exponent in the prestrain region of the
stress strain curve of BOS high N+B tested at J30°C and a
time delay of 600 seconds after prestrain
Figure C-9 Work hardening exponent in the prestrain region of the
stress strain curve of BOS low N tested at 200°C after a 30
seconds delay after prestrain
Figure C-10 Work hardening exponent in the prestrain reton of the
stress strain curve of BOS high N tested at 200 C after 30
seconds delay after prestrain
Figure C-11 Work hardening exponent in the prestrain region of the
stress strain curve of EAF low B tested at 200 C after 30
seconds delay after prestrain
Figure C-12 Work hardening exponent in the prestrain region of the
stress strain curve of BOS high N+B tested at 200 C after
30 seconds delay after prestrain
Figure C-13 Work hardening exponent in the prestrain region of the
stress strain curve of BOS low B tested at 200°C after 600
seconds delay after prestrain
Figure C-14 Work hardening exponent in the prestrain region of the
stress strain curve of BOS high N tested at 200°C after 600
seconds delay after prestrain
Figure C-15 Work hardening exponent in the prestrain region of the
stress strain curve of EAF low B tested at 200°C after 600
seconds delay after prestrain

xix
List of Figures

Figure C-16 Work hardening exponent in the prestrain region of the


stress strain curve of BOS high N+B tested at 200°C after
600 seconds delay after prestrain
Figure C-17 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of BOS low N tested at
130°C and 30 seconds delay after prestrain
Figure C-18 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of BOS high N tested at
J30°C and 30 seconds delay after prestrain
Figure C-19 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of EAF low B tested at
J30°C and 30 seconds delay after prestrain
Figure C-20 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of BOS high N+B tested at
J30°C and 30 seconds delay after prestrain.
Figure C-21 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of BOS low N tested at
0 .
130 C and 600 seconds delay after prestrain
Figure C-22 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of BOS high N tested at
130°C and 600 seconds delay after prestrain
Figure C-23 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of EAF low B tested at
J30°C and 600 seconds delay after prestrain
Figure C-24 Work hardening exponent in a portion of 'strain to failure'
region of the stress-strain curve of BOS high N+B tested at
J30°C and 600 seconds delay after prestrain
Figure C-25 Work hardening exponent in 'strain to failure region of the
stress-strain curve of BOS low N tested at 200 C and 30
seconds delay after prestrain
Figure C-26 Work hardening exponent in the 'strain to failure 're§_ion of
the stress-strain curve of BOS high N tested at 200 C and
30 seconds delay after prestrain.
Figure C-27 Work hardening exponent in the 'strain to failure 're§.ion of
the stress-strain curve of EAF low B tested at 200 C and
30 seconds delay after prestrain.
Figure C-28 Work hardening exponent in the 'strain to failure 'region of
the stress-strain curve of BOS high N+B tested at 200°C
and 30 seconds delay after prestrain
Figure C-29 Work hardening exponent in the 'strain to failure 're§.ion of
the stress-strain curve of BOS low N tested at 200 C and
600 seconds delay after prestrain
Figure C-30 Work hardening exponent in the 'strain to failure 're§_ion of
the stress-strain curve of BOS high N tested at 200 C and
600 seconds delay after prestrain

XX
List of Figures

Figure C-31 Work hardening exponent in the 'strain to failure 'region of


the stress-strain curve of EAF low B tested at 200 C and
600 seconds delay after prestrain
Figure C-32 Work hardening exponent in the 'strain to failure 'region of
the stress-strain curve of BOS high N+ B tested at 200°C
and 600 seconds delay after prestrain

-xxi
List of Tables

LIST OF TABLES

Table 2.1 Commonly used grades of low and high carbon steels in
BHP, Newcastle.
Table 4.1 Measurements of strain ageing kinetics in slowly cooled or
quench-aged Fe-C and Fe-N alloys(J.D Baird35J)
Table 4.2 Measurements of strain ageing kinetics in quenched Fe-C and
Fe-N alloys (J.D Baird35J)
Table 4.3 Distribution of boron atoms as observed by Watanabe et a/<74·
75)

Table 4.4 Typical steel compositions studied by Morgan and Shyne (79J_
Table 4.5 Mechanical properties of the steels listed in Table 4.4
(Morgan and Shyni79J).
Table 7.1 Image analysis
Table 7.2 Nitrogen analysis of steel grades after hot extraction with
hydrogen.
Table 7.3 Room temperature mechanical properties of steels used in
current study.
Table 8.1 Values of work hardening exponents of the four different steel
compositions under various test parameters.
Table 8.2 Values of maximum stress obtained in the four different
compositions in the prestrain and strain to failure regions
under various test parameters
Table 8.3 Values of total strain to failure of different steel compositions
under various test parameters.
Table 8.4 Values of strain ageing indices in samples tested at 130 and
200°C after a time delay of 30 seconds after prestrain.
Table 8.5 Values of strain ageing indices in samples tested at 130 and
20V°C after a time delay of 600 seconds after prestrain.
Table 8.6 Work hardening exponents of the four grades of steels as a
function of delay time.
Table 8. 7 Variation in work hardening exponents with the interstitial
content for the four steel compositions.
Table 8.8 Comparison of calculated values(using Gladman 's equation)
of flow stress and measured values of flow stress at a true
strain of0.8.
Table 8.9 Values of maximum stress obtained for the four different steel
compositions under various test parameters.
Table 8.10 Variation of total strain to failure as a function of
steel composition.
Table 8.11 Variation in the ageing index, interstitial content (C+N) and
the boron content, in the four steel compositions.

xxii
Chapter 1 Introduction

CHAPTER]

INTRODUCTION

1
Chapter 1 Introduction

INTRODUCTION

Steels are by far the most popular materials for various applications ranging from

simple structural load-bearing ones to the more sophisticated aerospace applications.

Steels are ductile and are capable of being formed by most forming processes,

including forging, rolling and drawing. Certain grades of steels are capable of being

drawn into very thin wires. The drawability of a metal or alloy is dependent on the

process variables and the mechanical properties(l>. The mechanical properties of a steel

(or any other engineering material) are controlled to a major extent by its chemical

composition and thermal processing. Ideally for optimum drawability the steel should

possess a combination of good strength and ductility. However, it has been noted that

factors which result in an increase of strength, normally decrease ductility and hence

drawability<2>.

The drawability of plain carbon steel wire can be particularly affected by two

phenomena, which may occur either before or during the drawing operation - strain

ageing and quench ageing. Strain ageing manifests itself by an increase in yield stress

on ageing after or during plastic straining< 3>_ Figure 1.1 illustrates the effects of strain

agemg. Suppose a steel specimen is strained to a point A and then unloaded. If it is

then reloaded immediately and strained further,, the stress-strain relationship would

follow curve (a). If, however, after unloading from point A, the specimen was aged at

room temperature for several hours and then tested, a higher yield point would reappear

and the stress-strain relationship would follow curve (b ). These features and other

characteristics of strain ageing are described more fully in Chapter 4.

2
Chapter 1 Introduction

Ageing after straining is called static strain ageing and ageing during straining is called

dynamic strain ageing. The process of strain ageing is important commercially as it has

both beneficial and detrimental effects.

A Y = chan1• in yield stress due to strain a1ln1


AU == chance in U.T.S. due to strain a1in1
.6E • chan,a In elon1ation due to strain a1in1

a
' I •I
INITIAL LOWER
YIELD
EXTENSION
(LUOERS STRAIN

PRE-STRAIN

STRAIN

Figure 1.1 Stress-strain curves for a low carbon steel strained to point A, unloaded
and then restrained immediately-curve (a) or after ageing-curve (b) (JD Baird<37 >).

Among the beneficial effects is strengthening. For instance, static strain ageing is a

viable strengthening mechanism in some alloy systems<4 >_ Dynamic strain ageing can

be incorporated into thermomechanical treatment to improve room-temperature tensile

properties. The detrimental effects include an undesirable decrease in the ductility and

hence formability and drawability; lowered toughness; and formation of stretcher strain

markings during forming.


3
Chapter 1 Introduction

Quench ageing may be defined as any change in the properties of the material resulting

from the precipitation of a second phase at low or moderate temperatures< 5>. The

phenomenon normally increases the hardness, yield and tensile strength; and decreases

ductility. It is effective in preventing strain ageing in many cases but at the same time

it also results in hardening of the matrix, which may adversely affect the ductility and

therefore, drawability.

Worldwide, production of rod and bar steel products is moving from the conventional

method of BOS (basic oxygen steelmaking process) to the newer and efficient EAF

(electric arc furnace) method. The advantages of EAF include flexibility of raw

material, which include use of coal and scrap, no necessity for the raw material to be

prepared in a special form e.g. agglomerates or sinters and the EAF process is

economical for small scale production. However, the EAF steel grades contain higher

nitrogen contents than the BOS grades. Due to the higher temperature produced in the

arc furnace , nitrogen in the atmosphere tends to dissolve to a greater extent in the steel

as compared to that in the BOS process. Nitrogen has been known to play a very

detrimental role during further cold rolling of stee1<2· 7, 8' 9 ). This is due to the solubility

of nitrogen in ferrite and its high diffusion coefficient at room temperature, and also a

high binding energy to dislocations<J, 4, 13, 14, 28, 29, Jo, 31, 36, 37, 44, 49, 51-55, 57-59, 103). These

effects are reviewed in Chapter 4, section 4.1.4.1 of this thesis. Since the EAF process

is being adopted more widely for rod and bar steels, it is worthwhile to study and

understand the effects of nitrogen on the strain ageing of such steels, particularly those

used in wire production. This project is therefore aimed at studying the strain ageing

phenomenon in low carbon steels as a function of temperature, time and other process

variables during wire drawing. The main focus of this study will be to determine the
4
Chapter 1 Introduction

effects of strain ageing on the mechanical properties and therefore the drawability of

low carbon steels, to be more specific. For completeness, the behaviour of EAF grades

will be compared with the BOS grades. The project has been sponsored by BHP Rod

and Bar Division, Newcastle and the Australian Research Council.

This project is an evaluation of the strain ageing phenomenon in low carbon steels

containing some residual nitrogen. Therefore the following sections in the document

will concentrate particularly on the physical and mechanical metallurgical principles

and factors affecting the strain ageing process in low carbon steels. Chapters 2, 3 and

4 contain a comprehensive literature survey of strain ageing and factors that affect

strain ageing. The objectives of this research project are stated in Chapter 5. The

experimental details are described in Chapter 6. Initially, the laboratory work

commenced with hot tension experiments. However, subsequently it became evident

that industrial drawing conditions are better simulated by hot torsion testing and this

technique was adopted as the main focus of the experimental work. Chapters 7 and 8

report the main results and discuss these findings in comparison to reported literature,

respectively. The main conclusions drawn from this research work are discussed in

Chapter 9. Appendix A contains some of the hot tension plots obtained for the four

grades of steels. For ease of reference, the stress-strain curves obtained from hot

torsion experiments for various combinations of times and temperatures, have all been

placed in Appendix B. Appendix C contains the work hardening plots derived from the

hot torsion data for the four materials under various combinations of temperatures and

times.

5
Chapter 2 Physical metallurgy of low carbon steels

CHAPTER2

PHYSICAL METALLURGY OF LOW CARBON

STEELS USED FOR WIRE/ROD PRODUCTS

6
Chapter 2 Physical metallurgy of low carbon steels

The focus of this Chapter is the physical metallurgy of low carbon steel grades

commonly used for wire/rod products. However, a section on the actual grades of steels

being produced in BHP, Newcastle and an introduction to the wire drawing process

have been included at the beginning of this Chapter.

2.1 STEELS COMMONLY USED FOR WIRE/ ROD PRODUCTS AT BHP, NEWCASTLE

The commonly used grades of low and high carbon steels at BHP, Newcastle are listed

in Table 2.1. The main applications are structural and as welding rods, hence properties

like drawability/formability, strength, toughness, corrosion resistance and weldability

are important.

2.2 WIRE DRAWING - A GENERAL DESCRIPTION

The reduction in the diameter of a solid bar/rod by a successive drawing operations is

called wire drawing<!). Drawing is normally carried out at ambient temperature but as a

result of large amounts of friction generated, the temperature rises during the drawing

process somewhere in the range of 40-70°C. In extreme conditions it may exceed

100°c< 1>. Reductions per pass may range from 5 to 50%. The drawing speeds normally

range from 15 cm/sec to 150 cm/sec and in most modem machines the speeds may

reach 2500 cm/sec<!). Wire drawing is an important commercial process and, a variety

of products are produced by this process. It is compatible with various materials like

aluminium and its alloys, copper and its alloys and a number of steel grades, both plain

carbon and alloy steels.

7
Table 2.1 Common grades of steels used for wire/rod products. Compositions listed are in terms of weight percent.

Grade C p Mn Si s Ni Cr Mo Cu N
WS1033 0.3/0.35 0.04 0.8/1 0.35 0.05
WS1043 0.4/0.45 0.04 0.6/0.85 0.35 0.04
WK1005 0.05 0.04 0.5 0.1/0.35 0.04
5
WK1006 0.08 0.04 0.3/0.6 0.1/0.2 0.04 0.2 0.2 0.1 0.4 0.012
WK1013 0.12/0.16 0.04 0.4/0.7 0.1/0.2 0.04 0.2 0.2 0.1 0.4 0.012
WK10B30 0.28/0.34 0.03 0.6/0.9 0.4/0.6 0.03 B0.0014/0.003 ;Ti0.02min;A10.02/0 0.3
.05
WK1033 0.28/0.33 0.04 0.6/0.8 0.15/0.35 0.05
WK1036 0.35/0.40 0.04 0.65/0.85 0.15/0.35 0.04
WK1043 0.39/0.47 0.04 0.6/0.8 0.15/0.35 0.04
WK1047 0.44/0.52 0.04 0.6/0.8 0.15/0.35 0.04
WK1050 0.48/0.51 0.025 0.6/0.75 0.15/0.35 0.02
5
WK1054 0.49/0.57 0.04 0.6/0.8 0.15/0.35 0.04
W2K1054 0.51/0.54 0.025 0.6/0.8 0.15/0.35 0.02
5
WK1057 0.54/0.62 0.04 0.55/0.8 0.15/0.35 0.04
WK1062 0.59/0.67 0.04 0.6/0.8 0.15/0.35 0.04
W2K1062 0.59/0.62 0.025 0.5/0.7 0.15/0.35 0.04
WK1066 0.63/0.69 0.04 0.8/1.05 0.15/0.35 0.04
WK1068 0.64/0.72 0.03 0.6/0.8 0.15/0.35 0.03 0.0 l 5/0.025Al
WK1070 0.70/0.75 0.025 0.4/0.6 0.15/0.35 0.02
5
WK1072 0.69/0.77 0.04 0.6/0.8 0.15/0.35 0.04
WK1077 0.77/0.81 0.04 0.6/0.8 0.15/0.35 0.04
W2K1077 0.76/0.80 0.025 0.65/0.75 0.25/0.40 0.03
W3K1077 0.74/0.81 0.025 0.6/0.8 0.25/0.35 0.04

8
Chapter 2 Physical metallurgy of low carbon steels

The die materials used for drawing must be such that they possess good wear

resistance, be able to resist high temperatures generated by friction arising during the

drawing process, and exhibit no reaction with the metal being drawn at any

temperature. The normal die materials are tungsten carbide or steel dies with tungsten

carbide inserts. The bearing surface is that portion which guides the wire /rod as it exits

from the die and the approach angle is that portion of the die where the actual reduction

takes place.

The starting material for a given steel wire drawing process is a wire rod. Prior to

drawing, the rod is cleaned by pickling in a sulphochrome bath (sulphuric acid -

chromic acid) in order to remove the mill scale present on the surface which otherwise

could result in surface defects and excessive die wear. The rod is then coated with a

thin layer of lime or plated with a layer of copper or tin. The lime serves as an absorber

or carrier of lubricant during dry drawing and it also serves to neutralise any acid

remaining from the pickling process. In dry drawing, the lubricant is either grease or

soap powder. In wet drawing, the whole die set-up is immersed in a lubricating fluid.

For coarse wires, reductions are usually given using a single bull block in which a

single reduction may range from 25 - 50%. For fine wires, a large number of bull

blocks are used in one continuous operation and the wire is passed through a number of

dies until it is reduced to the final size. Reductions in this latter case are 5 to 25% per

pass. For steels containing greater than 0.25% carbon, a special patenting treatment is

given (t)_ This consists of heating the wire above a certain critical temperature and

subsequent controlled cooling or transforming in a lead bath held at a temperature

around 500° C in order to produce a fine pearlite microstructure. This results in the best
9
Chapter 2 Physical metallurgy of/ow carbon steels

combination of strength and ductility for successful drawing. Defects that commonly

occur during wire drawing are (a) those already present in the rod like seams, embedded

scales and piping defects; and (b) those arising out of the deformation process, like

chevron cracks or a centre burst generally referred to as cupping. These internal defects

develop as a result of " secondary tensile stresses" which occur in situations where the

mean thickness of the die (h) and the length of the deformation zone, (L) is very high

(i.e 8 = h/L >> 1) (I)_ Temperature gradients resulting in chilled zones of metal caused

when the die meets the workpiece can cause regions of low or practically no

deformation leading to the development of shear bands. Localisation of flow into such

bands may result in shear fracture. Avitzur, as cited by Dieter< 1> modelled this wire

drawing process and found that the model can be successfully used to predict the

combination of semi-die angles and reduction which would prevent the occurrence of

defects.

2.3 PHYSICAL METALLURGY OF LOW CARBON GRADES USED FOR WIRE/ROD

PRODUCTS

The low carbon steels used for wire/rod products in Table 2.1 contain carbon contents

varying from 0.05% to 0.2% (all compositions will be given in weight percent unless

stated otherwise). The major alloying elements are manganese (0.25 - 0.90%) and

silicon (0.07 - 0.25 %). The elements chromium, nickel, molybdenum are in the range

of 0.05%, while sulphur and phosphorus are maintained between 0.025 to 0.040%.

Nitrogen levels vary from a minimum of 0.008 to 0.012 %.

10
Chapter 2 Physical metallurgy of/ow carbon steels

The compositions of steels produced by the BOS and EAF processes show a very

distinct difference in the total nitrogen contents. The EAF grade contains higher

nitrogen content than the BOS grade.

2.3.1 Fe - Fe3C phase diagram

Figure 2.1 shows the Fe- Fe3C metastable phase diagram. The carbon content of Fe3C

is 6.67%, is not shown in the above diagram because it has been truncated at 5 %

carbon. The diagram indicates the presence of different phases as a function of the

temperature and carbon content. The various phases in the diagram are defined below.

a) (a - Fe)- ferrite phase which is a solid solution of carbon in iron. This is a soft

phase and has a body centered cubic (BCC) crystal structure (Figure 2.2a) and

occurs at temperatures lower than about 910°C; however the exact temperature

depends on the carbon content of the steel.

b) (8 - Fe)- ferrite phase, which has the same body centred cubic structure as the low

temperature a - Fe, but occurs at temperatures higher than 1410°C.

c) (y- Fe) - austenite phase which is a solid solution of carbon in iron and occurs at

temperature ranges higher than 727°C and below 1410°C depending on the carbon

content of steel. This phase is face centred cubic (FCC) in crystal structure (Figure

2.2b).

d) Fe3 C - cementite, an interstitial compound formed between iron and carbon. This

phase has an orthorhombic crystal structure.

11
Chapter 2 Physical metallurgy of low carbon steels

Atomic%
0 2·5 5 7·5 10 12·5 15 17·5 20
1800

1700

1600
A Liquid

p- 1300
!
i 1200
l r
! 1100 1147·

1000

T+ Fe3C

727 K-

600 - a+Fe 3 c

500
0 0-5 1·0 1·5 2·0 2-5 3-0 3·5 4·0 4·5 5·0
Weight % Carbon

Figure 2.1 Fe-Fe3C equilibrium phase diagram (Krauss').

e) Pearlite (a - Fe + Fe3C) - this is a micro-constituent consisting of ferrite and

pearlite, and forms by the decomposition of austenite of composition "S" in Figure

2.1 at temperatures below 727°C. The phase commonly occurs as lamellae, with

ferrite and cementite as alternate layers.

12
Chapter 2 Physical metallurgy of low carbon steels

(a) (b)

Figure 2.2 (a) Body centred cubic structure (b) Face centred cubic structure
(Honeycombe 10>

The main constituents of a low carbon steel microstructure are typically (in descending

order of volume percentage)

1. ferrite (a - Fe)

2. pearlite (a - Fe+ Fe3C); and

3. finely dispersed constituents such as various nitrides / carbides / sulphides or

carbonitrides ( e.g AlN, VN, TiN, V4 C, MnS, TiCN etc.)

The equilibrium amount of pearlite in a Fe-C steel containing 0.2% carbon would be

about 25 volume% (Figure 2.1 ). There is often also a dispersion of very fine

nitrides/carbides in the ferritic matrix or grain boundaries of the low carbon stee1<3>· A

brief paragraph on the possible transformations on continuous cooling of steels 1s

included in the next section.

13
Chapter 2 Physical metallurgy of/ow carbon steels

2.3.2 Transformations during continuous cooling :

Many of the heat treatments performed on steel involve continuous cooling conditions,

and the transformations that occur are represented by transformation diagrams(?).

These diagrams can be experimentally determined through the use of dilatometry in

which changes in length are measured as a function of temperature and time. Figure

2.3 shows the relationship between a transformation diagram and the Fe- Fe3C phase

diagram for steels of three different carbon contents.

As illustrated in Figure 2.3, the transformation diagrams indicate the different types of

phases obtained when a steel of a certain composition is cooled. If the cooling rate

employed is such that it cuts the pearlite-ferrite curves, then these are the phases

expected in the steel. However, if a cooling rate that would result in avoiding

intersection with the pearlite-ferrite curves were used, the metastable and brittle

martensite phase would form. The minimum cooling rate that is tangential to the

leading nose of the curves is called the critical cooling rate. Any rates above this would

result in the formation of martensite.

Transformations are very much influenced by the steel composition i.e type and amount

of alloying elements present in steel. The exact effects are discussed very briefly in the

following sections.

14
Chapter 2 Physical metallurgy of low carbon steels

nnp•ratur•
.
C

900 900
100 800
700 700
600 600
n,
500 500
m P•B
400 400
300 300
200 200
100 100
0
(a)
0 0.2 0.4 0,6 0.1 ,.o 1,2¥, C
0
0.1 I 10 102 103 104 105 s
Tim•
T•mprrature
9c c
900
800
700
Ilf'•P
600
mP+B
500
400 400
300 300
200 200
100 100

(b)
0
0 0,2 Q,4 a., a.e 1,0 1,2 .,.c 0
0.1 I 10 10 10) 10 10 s
Tim•
T•mp.,.at-
c •C
900
800
700
600
SOO
400
300 300
200
100 100
0 0
(cl 0 0.2 a., o., a,,, ~o 1,2 °/o C 0.1 I 10 10 I 10 4 10 s
Tim•

Figure 2.3 Relationship between the Fe-Fe 3C and the structural transformations
resulting from various cooling curves in steels containing a) 0.8%C b) 0.45%C and
c) 1.0¾C (K.E Thelning< 11))

15
Chapter 2 Physical metallurgy of low carbon steels

2.3.3 Effect of alloying elements:

Plain carbon steels are quite limited in their range of applications due to their narrow

span of useful properties. Alloying elements are added to steels for purposes mainly

aimed at overcoming the limitations of plain carbon steels. Some of the important

advantages achieved are an increase in hardenability and improved mechanical

properties at room temperature and increased corrosion and oxidation resistance.

Each alloying element has a characteristic effect on the properties of the particular steel

being alloyed. Generally the effects may be divided into the effects on the Fe-Fe3C

diagram and the kinetic effects that occur during heat treatment. The effects during

heat treatment are well represented by the effect on the continuous cooling curves. By

affecting the phase stability and kinetics of transformation behaviour, the alloying

elements may raise or lower the transformation temperature, shift the eutectoid

composition, and alter the position of the continuous cooling curves to the left or right.

2.3.4 Classification of alloying elements in steel:

There are two distinct effects:

1. Alteration of the relative stability of ferrite and austenite by the element- hence

classification as austenite stabilising elements or ferrite stabilising elements.

2. Alteration of the relative stability of graphite and cementite/carbide phase -

hence classification as graphite stabilising or carbide forming elements.

16
Chapter 2 Physical metallurgy of low carbon steels

However, it must be noted that these are equilibrium effects and there are other effects

based on kinetic factors.

The specific effects of each of the common alloying elements added to low carbon

steels have been reviewed by Clark and Varney< 8>, Porter and Easterling<9> and

Honeycombe< 10>and the following text provides a brief summary.

• Manganese

The diffusion of carbon in austenite is retarded by manganese< 8>. Manganese provides

solid solution strengthening to ferrite. Manganese has an unlimited solid solubility in

austenite and has a 3% solid solubility in ferrite< 8>. Manganese shifts the CCT curve to

the right, lowering the critical cooling rate and increasing the tendency to form

martensite (increasing hardenability). Manganese decreases the extent of strengthening

with a reduction in the grain size. i.e. it decreases the value of 'k' in the Hall-Petch

relationship. Manganese also affects the modulus of elasticity and the shear modulus.

• Silicon

Silicon strengthens ferrite(s, 12> (Figure 2.4). It is a ferrite phase stabiliser and shifts the

eutectoid carbon compositions to lower carbon contents. Silicon delays the formation

of pearlite, shifting the CCT curve to the right. This decreases the critical cooling rate

and increases the tendency to form martensite. Silicon has a solid solubility of 2% in

austenite and 18.5% in ferrite cs, 12>. Silicon partitions to the ferrite phase and exhibits a

positive thermodynamic interaction parameter with carbon thus raising the activity of

carbon in austenite. This means that carbon will not only diffuse away from regions of

17
Chapter 2 Physical metallurgy of low carbon steels

high carbon concentration but also from regions rich in siliconC9>_ Silicon is used as a

general deoxidiser .

.........
HV
240
p Si Mn

220
: ,,I /
200 I I/ __,.. ,,,,,,, Mo
V

1eo I I /
r
~ 1-'P-
.,,,,,,,,,.._ i -'

,.o /, I __,,Ni _.,,,,,. ~ i.,.-'"


.' / ,,,, ~~ ~ -- - ~
....---- w

J I,, , ,,- ......~ ~ ~ ---- ---


140 i- Cr

120

100
~ ~
j , ".A'.. ~ --
~ ::;..-- -- --- ~

eo

0 2 · 4 8 8 10 ·12 14 19 18 20 22 24 .,.
Concentration of a l l ~ element in J..- iron I fenite J

Figure 2.4 Effect of substitutional alloying elements on the hardness of ferrite


(Bain and Paxton< 13>)

• Chromium

Chromium is a strong ferrite stabiliser as shown in Figure 2.5 and therefore raises the

eutectoid transformation temperaturecs, 13 >_ It also shifts the eutectoid carbon

composition to the left, i.e, to lower carbon compositions. Chromium has a high solid

solubility in ferrite and a 12.8% solubility in austenitec 8>. Chromium also has a

tendency of delaying the formation of pearlite, thereby shifting the CCT curve to longer

times and increasing the tendency for martensite formation. Chromium has a tendency

to retard the austenitic decomposition by formation of chromium-rich carbides, since

these carbides are thermodynamically more stable than cementiteC 12>. These carbides

lead to greater undercooling and therefore finer interlamellar spacing in pearlite02 >.

These chromium-rich carbides also act to improve the wear resistance of steel.

Chromium decreases the growth rate of pearlite and partitions to the cementite phase of
18
Chapter 2 Physical metallurgy of low carbon steels

pearlite02 >. Chromium also increases the corrosion and the oxidation resistance of

steel.

2800.....---r-----,r--...._--,r---r----.---------

- 2200t---'1f----bl~-'!-'-71r__-R-,::::!!lf,,-..::-=+-=:---+---+------t
lL
0

-;_ 2000 1---.,__-+---+-+---f.


E
~ 1800 If----+---+-+-,,,<-~-

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8


Carbon ( per cent)

Figure 2.5 Effect of chromium on the ferrite region (E.C Bain and Paxton< 13>).

• Nickel

Nickel is a very strong austenite stabiliser and thus increases the range of temperatures

over which austenite is stable. Nickel also shifts the eutectoid composition to lower

carbon content. Nickel has an unlimited solid solubility in the austenite phase and 10%

solid solubility in ferrite<S).. Nickel increases hardenability and shifts the CCT curve to

the right, thus lowering the critical cooling rate. But nickel tends to retain austenite

with high carbon< 8>.

Nickel contributes marginally to the solid-solution strengthening of ferrite. Nickel has

the effect considerably decreasing the ductile-to-brittle transition temperature< 8>. The

transition temperature is lowered considerably, thereby increasing the toughness of

steels. Nickel does not form carbides.

19
Chapter 2 Physical metallurgy of/ow carbon steels

• Molybdenum

Molybdenum is known to raise the grain coarsening temperature of austenite<S)_ It

provides slight solid-solution strengthening to ferrite and shifts the CCT curve to the

right, thus lowering the critical cooling rate and increasing hardenability (i.e the

tendency to form martensite ). The effect of molybdenum on the austenite phase field

is shown in Figure 2.6. Molybdenum is a strong carbide former and contributes to

secondary hardening during tempering. Molybdenum has a higher solubility in ferrite

(37.5%) than in austenite (3%i 8' ' 3)_ It is added to steels to counteract the tendency

towards temper embrittlement. Molybdenum forms abrasion-resistant carbides which

in tum raise the hot hardness of the steel(s, 11 >.

• Niobium

Niobium forms a very stable carbide, nitride and carbonitride. As such, the presence of

niobium in steels slows the recrystallisation process during hot working, leading to a

refinement in the ferrite grain size. Niobium has very limited solid solubility in

austenite and is a major contributor to precipitation hardening.

• Vanadium

Vanadium has an unlimited solid solubility in ferrite and up to 4% in austenite< 8>. It has

a moderate solid-solution hardening tendency in ferrite. Vanadium shifts the CCT

curve to the right, and correspondingly lowers the critical cooling rate and thereby

increases the tendency to form martensite. Vanadium forms very stable carbides and

nitrides.

20
Chapter 2 Physical metallurgy of low carbon steels

2aoo------.---,--~----------------,
26001--_..,_--+-...;.a,.~---+---+--+---+---+---t-----l

1400 1----+------li--;;;;:-=---s-.:::-:::-_-'"ff""lc..,_--+---t---+--t---1

12000 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8


0.2
Carbon ( per cent)

Figure 2.6 Effect of molybdenum on austenite region (E.C Bain and Paxton< 13>).

• Cobalt

Cobalt has an unlimited solubility in austenite and is soluble upto 75% in ferrite< 8> • It

hardens ferrite considerably by solid solution hardening. Cobalt in solution has a

tendency to decrease hardenability and therefore shifts the CCT curve to shorter times,

thus increasing the critical cooling rate.

• Nitrogen

Nitrogen forms an interstitial solid solution with ferrite< 3>_ Nitrogen addition to steel

results in an expanded y phase field(S, 10>. Interstitial, or free nitrogen is the most

common form in steel and gives rise to strain ageing problems at temperatures lower

than 100°C (as will be discussed in detail in further sections). The solubility of nitrogen

in ferrite, at various temperatures has been studied by Leslie <3, 14> (shown in Figure

2.7). It is seen that for a steel with a nitrogen content of 0.01 wt%, the solubility is only

21
Chapter 2 Physical metallurgy of low carbon steels

exceeded during cooling below 1173K (900°C). The extrapolated solubility of

nitrogen at room temperature is in the range 10·4 to 10-5 wt%.

0,05

0,04
l

0,0.3
N

Wt.% 0,02

0,01 ,,," s
,,
I

,/

0
700 1100 1500 1900

DEGREES, ° K

Figure 2.7 Solubility of nitrogen in different phases of steel (L.S Darken cited by
Leslie< 14>).

22
Chapter 3 Precipitation kinetics

CHAPTER3

PRECIPITATION KINETICS

23
Chapter 3 Precipitation kinetics

3.1 PRECIPITATION KINETICS:

As will be discussed in detail in section 4.1.1, strain ageing is due to the combined

effects of solute atom atmosphere formation around dislocations and also precipitation.

It is therefore very relevant to study the kinetics of precipitation. The process of

precipitation is typically diffusion controlled and takes place by nucleation and growth.

An understanding of the nature of these processes is very important in order to find an

effective way of controlling or offsetting strain ageing.

3.1.1 Diffusion:

Before understanding the kinetics of the diffusion process, a few basic thermodynamic

terms need to be defined.

Gibb's Free Energy: The relative stability of a system is determined by the Gibb's

free energy( 9), which is defined by the equation

G = H -TS-------------------------------------3.1

where H is the enthalpy, T is the absolute temperature and S is the entropy of the

system. Enthalpy is a measure of the heat content of the system and is given by

H = E + PV-----------------------------------3.2
where E is the sum of the internal energies of the system, P is the pressure and V the

volume.

A system is stated to be at equilibrium when it is at its most stable state, which in

thermodynamic relationships translates to the lowest possible value of Gibb's free

energy. This means that high stability is obtained with the best compromise between
24
Chapter 3 Precipitation kinetics

low enthalpy and high entropy. Therefore any transformation, which results in a

decrease in the Gibb's free energy, is possible. The transformation, however need not

go directly to the minimum free equilibrium state, but can pass through several

intermediate metastable states where the free energy values lie at a local minimum.

Most of the phase transformations that occur in steel are diffusion controlled and hence

depend on temperature and time in a non-linear manner<9>. The driving force for the

diffusion process is the inherent tendency of a system to attain the lowest possible

energy state where it becomes stable. Diffusion is typically related to the concentration

gradients across a system. The common principle that governs this process is that an

element diffuses across a concentration gradient, such that long-range diffusion ceases

only when the concentration of the element becomes equal in the two regions. Figure

3.1 illustrates this process in two systems, where point A indicates 100% of element A,

µ 1A, µ 18 and µ2A, µ28 are the chemical potentials of A and B in two different solid

solutions 1 and 2.

Chemical potential is the partial molar free energy of the element. This term is

introduced to understand the effect of addition or removal of atoms on the total free

energy of a given phase such that

dG - µA dnA -------------------------------------------- 3.3

This relation holds true when the temperature, pressure and the number of atoms of the

second element are constant. This proportionality constant in the equation is the

chemical potential, and it depends on the composition of the phase. Therefore dnA must

be so small that the composition of the phase is not significantly altered. As indicated

25
Chapter 3 Precipitation kinetics

in Figure 3 .1, diffusion of a given element always occurs down its gradient of chemical

potential.

~
8-rich A-rich
(o) (b) A ® (D B

~@I
8-rich A-rich
(c) (d)A ® (D 8

µl µi

µi µi
µr µi
µl µi
(e)
A ® (D 8
(f)
A ®CD 8

Figure 3.1 Free energy and chemical potential changes during diffusion - (a)i and
(b) down hill diffusion and (c) and (d) uphill diffusion (Porter and Easterling< >)

The two mechanisms by which atoms diffuse through a solid depend on the type of site

occupied by the element in question in the lattice of the matrix. The two mechanisms

are as outlined below.

(a) Substitutional diffusion- atoms diffuse through vacancy defects present m a

material.

(b) Interstitial diffusion - atoms force their way through the interstices between the

matrix atoms without permanently displacing any one of them< 15 )_

26
Chapter 3 Precipitation kinetics

Whether a solute goes into solution interstitially or substitutionally, depends on the size

of the solute atom (IS)_ An empirical rule based on x-ray data indicates that in cubic

lattices, a solute will dissolve interstitially only if it atomic radius is less than 0.59 times

that of solvent. However, practically, it has been observed that solute with radii up to

0.85 times that of the solvent can dissolve interstitially< 15 >_ Since strain ageing is related

to the diffusion of interstitial atoms only, this section will discuss only interstitial

diffusion.

3.1.1.1 Interstitial Diffusion

Fig 3.2 shows the interstitial sites in body centred and face centred cubic structures.

Usually the interstitial concentration is so low that only a very small fraction of the

indicated sites are occupied. This would mean that the interstitials are surrounded by a

number of vacant sites, which can be occupied as often as the energy conditions would

favou/ 9). This mechanism has been explained using figure 3.3. For the interstitial

atom in position I to move to position 2, atoms in positions 3 and 4 would have to

move apart, i.e a distortion in the lattice has to occur. The ease with which this happens

is the basis for diffusion of interstitials through interstice positions. This process could

take place as a "random jump" process in either a steady state situation or a non-steady

state one.

27
Chapter 3 Precipitation kinetics

(a) • Melal a1oms (b) e Metal atoms


•.) Oclahedral inlerslices o Tetrahedral interstices

Octahedral (a) and tetrahedral (b) int· stitial voids in fee struc-
turc.

·O·f,.o
··
.o ·:

(a) e Metal aroms (b) e Me1aI a1oms


o Octahedral inlerslices o Tetrahedral inlerslices

Octahedral (a) and tetrahedral (b) interstitial voids in bee


structure.

Figure 3.2 Interstitial positions in FCC and BCC structures (Porter and
Easterling<91).

28
Chapter 3 Precipitation kinetics

(a) (b)

o.oo 0 0
000
~
0
l
G '1Gm (c)
j
a.
X _,.

Figure 3.3 Interstitial atom in (a) equilibrium position (b) at the position of
maximum lattice distortion (c) Variation of free energy of the lattice as a function of
the position of the interstitial (Shewmon< 15>).

• Random Diffusion:

This mechanism of diffusion can be explained by assuming a dilute interstitial solid

solution <9>_ The parent atoms (A) are arranged on a simple cubic lattice and the solute

atoms (B) are accommodated in the interstitial positions in the crystal structure.

Another assumption is that the solution is so dilute, that every interstitial atom is

surrounded by six vacant interstitial sites. If the concentration of B varies along say,

only one dimension in the material, i. e direction x, then B will continue to diffuse to

regions of low concentration tuntil the chemical potential of B is the same everywhere.

The local flux of B atoms diffusing in unit time, is given by Fick's first law of diffusion

Je = - De (oCe / ox)---,--- ---------------3.4

The partial derivative indicates the concentration gradient and implicitly assumes that

(oCe I ox) a µe/ox --------------------------------3.5

29
Chapter 3 Precipitation kinetics

Ds is the diffusion coefficient of atom B which is independent of the concentration of B

and is given by

De = 1/6 (f' e/ a 2) - - - - - - - - - - ----3.6


Here, r s is the number of times an interstitial atom would jump per second and a is the

average separation between planes.

Diffusion is a process which is dependent on both temperature and time. Due to the

thermal energy of a solid, all atoms are in a continuous process of vibration. An

occasional extreme oscillation could result in a jump of the interstitial atom to another

position. In order to move an interstitial atom to a new position, the atoms of the matrix

have to be forced apart as shown in Fig 3.3(b). This causes an increase in the free

energy of the system by say, Ii Gm, which is in other words is the activation energy for

this process (Figure 3.3(c)). The energy of the atoms changes as they continuously

collide with one another. The fraction of atoms having an energy of Ii G or more than

the mean energy is given by the Boltzmann relation exp (-Ii G / RT). If the interstitial

atom vibrates with a mean frequency v in the x direction it makes v attempts in a given

time to jump to the next available site.

The fraction of the attempts that are successful is given by exp (-Ii Gm/ RT). If there

are z number of sites available, then the frequency of jumps is given by the following

equation.

f' 8 = zv exp (-Ii Gm/ RT)-------------------------- 3. 7

!l.Gm = !l.Hm -T!l.Sm ------------ -------3.8

where !l.Hm is the activation enthalpy and !l.Sm is the entropy term. The diffusion

coefficient at this temperature will be given by


30
Chapter 3 Precipitation kinetics

Ds = Dso exp (-Q IRT)------------·-----3.9


where, from equation 2.5,
Dso = 1/6 a 2 zv exp(~ Sm/ R)------------ --3.10
and Q- ~ Hm-----------------------------------3.11
D or r has been found to increase exponentially with temperature at a rate determined
by the activation enthalpy Q. Equation 3.9 is found to agree with experimental

measurements of diffusion coefficients in interstitial systems. It has been found that the

activation energy in these systems is only dependent on the activation energy barrier to

the movement of interstitial atoms from one site to another< 9 • 15 ).

Once atoms of a given element diffuse to regions of low chemical potential, there is a

point in time when the concentration of the element is such that it could combine in

stoichiometric ratios with other similarly diffusing elements to form a compound,

which could nucleate and grow as a function of time and temperature. The next section

attempts to explain mechanisms of nucleation and growth in very general terms.

3.1.2 Nucleation of a solid in a liquid:

The principles of nucleation can be most easily treated by considering solid nucleation

in a liquid cooled below its equilibrium melting temperature<9l_ Nucleation is a process

through which the smallest stable particle of a new phase is formed< 16l. The driving

force for solidification (GL - Gs) would allow the liquid to solidify (where GL and Gs

are free energies of the liquid and solid phase respectively). The nucleation process is

not instantaneous and liquids can be undercooled and held at the undercooling

temperature for a long time without anything happening. This is particularly observed

when the solid is forced to nucleate homogeneously, i.e when solid nuclei form within

the volume of the liquid itself. The process of homogenous nucleation requires a very
31
Chapter 3 Precipitation kinetics

high degree of undercooling. There are several theories proposed to explain the

nucleation process.

Gibbs explained this nucleation process thermodynamically <16>. Consider a liquid

system with a free energy G1, which is represented as

G1 = VL GvL-------------------------------3.12
If an embryo of a new solid phase forms, the free energy of the system changes to G2,

which is given by

where Vs is the volume of the solid sphere, VLthe volume of the liquid, AsL, the solid-

liquid interfacial area, Gv8 and Gi are the free energies per unit volume of solid and

liquid respectively. Due to the formation of a new surface, this embryo is associated

with some interfacial energy YsL- This in tum serves to increase the free energy of the

system. Any stable system sustains itself by attaining the lowest free energy state.

Therefore, an increase in free energy of the embryo would naturally tend towards

dissolution of the embryo in the parent phase. However, if an embryo of a specific

chemical and crystallographic orientation (eg. a spherical shape), forms, the increase in

the interfacial energy could be offset by a decrease in the chemical free energy, thus

increasing the stability of the embryo<9>_ Therefore, the net free energy now becomes

~ G = -4/3 7t r 3 ~ Gv+ 4n r 2 ysL ----------------------------3.14

Figure 3.4 illustrates this variation in the free energy change associated with nucleation

as a function of the embryo radius r*. As shown in the figure, the positive interfacial

energy term is proportional to the second power of the linear dimension of the embryo

(Gocr2); while the volume free energy change Gv is proportional to the third power of

32
Chapter 3 Precipitation kinetics

the linear dimension of the embryo. Therefore, creation of solid particles will always

lead to an increase in the free energy of a system. At small embryo sizes, the addition

of each atom will make the total free energy change AG, more positive. It can be seen

that for a certain degree of undercooling, there is a certain radius r*, which is associated

with a maximum excess free energy<9, 16>. If r < r* (the critical embryo/cluster radius)

then the system can lower its free energy by dissolution of the solid, while in a situation

where r >r* (nuclei), the free energy of the system decreases if the solid grows. The

effect of interfacial energy can be overcome only when 8AG/ 8r = 0, this defines the

critical size nucleus r* or the number of atoms in the critical size nucleus n* and the

free energy of activation for the formation of this critical size nucleus, A G*.

r* = 2 YsL I A Gv --------------------------- 3.15

and A Gv= Lv AT/ Tm-------------------,----3.16

AG* =l61t YsL3/3 (A Gv )2 =(l61t r3sL T2m)/ (3Lv2 A T2) 3.17

where Tm = melting point , A T = degree of undercooling and Lv = latent heat of

fusion. As the amount of undercooling increases, the critical size of cluster required to

be stable decreases.

Since AGv is often found to be proportional to the ratio of y3 to undercooling AT, where

y is the surface energy of the new surface created. The following general equation holds,

L1G* = k y I (L1T) 2 ------------------------------- 3.18

where k is the constant of proportionality, which depends on the slope of the solvus

boundary. Thus, the nucleation rate, both in a solid and in a liquid, increases for a

given system as the surface energy decreases and AT increases.

33
Chapter 3 Precipitation kinetics

.6G
+

0 r

Figure 3.4 Variation of free energy with embryo radius (Porter and Easterling<9>)

Gibbs related .1G* to the volume of the critical nucleus V* as follows <9, 16>

AG*= - (V*/2) AGv ----------------------------------------- 3.19

Thus any mechanism that would reduce V* and AGv would result in a reduction of.1G*

and hence, an increase in the nucleation rate at a given temperature. The value of

V*can be decreased by nucleation on a foreign surface or on a defect, such as a grain

boundary or dislocation< 15>_ This is referred to as heterogenous nucleation.

34
Chapter 3 Precipitation kinetics

Figure 3.5 Variation of r* and rmax with the amount of undercooling (Porter and
Easterling(9>).

The variation of ~G* with the amount of undercooling, for both homogenous and

heterogenous situations is shown in Figure 3.6.

It has been observed that that a much smaller undercooling is needed for the onset of

nucleation kinetics in heterogenous than in the homogenous process

The steady state nucleation rate N is given by( 16>

N = Z /3*0 exp ( - LIG* I kT) exp ( r-lt) -----------------------3.20

where Z is the Zeldovich non-equilibrium factor, /3* is the rate at which atoms are

added to the critical nucleus or the frequency factor, 0 is the number of nucleation

sites per unit volume, LIG* is the critical free energy for nucleation , k is the

Boltzmann's constant and T is the absolute temperature. The Zeldovich factor converts

the metastable equilibrium number of nuclei to the smaller number actually present

because some are lost to the growth stage of transformation. Normally Z= 10· 1- 10·2• f3

* is proportional to the diffusivity of an element in the matrix phase or nucleus: matrix


35
Chapter 3 Precipitation kinetics

boundary diffusivity. 0 is dependent on whether the nucleation is homogeneous or

heterogenous.; however, this parameter is unaffected by the alloy composition, reaction

temperature and time< 16>. T is the incubation time and is affected by temperature and

composition. Figure 3 .6 shows the variation of the steady state nucleation rate N with

the reaction temperature for a given alloy composition and the precipitate phase< 16>. It

is evident from the above equation that the rate of nucleation would increase if AG*

decreases. There is a large reduction in the amount of undercooling that is required for

an increased nucleation rate. This would mean that a number of nuclei could form

under conditions of heterogenous nucleation more easily than for a homogenous

condition, and more importantly, for a smaller degree of undercooling.

Although initially, the stage of solid nucleation has been studied on a system where

solid forms in the liquid, similar studies have been extended to formation of a new solid

in an already solidified system. The following paragraphs concentrate on the

nucleation of a solid in a solid system.

3.1.3 Nucleation in solids:

Solid-state nucleation is usually related to phase transformations which are thermally

activated atomic movements<9 , 15 >. There are several types of transformations possible.

However, this section will only be concerned with the mechanisms of precipitation.

Precipitation, in general, can be expressed as follows:

d ---+ a + P-----------------------------------1.21

36
Chapter 3 Precipitation kinetics

where a1 is a metastable supersaturated solution, f3 is the stable or metastable

precipitate and a is a more stable solid solution with the same crystal structure as a1,
but with a composition closer to equilibrium<9>.

'\
\
tJ.G~ \
\
\
\
\ tJ.Ghet
. Criticol value
for detectable
nucleation
\

' I'-
---~---- --'---
(a) 0
I

I
I
--- -- I

I
- ------AT
I I
I I
1 I
I I
I I
N I
I
Nhet I Nhom
I
I I
I I
I I
I I
I I
I I
I
(b) 0

Figure 3.6 Variation of ~G and N with undercooling for homogenous and


heterogenous nucleation (Porter and Easterling<9))

As dealt with in section 3 .1.2, this precipitation phenomenon will be explained in terms

of first homogenous process principles and then in terms of heterogenous process

principles.

3.1.3.1 Homogenous nucleation in solids

Consider a solid /3, which is rich in B and forms in a supersaturated A-rich solution.

For nucleation of /3, B atoms within the matrix must first diffuse together and for a

37
Chapter 3 Precipitation kinetics

compound to precipitate, its concentration at a particular temperature must exceed the

solubility product in the matrix phase. The solubility product K, is given by the product

of element concentrations in the phas, at a particular temperature.

K - [A]x [B]Y --------------------------------------- 3.22

where, A and B are the two elements in question that form a compound AxBy.

The larger the value of the difference from the solubility product at a particular

temperature, the higher is the driving force for the nucleation of the compound

precipitate. Figure 3.7 shows the relationship between % B and % N in austenite at

1000 °C. Points A and B are two B-N concentrations in two different steels. It is

evident that both the alloys have exceeded the solubility product of BN at this

temperature. It is expected however, the alloy at Point B will precipitate BN more

easily than that at A, due to a larger driving force available for it to do so.

%Bvs%N

0.000002

0.0000018

0.0000016

0.0000014

0.0000012
ID
~ 0.000001
~
0.0000008

0.0000006

0.0000004

0.0000002

0
0 0.002 0.004 0.006 0.008 0.01 0.012
wt%N

Figure 3. 7 % B versus ¾N (Maitrepierre cited by Franks and Kirkcaldy (73 >)

38
Chapter 3 Precipitation kinetics

As with the liquid to solid transformation, an a Ip interface must be created during the

precipitation process and this leads to an activation energy barrier. In addition, the

transformed volume will not fit perfectly into the space originally occupied by the

matrix and this gives rise to a misfit strain energy ~Gs per unit volume of p The

strain energy acts as a barrier to nucleation. The free energy change associated with

the nucleation process will have the following three contributions:

• At temperatures where the p phase is stable, the creation of a volume V of

p will cause a volume free energy reduction of V~Gv.

• The creation of a new surface will give a free energy increase of A 1.

• In general, the transformed volume will not fit perfectly into the space

originally occupied by the matrix and this gives rise to a misfit strain

energy ~Gs per unit volume of p

Summing this up leads to the total free energy term as

LlG = - V LlGv + A y+ V LlGs _ _ _ _ _ _ _ ___,,23

Assuming that the nucleus is spherical with a radius of curvature r ,

LlG = - 47!1 (LlGv - LlG.J + 4m-2 r------------------------------3.24

On differentiating equation 3.24, we get

r* = 2y/ (LlGv - LlG.J-------------------------------- 3.25


and

LlG* = 161fyl 3 I (LlGv - LlG.J 2-----------------------3.26

This expression is similar to that for a nucleation of a solid from a liquid (solidification),

except that the chemical driving force, which is the misfit strain energy, LlGs, is reduced

by a positive strain energy term. The homogenous nucleation rate is expressed as


39
Chapter 3 Precipitation kinetics

N = wCo exp (-LI G,,/kT) exp (-LIG*lkT)--------------------3.27


Where OJ is a factor that takes the frequency of vibration and the area of the critical

nucleus.

3.1.3.2 Critical energy for heterogenous nucleation:

As mentioned earlier, homogenous nucleation rarely occurs in practice and it is

heterogenous nucleation that is the most important factor. Both the situations of

coherent and incoherent nucleation at the dislocations have been extensively studied (9,

15 ' 16>. Coherency arises out of a perfect match at the interface plane of two crystals in

such a way that the two lattices are continuous across the interface. This can be

achieved if the interfacial plane has the same atomic configuration in both phases.

Incoherent interfaces have very different atomic configurations in the two phases. In

both these situations, it has been found that the volume strain energy may be reduced,

thus aiding nucleation. During incoherent nucleation on the dislocations, it was

observed that a destruction of a portion of the dislocation core and the subsequent

elimination of the dislocation's strain field leads to a change in the free energy which

accelerates nucleation( 16>. While in coherent nucleation, the nucleus forms on the

tensile/compressive side of the dislocation core, depending on whether the volume

change is negative or positive and at a distance away from dislocation core so as to

minimise the volume strain energy of the nucleus. This could result in the partial

nullification of the strain field around the dislocation and reduce ~G* further.

Although the dislocation does aid coherent nucleation to a much smaller extent than

incoherent nucleation, this effect is decreased to a large extent as a result of the high

nucleus: matrix interfacial energy associated with incoherent nucleation. However,

40
Chapter 3 Precipitation kinetics

both incoherent and coherent nucleation on dislocations have common features such as

higher rates of nucleation at edge that at a screw dislocations< 16>. The nucleation rate

also increases with increasing values of ~Gv , and at dislocations with longer Burgers

vector and at planes parallel to the particular habit plane which allows a maximum

relief of strain energy.

Strain ageing is a phenomenon associated with the segregation of free nitrogen and

carbon to the dislocations and precipitating on them. It is therefore very relevant to

study the principles of heterogenous nucleation on dislocations. In order to determine

the nucleation rate, the free energy barrierL1G* <16> that must be overcome by nucleation

must be understood. As indicated in the above equations, it is obvious that the

nucleation rate decreases exponentially as the critical free energy decreases. This

critical free energy term L1G* can be determined from the free energy - size relationship

during nucleus formation. According to the classical nucleation theory<9, 16) , the free

energy change that accompanies the formation of a nucleus on a dislocation

(heterogenous nucleation) L1G°, is represented in equation 3.29.

L1G° = V ( L1Gchem + L1Gc) + S~y----------3.29

where L1Gchem is the chemical driving force for nucleation, L1Gc is the volume strain

energy that is created by the formation of a new volume (nucleus) , y is the interface

energy and ~ is a y modifier that is associated with the presence of dislocations. V and

S are the volume and surface area of the nucleus Assuming that the precipitate is

spherical, the critical free energy L1G*, for the formation of a particle of critical

diameter d can be written as

L1G* = J6,r~ 3 y 3 I 3( L1Gchem + LiGc/----------------------------3.30

41
Chapter 3 Precipitation kinetics

This Chapter hitherto has described mechanisms of nucleation and growth of a

precipitate. The following section is focussed on the description of three important

experimental methods of monitoring the kinetics of the precipitation process.

3.1.4 Experimental methods of monitoring precipitation kinetics:

A number of techniques have been used to study the precipitation process. These

include microhardness, internal friction measurement( 122' 123 ), electrical resistivity(!?, 18)

and more recently constant strain rate hot compression testing(l 9)_

In the chemical and electrochemical extraction methods (20 , 21 ), precipitates of the order

of 10 °A in diameter or less are extremely difficult to extract. While in the method of

evaluation by electron microscopy, extraction replicas vary in quality from operator to

operator. Moreover small precipitates do not always remain on the replica upon

extraction. Another disadvantage is that the area observed is extremely small,

sometimes making data obtained not statistically meaningful. Evaluation methods

using bulk properties as a measure of the precipitation process, i.e hardness and

electrical resistivity are dependable (l 3)_ But the results obtained should be carefully

interpreted as various other phenomena could influence the measurements.

It is prudent therefore to use more than one of these methods in order to clearly

evaluate the situation. Three methods of analysis are highlighted in this literature

review.

42
Chapter 3 Precipitation kinetics

3.1.4.1 Electrical resistivity

The electrical resistivity or, inversely, the conductivity of pure metals follows a

periodical variation with atomic number similar to other electronic properties, but with

individual differences< 17• 18>. The conductivity of the noble metals is the greatest while

the transition metals are considerably lower. Electrical resistance in single crystals

varies with crystallographic direction, a characteristic, which is not very distinct in

polycrystals unless a strong texture develops.

The elementary interpretation() 7> of resistance assumes that it results from the vibration

of atoms which scatter the electrons moving through the lattice. These vibrations

increase with temperature. The mean kinetic energy of the vibrating atoms is given by

mean kinetic energy=½ 4,I V MX1 ------------------------3.31

where, M is the mass of the atom ; v is the frequency of vibration and X is the mean

square of the displacements from the mean lattice position.

The resistance R is proportional to the probability of scattering<22 >, which is

proportional to X 1 and inversely proportional to the mean free path of an electron.

Therefore the kinetic energy per atom for each of the three directions at 90° is given by

2,r 1 v 1 MX1 = l/lkT--------------------------------------3.32

Quantum theory associates an energy h v with a fundamental frequency of vibration v.

The relationships for temperature variation naturally involve functions of h v lkT.

Therefore the above equation becomes

X1 = (h 1 I 4 ,Ik OM)* T/0-------------------------------3.33

43
Chapter 3 Precipitation kinetics

and

PIT= Constant * X 2-------------------------------3.34

where /JJT is the main resistance component that varies with temperature and 0 = h v

lk, a characteristic temperature. This, therefore gives the relationship between

temperature and resistivity such that resistivity increases with increasing temperature.

It is obvious that there will be a change in the frequency of atom vibration when there is

a change in the composition, i. e., the size of the atoms may vary resulting in varied

scatter of the electrons and therefore a different value of resistivity. The application of

resistance measurements therefore proves to be an excellent tool used to study the

metallurgical changes. For example, an important metallurgical change that can be

studied by resistance measurements is the progress of precipitation of a second phase

by nucleation and growth. A typical example is shown in Figure 3.8. The resistance of

an alloy system of A-Bis shown in (a) where a is the primary solid solution and p the
phase that precipitates. For a temperature T0 , and a composition x % of B, only the a

phase is present. The resistance of this is indicated by curve (b1). At room

temperatures, the resistance would be as indicated in curve (b2). This would apply if

there was no precipitation or a phase change. Equilibrium at T2 (room temperature)

would however involve precipitation to give a p phase of composition c2 leaving a of

composition a2. In practice an alloy would be quenched from above To and aged at T,,

between T O and T 2. In this case the p phase would have a slightly different

composition, c 1 and the matrix composition would move towards a composition a,.

The resistance of the alloy of composition x would be ao at To and b1 at T, at the start of

ageing. The equilibrium resistance at T I is p 1 the mean of a, and c1 weighted according

to the proportion of the phases. The rate of change in the resistance will depend on the
44
Chapter 3 Precipitation kinetics

nucleation and growth rate and will be affected by several variables. A general form of

the curve is shown in figure c; showing a slow initial rate and then a faster change

followed by a decreasing rate towards the end of ageing. On cooling to room

temperature the final resistance will decrease according to the temperature coefficient

between T1 and T2.

The activation energy for diffusion, Q, plays an important role in precipitation process.

The rate of precipitation can be represented by the rate of change in resistance in the

following way:

dpldt = K exp - (A+Q/RJ)-----------------------------------3.35

where A is a second energy which is not necessarily constant with varying temperature.

Taking ~p from the start of precipitation at time t =O,

log Ap = log t- (A+QIRT) + log K---------------------------3.36


log t = B+ (A+QIRT)------------------------------------3.37
where B = log Ap- logK = constant for a constant change in resistance ~P-

The resistance method may make use of the classical four- probe technique, which

utilises two outer probes for current and two centre probes for voltage. The two

voltage probes are maintained at a fixed distance apart for all samples. The sample

dimensions are also maintained constant throughout a set of experiments. The

resistivity p is given by

p - RA/ L------------------------------------------3. 38

where R is the resistance measured on the samples, A is the area of cross-section and L

is the constant distance between the two voltage probes. The resistivity values so

measured are temperature sensitive, and also dependent on orientation of the samples.
45
Chapter 3 Precipitation kinetics

It is therefore necessary to keep these variables constant. The changes in resistivity

measured are very small and accuracy in measurements necessitates the use of a high

precision instrument that is capable of registering such a minute change.

The errors associated with the resistance method include inconsistent readings as a

result of variation in temperature and contact between probe and specimen. The

sensitivity of the instrument must be very high in order to measure the extremely small

changes in resistance values.

3.1.4.2 Constant Strain Rate Hot Compression Test:

Earlier experiments attempted to follow the progress of precipitation through its effect

on the high temperature yield strength<23 )_ This procedure is the isothermal equivalent

of using hardness measurements or yield strength at room temperature after periods of

ageing or tempering at elevated temperatures. A shortcoming of this technique,

however, is that throughout the temperature range of precipitation, there is often hardly

any change in the yield strength.

The inadequacy of the above technique led to the development of an alternative test

method. This was based on the measurement of strain to the peak stress <23 •28 >, Ep, This

method could be used to measure the times and amounts of both static and dynamic

precipitation. The peak strain is sensitive to both strain rate and temperature. In

situations where all other parameters are constant, the strain is a maximum in a given

material subject to precipitation soon after the initiation of flow (i.e dynamic

46
Chapter 3 Precipitation kinetics

~ .. p,
To

(a)

T1

T2

x.,. -e

p -- ------ (b)

-
- To

T,

T2
A .x ,.
B

p Cc)

TIME ~
-
Change in resistance during precipitation.
(a) phase diagram (b) resistance for equilibrium at three temperatures
and for quenching from T0 to Ta followed by ageing at T 1 (c) general
form of resistanctHime curve for ageing

Figure 3.8 Change in resistance during precipitation(Willard et a1< 22 >)

47
Chapter 3 Precipitation kinetics

precipitation). By determining the dependence of peak strain on strain rate and

temperature, the precipitation start time can be detected more sensitively than the

conventional techniques.

Figure 3.9 shows three flow curves obtained in a 0.018%Nb low carbon steel, after

successively longer ageing times. It is observed that by increasing the holding time

before deformation, a distinct change in Ep is observed.

A large holding time results in a small peak strain<23 >_ This corresponds to the increase

in the amount of static precipitation during the holding period and therefore decreasing

amount of dynamic precipitation during the work hardening interval. A decrease in the

amount of dynamic precipitation in tum decreases the amount of retardation of dynamic

recrystallisation and therefore the peak strain. In the absence of static precipitation, a

constant and maximum Ep is obtained associated with the maximum amount of

precipitate being available to inhibit nucleation of dynamic recrystallisation during the

rising portion of the flow curve. An increase in the static precipitation, leads to a

decrease in potential for dynamic precipitation, and therefore a decrease in Ep. Once

all the element has been precipitated out, before deformation, there is no further change

in Ep. Thus the end of the first plateau (marked in Figure 3.9) corresponds to the

precipitation start time and the beginning of the second corresponds to the precipitation

finish time. From this data a precipitation time-temperature diagram is constructed,

illustrated in Figure 3.10.

48
Chapter 3 Precipitation kinetics

45
T•mp .• °C
• 875
+ 900
40 A 925
a 975
o 1025
35
...
"'"'
~ 30
:;;
"
,c
w
a. 25
l=
z
c
a: 20
:;;

15
0 .Q. Ooi-0-0- 0 0 - - -
~
10'-------'------:------'-----_.__ _ ___,
10 1 10 1 10 3 104 105 10'
HOLDING TIME, s

Figure 3.9 The peak strain as a function of the holding time at temperature before
deformation. The first plateau indicates precipitation start time and the second,
the finish time(Weiss and Jonas( 23>).

,o,ol /
/
/
,;
/
/

•/
/

.
u
1000
I
I
/

(
w
a: I
...=> I
p
\
<I 950 s P,
a:
\•
"':I
Q.

...w 900
\. "·"'."
850
\
·, UNDEFORMED
0 05-,. C , 0,035% Nb

800 I 2 3 4 5 6
10 10 10 10 10 10
TIME

Figure 3.10 A precipitation time -temperature diagram for as obtained from data
in figure 3.9 (Weiss and Jonas( 23 >).

49
Chapter 3 Precipitation kinetics

It has been observed this method apparently indicated " faster kinetics" than other

techniques< 23 )_ However this was explained as due to the differences in the amount of

prestrains used, the alloying content and other factors. The Ps line for dynamic

precipitation as deduced by this method is 'C shaped' and is located well to the left of

the others indicating that the curve represents the limiting case for static process when

large prestrains are applied at high strain rates.

3.1.4.3 Hot Torsion Test:

This method is similar in approach to constant strain rate hot compression testing.

Torsion samples are austenitised and then cooled to the testing temperature. The

testing is commenced within a specific period of time which prevents initiation of static

precipitation of the carbide /nitride. The samples are tested at varying strain rates and

temperatures. The flow curves are plotted between equivalent stress and strain as

shown in Figure 3.11. The equivalent strain at the surface is given by

&eq = r I (3)112 = R 0 I (3)112------------------------------3.39

where y = surface strain, R = specimen gauge radius, 0 = angle of twist and L is the

gauge length.

The equivalent strain is given by

&eq - rI (3/12 ----------------------------------------3.40

and equivalent stress is given by

a eq = T (3)1 12------------------------------------------3.41

50
Chapter 3 Precipitation kinetics

.084 Al - .016 N

f 150

-
~

Ill
6.10" 1
J.I0· I
Ill
lo.I
~ 100
Ill

I-
z
II.I
..J
<r
> ,o
3
0
II.I
T•9oo•c

0 0., 1.0
EQUIVALENT STRAIN

Figure 3.12 Flow curves in terms of equivalent stress versus equivalent surface
strain for a low carbon steel determined at 600 °C at nine strain rates (Michel and
Jonas<21>).

where -r is the surface shear stress such that

a eq = T (3/12 I 21rR3 ( J+m+n)-------------------------3.42

where T is the measured torque, m is the strain rate sensitivity index given by (8 ln TI

cS ln 0 )0 and n is the work hardening exponent given by ( cS ln T / cS ln 8)0

The flow curves in figure 3.11 indicated a peak in stress (indicative of dynamic

recrystallisation) and the peak strain increased with the strain rate and decreasing

temperature. Figure 3.12 represents the plots between peak strain as a function of strain

rate. Similar to the hot compression technique, the start and finish times of

precipitation can be obtained from the hot torsion tests.

51
Chapter 3 Precipitation kinetics

./tl/
70

j
875 • C

..."'"'
60
//j
--·
0.05"/o C, 0.0!15Nt, ----·--

-·-1·-·
...
a::

"'
~
...
Q.
50 _,,,.,,,,.-•'• /•
/. EFFECT OF
/ SOLUTE
...
0
• EFFECT OF
z
:.i / FINlc: PPT'S /•
~ 40
"' _/ ./
!IO
----,·--
EFFECT OF /
/ 0.0 5 "1eC

COARSE PPT'S ~

20 ---~ j ------- .

-4 ,ciz
_,
10 10
STRAIN RATE i 1
fo)

Figure 3.12 Flow curves in terms of equivalent stress versus equivalent surface
strain for a low carbon steel determined at 600 °C at nine strain rates (Michel and
Jonas< 27>).

Following this chapter on the discussion of precipitation kinetics and the experimental

methods of measuring these kinetics, is Chapter 4, which will focus on strain ageing and

factors affecting the process.

52
Chapter 4 Strain ageing of low carbon steels

CHAPTER4

STRAIN AGEING OF LOW CARBON STEELS

53
Chapter 4 Strain ageing of low carbon steels

STRAIN AGEING OF LOW CARBON STEELS

This chapter defines the process of strain ageing and the effects of strain ageing process

on the mechanical properties. Subsequently, discussion is focussed on the factors that

affect strain ageing, with special emphasis on the compositional effects.

4.1 GENERAL FEATURES

Strain ageing is associated with the yield point phenomenon in which the strength of a

cold worked material increases and the ductility decreases on heating or ageing at a

relatively low temperature after cold working<4. 6>. Features of this phenomenon are

summarised in Figure 4.1. If a sample is left at room temperature after it has been

loaded to a stress beyond its elastic limit, on further testing the sample would show an

increase in the strength properties while recording a decrease in the elongation. This is

the main manifestation of the strain ageing process. The occurrence of strain ageing is

due primarily to the locking up of dislocations by interstitial carbon and nitrogen

atmospheres; although, heterogeneously nucleated precipitates may also play a ro1e<29,


30-31)

Prior to describing the effects of strain ageing on mechanical properties, it is necessary

however to define a few of these mechanical properties. Deformation in materials can

be of two types - elastic and plastic. In elastic deformation, the material regains its

original shape after the load is removed. This type of deformation is called elastic
54
Chapter 4 Strain ageing of low carbon steels

deformation (marked by OE in Figure 4.1). However there is an upper limit to loads

that can result in such elastic deformation. When loads beyond this limit are applied,

the material begins to undergo a permanent change in shape. This is plastic deformation

(all deformation beyond point E in Figure 4.1), which involves the large-scale

generation and movement of dislocations.

AY =chan1• in yield str•ss due to strain a1ln1


AU =chan1• in U.T.S. due to strain a1in1
AE - chanp in elonption due to strain a1in1

STRAIN

Figure 4.1 Stress-strain curve for a low carbon steel strained to point A, unloaded
and then restrained immediately-curve (a) and after ageing-curve(b)( Baird<29>).

Engineering stress is defined as the load borne per unit area.

cr - P/ A -------------------------------------------4. 1

55
Chapter 4 Strain ageing of low carbon steels

where P is the applied load and A is the initial cross--sectional area of the stressed

material. Engineering strain is defined as the ratio of the change in dimensions of

the specimen, from its original length.

E= (lr-10 ) / 10 --·-----------------------4.2

where lo is the original length of the test specimen, lr is the final length of the

specimen. Strain is usually expressed as %elongation.

Ultimate tensile stress: This is the maximum stress that a material can withstand

without failure (marked as U in Figure 4.1 ).

Yield stress: This is the minimum stress at which the material starts to yield plastically

(i.e a permanent change in shape).

Work hardening exponent: This measures the capacity of the material to harden as

deformation proceeds beyond the plastic strain at yield stress. Strain hardening occurs

due to multiplication of dislocations, which eventually become entangled, therefore

causing a resistance to further deformation and thereby resulting in an increase in the

strength of the material.

Figure 4.2 shows the variation in the mechanical properties as ageing proceeds at room

temperature. It is observed that the yield-point elongation and yield stress increase

continually from the commencement of ageing, while there is little change in the tensile

strength or ductility even after ageing for a week<27-33 >_ Strain ageing leads to a general

rise in the level of the stress strain curve and an increase in the rate of work hardening

during the early stages of straining after ageing. The total strain at fracture (%reduction

in area), is reduced by the ageing process. Any process which affects the level (higher

56
Chapter 4 Strain ageing of low carbon steels

stress values) of the stress-strain curve more than the work hardening rate leads to a

decrease in the absolute ductility as well as the uniform elongation<29-33 >_

Various experiments have proven that the effect of ageing on the mechanical properties

is quite dependent on the amount of working (i.e the amount of plastic strain) the steel

has been subjected to <4-6• 29•30• 31 • 33 >_ It has been observed that the changes in the tensile

strength and ductility in a more heavily worked steel was much more pronounced than

in the lightly worked steel(31)_ By contrast, the change in the yield stress and the yield

point elongation becomes less as the amount of working increases. The yield point

elongation is affected by the amount of working in a similar manner as the yield stress,

the maximum change occurring after a low degree of plastic working. A more detailed

discussion of the effect of strain ageing on mechanical properties is presented in section

4.4 of this chapter.

It has been generally observed that steel has a higher solubility for nitrogen than that for

carbon. Specifically, nitrogen solubility decreases from 83.9 ppm to 0.079 ppm at 300°

C and 25° C, respectively while carbon solubility decreases from 6.8 ppm at 300° C to

0.003 ppm at 25° c<2 - 6• 14• 29 -33 >_ As a result of a higher solubility of nitrogen coupled,

with a very slow and less complete nitride precipitation during cooling, nitrogen plays a
. the stram
greater ro1e than carbon m . agemg
. of plam
. carbon stee ls(2-6' 14 29-33)
' .

57
Chapter 4 Strain ageing of low carbon steels

ZDAYS IWEEK IMOHTH 3MONTHS IVEU

TOTAL ELONGA"TION (2°G.L)


50

>
~---.---1--1---- f-x-1-l
5-40
~~
..,-
:,
0
30
.

1-1-1 X I
X
UNIPORM ILONGATICN
X

X .
I
x I x-l-:--it-l-4
X I

-g
! 201-----------------'
"'a

t;z
~
ULTIMATE TENSILE STRENGTH
;~.,. 195
19
I
l
1 - i - - x - - x.,..
YIELD
l ,._
t
POINT ELONGAT::.:-+-t
I
x
-·-~ I

u;,!.
~-
i-
l
I
/--x x x--x----1
X
x I
X
~
w 01----="---------...__,,_.,,.,,.
YIELD S T R E S ~ :
. '
X

JC X l
X/ X
•/x
. /.,_,,...
100

8 L------i......----..__ _ ____,,......~
0 100 1000 10,000
AGEING TIME (HOURS)

Figure 4.2 The effect of ageing on the properties of mild steel temper rolled
1.4% ( B.B Hundy <4>)

Strain ageing is an important phenomenon that needs to be controlled since it not only

increases the yield stress after ageing and results in a decrease in ductility, but it also

gives rise to stretcher strains or Luder bands due to localised heterogenous deformation.

The process is also responsible for the occurrence in the stress -strain curve of what is

termed the Portevin-LeChatelier effect or dynamic strain ageing. This is a

manifestation of the non- uniform character of plastic flow< 35 -49) and is described in

detail in section 4.5.6.

58
Chapter 4 Strain ageing of low carbon steels

There are a number of theories to explain strain ageing, that are described in the

following section.

4.2 THEORIES OF STRAIN AGEING

Early theories of strain ageing assumed that the process was either due to some form of

strain-induced precipitation within grains or reformation of grain boundary films which

had ruptured during initial yielding<50• 51). However these early theories could not

account for all the observed changes that occur during strain ageing and therefore were

generally not accepted. A more applicable theory that was proposed later was the

formation of solute atmospheres around the dislocations pinning them, thus leading to

an increase in the yield stress. The kinetics for this atmosphere formation as suggested

by Cottrell and Bilby< 52) was based on an assumption that a driving force existed which

attracted the solute atoms towards a dislocation. The driving force, generally referred

to as the drift force, is associated with a decrease in the dilational energy of a solute

atom in sites near the tension side of an edge dislocation as represented schematically

in Figure 4.3. The decrease in the potential energy towards the core of the dislocation

gives rise to the drift force attracting the solute atoms towards the dislocation. It has

been shown by Nabarro and other workers, that a similar force would exist near a screw

dislocation also <50)_

The number of solute atoms that would reach a dislocation (edge or screw) in time t; Nt

is given by

N t = Tl o 3( 1t /2) 113 L (ADt/RT) 213 ---------------------4.3


59
Chapter 4 Strain ageing of low carbon steels

where 110 is the average concentration of solute (atoms /unit volume); L is the

dislocation density; D is the diffusion coefficient of the solute atom and A defines the

magnitude of interaction between the solute atom and the dislocation.

'Sr----r~-r~----liliC-
>-
<.,
Ck:
w
z.___...,,,,
w ...../
,,.... Core region
I
I
I

Figure 4.3 Schematic diagram showing potential energy of a solute atom as a


function of the distance R from the centre of a dislocation (J.D Baird<50>).

The value of A has been estimated in the range of 1.4 to 3.0 x 10 ·20 dyne cm 2 for both

carbon and nitrogen in iron <50>. It was estimated by Cottrell -Bilby that a dislocation

would be saturated with carbon when 1-2 atoms of carbon per atom plane of dislocation

had segregated to it< 50, 52>. Useful information about dislocation densities L, can be

estimated from this equation, if the other parameters are known. Equation 4.3 can also

be used to estimate the atmosphere densities at various rates of strain ageing (rate of

strain ageing would be given by rate of increase in yield stress with time) if the

dislocation density can be measured experimentally. However, this equation is relevant

only at atmosphere densities below the saturation level as stated by Cottrell and Bilby
(52)
As saturation is approached, back diffusion from the core region tends to

counterbalance the inward flow of solute atoms to dislocations due to the drift force.

60
Chapter 4 Strain ageing of/ow carbon steels

Further, it has been suggested that relief of the dislocation stresses by the segregated

atoms may reduce the drift force <51 )_

Harper modified the above equation so as to accommodate the decrease in the

interstitial solute content surrounding the dislocations as strain ageing proceeds< 50, 51 ' 53 )_

Based on an assumption that the rate of segregation of interstitial solute atoms at a

particular time would be proportional to the solute concentration remaining in the

matrix between dislocations, the following equation was derived< 51 , 53 ),

W= N!Nt = 1-exp [ -3L {1t/2) 113 (ADt/kT) 213 ]--------------4.4


where W is the fraction of solute that has segregated to the dislocations, N is the number

of solute atoms that would reach a dislocation and N1 , the total number of solute atoms

present and the other terms are as in equation 4.4. Equation 4.4 can be restated as

In (1-W) = - (t /-c) 213 ------,---------4.5


where -c is a constant. The preceding two equations do not take into account the

saturation or back diffusion effects and therefore hold good only for low atmosphere

densities and the early stages of ageing.

The kinetics of the ageing process has been studied by a number of investigators by

measuring the change in the concentration of interstitial solute remaining in the ferrite

lattice. However, this is based on the assumption that all solute leaving the lattice

segregates to dislocations. The change in the solute concentration can be measured

either by the internal friction method (Snoek peak) or by electrical resistivity change.

The former technique has generally found to be more accurate <51 , 54 )_ Extensive work by

a number of investigators (listed in Table 4.1) revealed that the ageing kinetics
61
Chapter 4 Strain ageing of low carbon steels

predicted by Cottrell- Bilby equation was satisfied only up to about 50% completion of

the total ageing process, while Harper's equation (equation 4.4) held good up to near

completion of the ageing process< 51 >. In general, the activation energy of the ageing

process was found to be equal to the activation energy for diffusion of the interstitial

solute present<Sl)_ Although several workers agree with the adequacy of the Cottrell -

Bilby and the Harper equations, one serious limitation is that these cannot be applied to

strain ageing situations with a high solute concentration <51 >. From Table 4.2, it is

evident that Carswell's studies on Fe-N alloys containing as much as 0.003% nitrogen

have shown that the Harper equation satisfactorily explained kinetics <50, 51 >. It is also

apparent from Table 4.2 that there is no mention of alloys containing more than 0.003%

of which successfully obey Harper equation <50, 51 >. This limitation is apparently due to

the fact that strain ageing can be associated with more than atmosphere formation.

Other theories therefore seek to explain strain ageing in two steps - the first step is

associated with atmosphere formation and the second, with precipitate formation

following atmosphere formation.

4.2.1 Atmosphere formation

A lack of knowledge of the exact equilibrium form of a solute atmosphere at the core of

a dislocation currently hinders the formulation of a satisfactory theory for the proposed

first stage of strain ageing <51 >. As a matter of approximation, the Maxwell - Boltzman

law gives the solute concentration at a few atomic spacings away from the dislocation

core.

C =C O exp ( - ~ /kT)------------------------------------4. 6

62
Chapter 4 Strain ageing of low carbon steels

where C0 is the average concentration in the crystal and ~ is the interaction potential or

binding energy between the solute atom and the dislocation.

For an edge dislocation,

~ = A (sin 0 ) /R ---------------- ---4.7

where R and 0 are the polar coordinates of the solute atom with respect to the

dislocation line and A is the solute - dislocation interaction parameter. These equations

give an impossibly high solute concentration when R ~ 0< 51 >. It was suggested by

Cottrell and Bilby that the relief of dislocation stresses by segregated atoms would give

rise to a saturation condition in the core region {R<1OA0 ) when it contained one

solute atom per atom plane of the dislocation< 51 >. It was postulated that these atoms

would tend to arrange themselves in a line just below the extra half plane of an edge

dislocation to give a condensed atmosphere. Surrounding this, a dilute Maxwell -

Boltzmann type distribution of the solute would be present, containing one additional

solute atom per plane of dislocation.

However it was suggested by Beshers that the distribution of solute atoms near a

dislocation core follows a Fermi-Dirac distribution dictated by energy levels rather than

a Maxwell-Boltzmann type of distribution<51 >. This would imply that solute atoms have

various energy levels associated with them and arrange themselves around the

dislocation core in the lowest possible energy states. At higher temperatures some

atoms raise their energy levels due to the additional thermal energy supplied. A serious

limitation of this latter theory is that it is not sufficiently quantitative to calculate the

amount of solute atoms near a dislocation core with any confidence.

63
Chapter 4 Strain ageing of low carbon steels

In the dislocation potential field, the interstitial solute atom acquires a drift velocity 'v'

that would allow for it to be attracted towards the dislocation.

v = - (DI RT) d~/ dR-------------------4.8

This will be the controlling step governing the early stages of strain ageing when the

concentration gradients arising due to segregation are negligible <51 >. In this case the

Cottrell - Bilby equation should be applicable. But with time, the solute atmosphere

would have attained an appreciable value, causing the onset of back diffusion. This

counteracts the drift flow, and at equilibrium, the two fluxes of solute atoms balance

each other. Further, atmosphere formation is results in the segregation of solute from

the matrix, thus causing the depletion of the interstitial solute from the matrix

(surrounding the dislocation). Cottrell - Bilby's equation does not account for this, thus

explaining the deviations from the equation in the early stages.

Bullough and Newman studied the deviations from the Cottrell-Bilby kinetics and came

up with a theory in which the potential field around a dislocation is considered to be

radially symmetrical, and not as described by the Maxwell - Boltzmann distribution<55 ,

56>. They also assumed that the interaction potential between a solute and a dislocation,

is not affected by the solute atoms that have already segregated to the core. They

showed that up to 30 - 40 % completion of the ageing process, the predicted kinetics

followed the Cottrell - Bilby equation<51 , 56>. In the early stages of the process, diffusion

assisted the segregation process by supplying solute to a depleted zone around the

dislocation, but in later stages, back diffusion from the core retarded strain ageing.

Therefore the overall kinetics of the strain ageing process did not obey the Cottrell-

Bilby equation.

64
Chapter 4 Strain ageing of low carbon steels

Atmosphere formation was found to account for only a small fraction = 10% of the
solute which was predicted to segregate to the dislocations< 51 • 56 >. However, in most

strain ageing cases, experimentally it has been observed that 100% of the solute was

found to segregate. Therefore it was concluded by Bullough and Newman that

atmosphere formation, alone, was not responsible for strain ageing< 51 >. During the stage

of atmosphere formation, back diffusion from the core was found to be more

predominant than the depletion of the matrix surrounding the dislocation in slowing

down the rate of solute atom segregation with time. This meant that the correction to

the Cottrell-Bilby equation as postulated by Harper was not important in this case and

does not fully explain the observed kinetics of Bullough and Newman as evident from

Table 4.1. It was therefore believed that strain ageing occurs in two stages, the first

being atmosphere formation and the next, precipitation on dislocations.

4.2.2 Precipitation on dislocations

A review of literature shows that there is very little agreement between the predicted

and observed kinetics in the stage of atmosphere formation despite the initial success of

the Cottrell - Bilby equation. However, there appears to be a good agreement between

the observed and predicted kinetics in the precipitation stage of the ageing process.

The densities of solute atmospheres formed during strain ageing as experimentally

estimated by internal friction and electrical resistivity techniques are often well above

the saturation limit for atmospheres predicted by the Cottrell - Bilby equation< 51 >. In

65
Chapter 4 Strain ageing of low carbon steels

Table 4.1 Measurements of strain ageing kinetics in quenched Fe-C and Fe-N alloys (J.D Baird<51 >)

.I .I •Asiq Solute Content (wt.%)


..
Calculated Dhlocatlon
. . ..
Demlq,ftf.Solllte
lnvestlptor !Method• I Preatrain % Temperature, 0 C. C N Dcmlty cmkm1 x 10-1• Atoma/Atom •·
Plane .,
H.\Jut!R lU ' I.F. 5 I 30
II 0·015 -- 7·7 13
-
i
10
15 i 30
39
0·009
0·013 - 10·2
15·7
6
6
DAHL AND Luca
!
E.R. 4·5 120 I 0·015 - . 0·82 75t
II
i 5·1 ISO 0·015 -- 1·67 45+
11st
;
5·3 90
I 0·018
I
0·95
THOMAS AND La.u:
I
I.F. 7 i
24-54 I - 0·015 I 7·7 11·5
---
IT
7 35-67 0·015 6·6 15·6
7 35-67 0·0095 I S·O 13·1
~
I 7 I 35-67 0·0055 4·0 9·-4
I
i
I -- II
PITSCH AND LOCJCB HI
I I.F. 4·8
4·8
I
I
90
120
0·01075
0·01575
2·6
2·2
-40•5
58
I
i
i
4·8
4·8
149
170
0·01835
0·01915
-- '
I 3·0
2·4
66
65·5
~CUN AND HBu.ml 111
i
I.F.
i
12·2
12·2
20-1-40
2~140
0-0203
0·0203
-- I 13·6
13·8
13
13
'
'
!
12·2 20-140 . 0·0127 - 13·2 9
12·2
12·2
5-2
20-140
20-140
20-140
0·017
0·0185
0·0185
--- 13·2
17·
11
11
10
15
WU.SON AND RUSSBLL ' E.R.
30·3
4
20-140
20 and 60
0·0185
0·01
-
0·004
36·4
10
4
12
I

KosTl!ll, BANGIRT AND HAHN 41 I


I
I.F. 33
I
40 0·018
- -..
- 30
-- -- ·- -. . ..

66
Chapter 4 Strain ageing of/ow carbon steels

addition, the alloys that have been shown to exhibit strain ageing (Table 4.2) contained a

high concentration of carbon and nitrogen. Atmosphere formation alone could not

account for this solute content. Therefore it was assumed that the solute atoms, in

addition to the formation of atmospheres around the dislocations were also

heterogenously precipitating on dislocations. It was suggested by Bilby that the solute

atoms drawn into the dislocation core could diffuse along the core to feed the precipitate

particles spaced at intervals along the dislocations(51 • 52• 57>.

However the fact that atoms in the dislocation core could feed the precipitate particles,

was refuted by Bullough and Newman on the basis that the binding energy of a solute

atom to a dislocation is greater than that to a precipitate particle(53 >_ But this was

disproved by Harper, Thomas and others, who studied quenched alloys; and argued that

there are too many solute atoms to be accommodated in the low energy sites at the

dislocation cores. Therefore, once these energy sites are occupied, further atoms,

which diffuse to the dislocation due to the drift force, occupy higher energy sites. This

makes the situation energetically favourable for atoms in the higher energy sites to

diffuse along the dislocations to feed precipitate particles (lower energy).

67
( 'hapter 4 Strain ageing <~(low carbon steels

Table 4.2 Measurements of strain ageing kinetics in slowly cooled or quench-aged Fe-C and Fe-N alloys(J.D Baird<st >)

Solute Co11tent Calculated fJul ~ i ;·: ~ _,?


0 ~:: _·

IDwatlptor Method' I PrestraJn Heat Treatment I A,m1 (wt.%) DJalocadon of Sohate llatdill ·.. '.;
% Temperature, °C Dt111lty cm/ Atoma/Atom' fo8owl4 ·., ·.
C N cm1 X10-" Plant . , . ·•..
-----1---1---1
E.R.
-------!
Slow cooled to 200°c, I
1
1

20 and 60 - 0·0022 5•1


,, .-
"--"
0- - , ·• •

WIUON
Rusm.t'
AND4
aged anel then quenched.
Wet hydrogen treated
'"l
. . 1/
UllllaP ' , . '
nn,... · .,,
aAMIJ .. ....,;
4
8 hr, quenched from
500°C. .
20 and 60 0·0005 - ..... 4
~.·
aptoeG% ,·

CA;lswll.L" I.F. 3 Quench aged 180° 20 0·0031 9·6 N01 • } / , I

3 Quenclt aged 120° 20 I 0·001 JO N01


3 Quench aae<f 80° 20 0·0009 10·7 N~
Harper I

3 Quench aged 60° 20 I


I
0·00083 11·4 "'01 eqllldoa
2 Quench aged 120° 20 I 0·001 7·5 N~ ltfint..
2 Quench aged 80° 20 0·0009 8·0 ..,H
2 Quench aged 60° 20 1= 0·000831 8·6 NO~
IKosra, BANODT I.F. 33 Normalised. 40
j

j 0-0032 - I Harper
I
AND HAHN .. I equation
i
•E.R. •Electrical resistivity, I.F. ""'.Internal friction,

68
Chapter 4 Strain ageing of/ow carbon steels

All of these workers have based their analysis on the kinetics of the precipitation on the

dislocations, assuming that the dislocation core is a cylindrical region in which the

diffusivity of carbon or nitrogen is sufficiently high to be taken as infinite<51 , 57 >_ While

Bullough and Newman assumed various rates of transfer of solute atoms across the

core - precipitate interface< 51 • 53 >, Ham assumed that the transfer of solute atoms from

the dislocation core into the precipitate was sufficiently fast for the dislocation core to

act as a sink for these atoms< 51 >. These two calculations agreed at higher rates of

transfer. Apart from the initial (small) time for atmosphere formation, the kinetics are

predicted to follow

In (1-W) = - (th )-------------------------------------4. 9


which again, however, does not agree with the observed kinetics.

Bullough and Newman further assumed that the drift flow is important in controlling

the kinetics of the ageing process, more importantly the rate of growth of precipitates

on dislocations. Accordingly, they set about finding a model of the precipitation that

would explain the observed kinetics. Rod- shaped precipitates were assumed to form

along the dislocations centered at the dislocation core <50, 51 >. The chemical binding

between the precipitate and its elastic interaction with the dislocation strain field was

assumed to stabilise the precipitate. In order to find the exact order of the time

constant, they also assumed that the transfer of solute atoms across the matrix-

precipitate interface was the rate-controlling step. This transfer also decreased as the

precipitate grew, possibly as result of an increasing hydrostatic pressure build up inside

the precipitate. These assumptions led to the observed kinetics following an equation

close to that developed by Harper. The above explanation by Bullough and Newman
69
Chapter 4 Strain ageing of low carbon steels

assumes a high activation energy for the interface transfer process. However, without

this assumption the process kinetics would follow equation 4.9.

Equation 4.9 does not give the observed kinetics of the ageing process if a uniform

dislocation density is assumed. In practice, however, it has been observed by Keh, that

the dislocation density is not uniform throughout the materia1< 5s, 59 >_ Keh has shown

that in regions of clusters and cell walls, which is assumed to occupy one-fifth of the

volume, the dislocation density is about four times the average dislocation density.

Therefore, the density in the remaining four-fifths of the volume must be only about a

quarter of the average density. The time constant (i-) was found to be inversely

proportional to dislocation density <51 >. The overall rate of segregation m an

inhomogenous structure as predicted by Bullough and Newman can be given by

1 -W = 1/5 exp (- t / 't 1) + 4/5 exp (-t /16 't 1) ------------4.10

and a plot between log log (1 -W) and log ( t / 't 1 ) gives a reasonably good straight

line relationship.

Observations by Nacken and Heller showed that precipitates formed as a result of

segregation of the solute atoms on dislocations did not redissolve upon heating as easily

as those produced by quench ageing< 51 >. And overageing was found to be more rapid

and complete during quench ageing than during strain ageing. This indicates that during

the early stage of precipitation on dislocations, the solute atoms are more tightly bound

than if they were in the form of discrete particles formed during quench ageing. These

observations support Bullough and Newman's hypothesis that a dislocation - precipitate

is a specialised form, partly stabilised by interaction with the elastic strain field of the

70
Chapter 4 Strain ageing of low carbon steels

dislocation. With these assumptions, they found that they could explain the kinetics as

shown by Harper's equation well.

Some deviations from the Harper equation were noticed, and these deviations were

attributed to the precipitation of solute between dislocations. Studies by Leslie et al,

confirmed this phenomenon by observations made in some commercial steels, where

fine precipitates were noticed between dislocations< 5B, 59 >_ This precipitation has been

observed to occur up to very high prestrains, especially in regions of low dislocation

density.

Hundy developed express10ns relating the time of strain agemg of steel at room

temperature to that at higher temperatures<4' 5>. Assuming that carbon causes strain

ageing, the equation is

Log 10 t r!t = 4400 (1/f r - 1/f) - log 10 Tff r·----- 4.11

While, for nitrogen, the equation is

Log 10 t r!t = 4000 (1/f r - 1/f) - log 10 Tff r ------------------4.12

In these equations, 'tr' is the ageing time at room temperature, Tr, and t is the time of

ageing at elevated temperature T. These equations are based on the assumption that the

number of solute atoms are the same at room and elevated temperatures, which has been

found to be true only up to about 200°C.

4.2.3 Effect of dislocation sources

Cottrell postulated that if a specimen is prestrained in one direction, all inclusions

capable of acting as dislocation sources under this stress system will be inactivated<60>.
71
Chapter 4 Strain ageing of low carbon steels

This should help in achieving a return of the yield point in ageing, if restraining is in

the prestrain direction rather than in any other direction. Studies by Kennett and Owen

have revealed that inclusions elongated by rolling act as more efficient dislocation

sources during transverse tensile straining than by straining in the direction of

rolling< 61 >. However studies by Tipper showed no effect of rolling direction on the rate

of return of the yield point<35 >_ Experiments by Mura and co-workers revealed that a

steel aged at 30°C until it showed a yield point on testing at the same temperature, did

not show a sharp yield point when tested at -196 °C after the same ageing treatment<62>.

This could be explained by the fact that inclusions or precipitate particles may also give

rise to dislocations by thermal stresses when the temperature of the specimen is

changed.

Suggestions were made by Leslie and Keh that the stress required to activate grain-

boundary or inclusion sources may be raised during strain ageing by the segregation of

solute atoms<63 >_ This could be responsible for the slow secondary rise in yield point

extension that takes place on prolonged ageing after all types of prestrain.

4.2.4 Factors affecting strain ageing:

The factors that determine the magnitude of strain ageing are:

• the dislocation density

• the concentration of interstitial solutes namely, carbon and nitrogen

• the time available for diffusion to occur

• the temperature at which diffusion takes place.

72
Chapter 4 Strain ageing of low carbon steels

Strain ageing can be of two types - static strain ageing and dynamic strain ageing<29).

Static strain ageing is referred to that process of ageing that occurs in a previously

strained material that is left exposed to some temperature (typically room temperature)

for a period of time. When this material is tested for mechanical properties after this

period, it is found to exhibit strain ageing (rise in strength or a decrease in ductility).

Dynamic strain ageing, on the other hand occurs during the process of straining itself.

The following sections describe the two phenomena.

4.2.5 Static Strain Ageing

Static ageing can be divided into four stages as described from the work of Wilson and

Russell< 56 ' 57 ), illustrated in Figure 4.4. Stage 1 of the process shows a rise in the yield

stress and the yield point elongation (Luders strain) while the other properties remain

unchanged. There is however a fall in the strain hardening exponent. During Stage 2

of the phenomenon, the yield stress continues to increase while the yield point

elongation remains constant. The flow stress rises while the rate of strain hardening

remains constant. In Stage 3, the strain hardening rate increases giving rise to an

increase in the ultimate tensile stress and a decrease in the elongation. However there

is a decrease in the uniform elongation. Stage 4 leads to an overageing effect, which

results in softening. Stage 1 is purely the yield point phenomenon while stages 2 and 3

reflect strain/age- hardening processes.

Wilson and Russell proposed that during Stage 1 of strain ageing, Cottrell atmospheres

were formed around the dislocations, which were sufficient to completely pin the

73
Chapter 4 Strain ageing of low carbon steels

STAGE I STAGE II STAGEIII

~ 8
" 6
"I: 4
i LOWER YIELD STRESS

... 4
Q
.... 2

i 01----+----r
~
# 0
~ -5 ELONGATION
-10
TO FRACTURE

0·04
i--;---r--r-"===*=:::::::"1
,ao-02
0

10 10 IOI I~ 105
AGING TIME AT 6QOC,mln

Figure 4.4 Four different stages of static strain ageing (Wilson and
Russell <56 •57>)

dislocations< 56 • 57 >_ These atoms were dispersed on further straining, thus resulting in

the freeing of dislocations while the other properties remain unaltered. In stage 2 of the

process, the hardening that comes about was postulated to arise from the pinning of

dislocations by fine particles(29• 30• 33 >. It was further suggested that it was easier to

activate fresh dislocations on further straining rather than unpin the already strain aged

dislocations. This led to Luders strain achieving a constant value and marked the

beginning of the strain age hardening stages 2 and 3. The onset of stage 3 was

delineated as the point when the fine precipitates became sufficient in size to accelerate
74
Chapter 4 Strain ageing of low carbon steels

strain hardening. During stage 4 of the process the increase in Luders strain was due to

the precipitates going into solution and segregation along grain boundaries leading to

an increase in the unpinning stress<64 • 65 >. The rate of ageing and the final yield and

tensile stresses increase with increasing solute content<29• 30>. The ageing index has

been defined as the difference in the tensile strengths or the yield strengths at room

temperature and the temperature of strain ageing< 4, 66>. Alternatively it could represent

the difference in the strength values at two different times, for example on receipt and

after storage at room temperature after a few days.

Static strain ageing has been found to be affected by the following factors< 4• 29• 30•
33).

o composition

o temperature

o pre-strain

o prior heat treatment

The effect of each of these parameters on strain ageing is discussed in the following

sections.

4.2.5.1 Effect of Composition

Solute alloying additions can contribute in two ways to strain ageing. There are solutes

which lock dislocations and diffuse quickly to produce strain ageing while there are

75
Chapter 4 Strain ageing of low carbon steels

other solutes which can alter the concentration or mobility of those solute atoms that

directly produce strain ageing( 33 )_

The first class of solutes must have a high diffusion coefficient such as interstitial

alloying elements in ferrite, i.e, carbon, nitrogen, oxygen and hydrogen. While oxygen

and hydrogen do not appear to be the main elements causing strain ageing, carbon and

nitrogen have been proven to be very important<3, 33 )_ The effectiveness with which an

element influences or causes strain ageing in steel is based on

• its solubility in ferrite;

• its diffusion coefficient and

• the degree to which it locks the dislocations.

a) Nitrogen:

Steel has a very high propensity to absorb nitrogen at high temperatures from the

atmosphere during steelmaking<3, 33 • 66' 67 ). Once nitrogen is absorbed into the metal, it

is difficult to remove the element. Vacuum degassing removes only 10-20%<66' 67 ).

Figure 4.5 shows the Fe-N equilibrium diagram in the Fe-rich region< 3, 14). The

solubilities of the two main phases, Fe4N and Fe 16N 2 at various temperatures are shown.

The variation in the solubility of nitrogen in ferrite as a function of temperature, as

studied by Leslie is shown in Figure 4.6(3, 14). It is seen that for normal nitrogen contents

of 0.002 to 0.01 %, the solubility is only exceeded during cooling below about 300 °C.

However, below about I 50 °C, the mobility of nitrogen is quite low. The extrapolated

76
Chapter 4 Strain ageing of low carbon steels

solubility of nitrogen at room temperature is 104 to 10-5 wt%. The higher equilibrium

solubility of nitrogen and the reduced rate of precipitation of nitrogen in the presence of

manganese results in an amount of nitrogen in solution which is at least one order of

magnitude higher than the amount of carbon in solution for most commercial grades of

steel 0 3>_

Studies using internal friction have revealed that peaks corresponding to nitrogen and

carbon disappear as strain ageing proceeds while a new peak appears at a higher

temperature, as shown in Figure 4.7. This new peak was found to be due to the carbon

and nitrogen atoms, which lock up the dislocations. It was believed that the

supersaturation of nitrogen was relieved by the presence of fresh dislocations.

However, later work by Thomas and Leak showed that the equilibrium solubility of

nitrogen is less in the presence of fresh dislocations than in their absence, as illustrated

in Figure 4.8< 69' 70>. It has been shown that the binding energy of a nitrogen atom to a

dislocation is about 0.75eV while the binding energy in the precipitates of Fe4N and

Fe16N 2 is 0.25 - 0.35 eV in the former and even less in the latter. Thus the N atoms are

more firmly bound to dislocations than to iron in nitride particles. This suggests that

during straining, the nitride precipitates dissociate, thus providing a steady source of

nitrogen for strain ageing. Support for this theory came from a study by Dijkstra who

found that the solubility of nitrogen in the presence of nitride particles was 0.012% at

300° C and 0.02 wt% at 400° C, while solubility in the presence of dislocations was

only 0.002% at 300° C and 0.006% at 400° c<68>.

77
Chapter 4 Strain ageing of low carbon steels

N1tro9e"I la,. •1.1


C :;2 :>? '.)4 05

900

800

u 6GO
•..,, 0.0"
•~ 500
.,...
ci
f 400
~

1000~--,~-~~~..,._.....,,..~---~_,.......- - - J
002 o 04 006 ooa 010 o ,z o 14
N1tr09en ( wt%J

Figure 4.5 Ferrite field showin? equilibrium between Fe-N2 gas, Fe-Fe4N and Fe-
Fe16N2 (Bever cited by Leslie<3, 1 >)

IOO 0-IOO
-I
0050
0
i: 0025
z !z1£1
0-010
8-2
~ §
3
';t
e.,
I,) -3
9 0·00025
.4L,,__......__ _ _ ___,..,,.--~-~-=---="'==----=-O·OOOI
3·2 2·8 2·4 2·0 1•6 1·2 0·8 0·4
I xl01
i'

Figure 4.6 Solubility of carbon and nitrogen as a function of temperature


(Leslie<3>).
78
Chapter 4 Strain ageing of low carbon steels

Freq. • I cp•.

DI&

.00

Temp. •c

Figure 4. 7 Internal friction curves of a dual solid solution of carbon and nitrogen
in ferrite vs temperature. Curves a and b are the internal nitrogen and carbon
peaks(122).

100 125 150


TEMPERATURE
250
:c
200
O•OI-..:.:;::.;::;...--=.::;.=,-~~--~;;;;.-.~.,-~,....,,..,
,,,,,
,
~ ,,,,,
,
0 ,,,,,
..... ,,,,,
,,,,,
z
,
~z
/

,,,,,
,
O•OOI ,,,,, _ _ 1N E0Ul.l8RIJM WITH
,,, ;
PIIECIPtTATED NITRIDE
/ IN EOUILIBRIJM
Q •0005 , ' ---- - WITH DISLOCATION
"" ATMOSPHERES
O •00O1'-.e=-"__2_.·..... s
6-~2""""".4~~2~-2-::---~26-::-o=--~,.
.
I ,c 10
T
Figure 4.8 Solubility of nitrogen in equilibrium with dislocations and precipitated
nitride as a function of temperature (Thomas and Leak(69>).

79
Chapter 4 Strain ageing of/ow carbon steels

These results imply that the effect of nitrogen on strain agemg may not be very

dependent on the prior heat treatment used. The diffusivity of nitrogen in ferrite is

given by the equation<3>

DN = 6.6 x 10 ·3 exp (-77826/RT) cm2 sec·•-----------------------4.13

where DN is the diffusion coefficient for nitrogen, R is the gas constant and T, the

temperature. DN was calculated as being 1.5 x 10 · 16 cm2 sec· 1 at room temperature (as

shown in figure 4.9), which corresponds to one nitrogen atom making a jump every

three seconds and in one month the diffusion distance would approximately be equal to

the distance between dislocations< 59 , 63 , 69>.

To summarise, nitrogen was found to satisfy all the three criteria for the occurrence of

strain ageing. The nitrogen atoms have quite a high solubility in ferrite at room

temperature, high diffusion coefficient in ferrite at room temperature and a high binding

energy to dislocations.

Fast and co-workers studied the effect of nitrogen alone on the strain ageing in two iron

- nitrogen alloys, one containing 0.02% nitrogen and the other 4 x 104 % nitrogen< 29,
34)
Study of the hardness values after a 10% prestrain and ageing treatments of 2

hours at temperatures between 100°C and 700°C revealed higher hardness values in the

alloy containing higher nitrogen and therefore a higher degree of strain ageing. The

results are shown in Figure 4.10.

80
Chapter 4 Strain ageing of low carbon steels

TEMPERATURE,°C
25 50 75 100 I 50 200 250

101

16'
-10
10

-I I0-11
s
s 10
N
2

. 10-,,
-14
10

-15
10

3·0 2·6 2·2 1·8 1·4 l·O


+.·,o•

Figure 4.9 Diffusion coefficients of nitrogen and carbon (Leslie<3>)

150 _ _ _ _ _,Fc-0·02'YoN
zl40 - • ~
~130 ~
~120
~ Fc-0•040/oC
;,10 /0~~ •
i, --:::...-,- • "
90 F~-0·0001°/o~ "'-

80 ~
70~-------------'-------..;a
0 100 200 JOO 400 SOO 600 700
TEMPERATURE ,0 c

Figure 4.10 Strain ageing of iron -nitrogen and iron carbon alloys strained 10%
and aged for 2 hours at different temperatures (Fast cited by Baird<29>).

81
Chapter 4 Strain ageing of low carbon steels

Figure 4.11 illustrates results of the experiments on iron nitrogen alloys, which show

the changes in mechanical properties at two temperatures of 51 °C and 42°c<29, 70>. The

changes in mechanical properties are well correlated with the amount of nitrogen

dissolved in ferrite as measured by the internal friction technique.

51°C

Figure 4.11 Strain ageing of 0.029% nitrogen-iron alloy followed by change in the
yield stress(upper curves) and solute content of ferrite as measured by internal
friction(Thomas and Leak cited by Baird<29>).

Josefsson and Backstrom studied the influence of varying nitrogen contents in three

carbon saturated low carbon steels containing (in weight%) 0.02%C, 0.003%N;

0.03%C, 0.006%N and 0.03%C, 0.008%N<71 >. The rate of strain ageing increased with

the total nitrogen content as illustrated in Figure 4.12. Figure 4.13 shows the

behaviour of the three grades of steels mentioned above at 60°C. The amount of ageing

apparently has undergone no change, but the ageing time has decreased.

82
Chapter 4 Strain ageing of low carbon steels

80

70

60

..
E
50

e
";. 4 0
.,.
>-. .
<I
lO
0

2-0

1·0

I DAY 45 DAYS
4l 4' 41 ... ...
AGING TIME, m1n

Figure 4.12 Increase in the yield stress after prestraining 3% and 8% vs ageing
time at room temperature (Josefsson and Backstrom<11>) o-----o 0.02 C, 0.003 N ; •-
·•0.03C0.008N; /l-- /l0.03C0.008N. ----------0.002N(Wilson and Russell); _ __
_ 0.014%C+N(Wilson and Russell).

8·0

7·0

6·0

5·0

e
~...
,..
4·0
>-~
<I
5-0

2 ·0

;
IHR IDAY 45 DAYS
O·O
4·01 40 4' 4l 4l
•• 41
AGING TIME I m,n
41 47 41 4'

Figure 4.13 Increase in the yield stress after prestraining 3% vs ageing time at
60°C (Josefsson and Backstrom<71 > )-0-----0 0.02 C, 0.003 N; •--•0.03C0.008N; /l--
/l0.03C0.008N(water quenched from 250°C); ---0.002N(Wilson and Russell);-
_ _ _ _ _0.014 ¾C+N(Wilson and Russell).

83
Chapter 4 Strain ageing of low carbon steels

Taheri and co-workers studied strain ageing in various grades of steels with carbon

contents in the range of 0.05 - 0.08 wt¾ and nitrogen contents between 0.003 and 0.43

wt%< 72 >_ Figure 4.14 compares the strain ageing susceptibilities of these steels which

were aged at 100 °C for 20 seconds after 6% strain. It is evident that as the level of

solute nitrogen content increases, strain ageing increases (flow stress increases). It was

also proved that N was responsible for the decrease in the ductility, by computing the

activation energy of the ageing process from the following equation.

In (t2/t1) = Q/R (1/f 1 - lff2) -------------------------4.14

where T1 and T2 are two different test temperatures of strain ageing and t 1 and ti are

times of strain ageing at these temperatures which produce the same degree of strain

ageing, Q is the activation energy and R, the gas constant. The activation energy was

calculated to be 67kJ/mol by Taheri et al.

10u rr====-===::::::"'-----------,
---hlN,med
.
..~·······~········~·-······
80 ---hlN,•1-
. ...
- - - ~ - lo N, nwd
. .
60

40 ........ :- ..... .
. .
20 • • • • • • • • .~ • • • • •. ·-.-!:- • • • • • - • -:-. • ---- -- ~.• • • - - ---
.. ....,_ _.,._ _ _..,_ .
0 ..,__ _...., _ __.,.·_ __.

0 50 100 150 200 251


Aglng temperature re)

Figure 4.14 Effect of nitrogen content on the flow stress of low carbon steel aged at
100°C for 20 seconds after 6% strain (Taheri et a1< 72 >).

84
Chapter 4 Strain ageing of low carbon steels

Figure 4.15 shows the effect of nitrogen content on the decrease in total elongation due

to ageing.

I• S% ROLLINC REDUCTION

--
~ 30
AGED b MTHS AT ROOM TEMPERATURE

, "'
~
"'m:u
o 20

a JOf,IS t, OWEN • &ARNE TT ( 35)


• HUNO]

0-008 0·01 0-01 0-020


NITROGEN CONTENT( •k )

Figure 4.15 Effect of nitrogen content on decrease in total elongation due to ageing
(B.B Hundy<4> 1956).

• Carbon

The solubility of carbon in ferrite is given by the equation<3>

%C = 2.55 exp (-9700/RT------------4.15


while the diffusion coefficient is given by<3>

D c = 0.02 exp (-20,000/RT)---------------------------4.16

From Figure 4.6, it is seen that the equilibrium solubility of carbon decreases to below

10 -3 to 10 -4 wt% at a temperature just over 200°C. Above 200° C, the mobility of

carbon atoms is reasonably high and on slow cooling, equilibrium should be

approached if there are sufficient nuclei present for the precipitation to take place.

Studies using internal friction measurements indicate that it is easier for nitride

precipitates to dissociate than for carbides(Jo, 37 >_ However if carbon is held in

supersaturated solid solution by quenching or by slower cooling in the absence of


85
Chapter 4 Strain ageing of low carbon steels

precipitate nuclei, then the excess carbon should result in strain ageing. At a

temperature greater than 200°C, enough carbon should be in solution at equilibrium to

give some ageing directly. It is also seen that the diffusion coefficient of carbon is

slightly lower than that of nitrogen at low temperatures while it rises above that of

nitrogen above 390°C as shown in Figure 4.9. Locking of dislocations by carbon is

slightly weaker than by nitrogen.

Fast and co-workers observed the change in hardness in a Fe-0.04% C alloy after

ageing for 2 - 40 hours at 50 °C- 200°C and a prestrain of 10%<29>_ It was noticed that

there was no appreciable increase in the hardness even after ageing at 50°C for about

40 hours. Two hours at 100°C produced a slight increase while two hours at 200 °C

produced the peak hardness. This was as compared to the fact that a Fe - 0.02% N

alloy showed maximum hardness after ageing for 2 hours at 50 °C. Similar studies

using the internal friction method by Butlei13> and Stephenson <74 >, after ageing low

carbon steels and irons to 100 °C for 1 hour after 7.5% and 10% prestrains revealed

rapid increases in the yield strength as dissolved carbon was increased from 0 to

0.0001 % (Butler) and 0 to 0.0026% (Stephenson) and a slower increase thereafter.

Figure 4.16 illustrates results of studies by Butler. Even with as low as 0.0005%C,

ageing indices of 5000psi (35 Mpa) are obtained. Carbon was found to be as effective

as nitrogen in producing strain ageing effects in polycrystals while to a lesser extent in

single crystals.

86
Chapter 4 Strain ageing of low carbon steels

B BO
(G) NITROGEN

6 60
AY,
k 1 f/mm 1
4 ....-- -----------------------
,,":,_..-----·------------·-
~
«>

2
,..,-,,· 20
- ENRIETTO (30)
- - - - SEKINO I FUJI SHIMA (30
- - · - SEKINO & FUJI SHIMA (3~
10 100

8 BO
{b) CARBON

60
AV,
kcf/mm 1
6

4 ,,,,, ---·---
-------
_,,,.. ..
- · - · - - - - - - - - - - IIIN,_
.co

,.
2
•/ -
,~-STEPHENSON (28) 20
• - - - - BUTLER (;311)
0 __________ & FUJI
- - - SEKINO.._ SHIMA
__...._ (30 (3~
_ _.__~
0 2 3 4 5 . 6
o/o N °" o/oc x10,

Figure 4.16 Effects of (a) nitrogen level in solution and (b) carbon level in solution
on the change in yield stress, on prestraining and ageing for lhour at 100°C using
the following prestrains Stephenson 10%; Enrietto 12%; Seking, Fujishima, Butler
7.5%. Dissolved carbon and nitrogen were determined by internal friction (Baird
(33)).

• Other alloying additions

The elements, which directly do not themselves produce strain ageing can be divided

into four categories <33 >:

i) Elements that interact weakly or not at all with nitrogen and carbon

ii) Nitride formers

iii) Carbide formers

iv) Nitride and carbide formers

87
Chapter 4 Strain ageing of low carbon steels

i) Elements that interact weakly with nitrogen and carbon:

Manganese and Phosphorus, Copper and Nickel are elements which tend to go into

solid solution in ferrite with little or no affinity for either nitrogen or carbon and this is

reflected in their effects on strain ageing. Edwards and co-workers have studied the

effects of copper and nickel and have found that they only increase strain ageing to a

very marginal extent in a 0.025%C-Fe alloy at 250°c<33 >_ Therefore copper and nickel

will not be considered further.

• Manganese

Manganese has been known to interact weakly with nitrogen in solid solution. Effect of

manganese content on strain ageing rate is shown in Figure 4.17. Bergh et al suggested

that the variation in the ageing rates of rimmed steels was as a result of a variation in

the manganese content in steels<75 >. It has been shown that the activation energies for

strain ageing increase with increasing manganese content from 18 kcal /mol in pure

iron to 22 kcal/mol in a 1.2 % Manganese steel. This higher activation energy for strain

ageing in steels containing higher manganese is related to a decreased mobility of

nitrogen in the presence of manganese<33 >_ Internal friction studies by Fast <34 > and

Sladek <68 > show that nitrogen atoms prefer sites beside manganese atoms rather than

iron sites. Manganese has also been known to slow down the precipitation of carbon

and nitrogen during quench ageing <34• 76>.

• Phosphorus

Phosphorus has been shown to reduce the strain ageing especially at low nitrogen levels

(0.001 %) while at high nitrogen levels it apparently has no effed33 >.

88
Chapter 4 Strain ageing of low carbon steels

4.----.----.-----.-------,.-~~

.. --~::::::::::=:
-~ ~ 50

INCREASE
INYIELO
s /0/ -------· 40

MN/•&
sTAess

/
30
2 /

61-------'"'1'-,A
,,,,-·

I
{6Y) 1
ton1'-/in 2 Mn IN SOLID SOLUTIOH 1 9/e 20
/ K 0·04
./t,. 0 0•11
10
6 • 0•18
6 0•26

5 10 15 20

AGEING TIME,DAYS AT 2o•c

Figure 4.17 Effect of level of manganese in solid solution on the rate of strain
ageing of commercial rimmed steels prestrained 10% in tension (Baird<33>).

ii) Nitride formers

Since nitrogen accounts for a major portion of strain ageing that takes place below 100

°ሥ
C, this process can be reduced by the addition of nitride forming elements. However it

is essential that an earlier heat treatment is employed, which allows full precipitation of

nitrogen as alloy nitrides and carbon as carbides on final cooling <33 >_

• Aluminium

Aluminium is a strong nitride former and the compound AlN, is an extremely stable

product. Leslie at al studied the effect of prior heat treatments on strain ageing of

aluminium killed stee1s< 14>_ It was noted that in the absence of silicon, low temperature

strain ageing can be reduced by aluminium additions but not completely eliminated.

Aluminium is most effective when low austenitising temperatures are used so that AlN

89
Chapter 4 Strain ageing of low carbon steels

does not dissociate, and also in situations when subsequent cooling is slow. Aluminium

is slow to precipitate during cooling, unless the cooling rates employed are

considerably lower than that used for commercial rod< 77 , 78 ' 79>_ Additions of aluminium

therefore would not prove useful on commercial scale due to the fact that AlN

precipitates during isothermal treatment at relatively low temperature but not during the

billet reheating temperatures. Another disadvantage of using aluminium killed steels is

the poor surface finish as a result of the presence of aluminium oxide inclusions<18>.

Laxar and co-workers studied the effects of various aluminium contents on the strain

ageing of low carbon stee1< 79>_ Small additions upto 0.08% decreased strain ageing

while additions over 0.3% increased strain ageing. The effect of aluminium on the

strain ageing of low carbon steel, as studied by Epstein et a1< 29> is illustrated in Figure

4.18.

The influence of Al/N ratio on the solute nitrogen content after continuous annealing is

illustrated in Figure 4.19.

Internal friction studies have proved that the effect of aluminium was to slow down

diffusion and hence the precipitation of carbon during cooling<29>_ The observed ageing

was therefore due to a supersaturation with carbon. Aluminium can also form a

reasonably stable carbide Al4C3 but Richardson proved that the association of Al and C

in steel is made less favourable thermodynamically by the strong affinity of iron and

aluminium< 80 >.

90
Chapter 4 Strain ageing of low carbon steels

ALUMINIUM-Kil.LEO NON-AGING STEEL

IO 20 30 40 10 20 30 40
ELONGATION ON 2in, OJ,,

Figure 4.18 Stress strain curves at 20°C and 205°C for (b) rimmed steel which
showed pronounced strain ageing and (c) aluminium killed steel which showed no
strain ageing (Epstein et al cited by Baird <29>May 1963).

20
COIL TEMP.
E 0 C.R. 6SO°c 7'1Jlc
8:
.:::
0
II)
15 0
l2°c/s
500c/s

0

6
z \.I i
6~
-
UJ
§ 5
0~

~- -- i.i...

0 .. 2 3
~
......... 8
a
E---,.l...

5
.6.

6
- 7 8 9

lii,.
0

AUN

Figure 4.19 Influence of Al/N ratio on the solute nitrogen after continuous
annealing at 825°C (Meissen and Leroy<81 >).

• Silicon

Silicon combines with nitrogen m ferrite during tempering in the range of 600 -

6S0°c<29, 33 >_ It was postulated that the product SiN being similar in structure to AIN

would precipitate on AlN particles on cooling. This makes combinations of aluminium

and silicon more effective in suppressing strain ageing than aluminium alone. Silicon

91
Chapter 4 Strain ageing of low carbon steels

also forms a stable carbide but the strong affinity between Fe and Si prevents formation

of these carbides.

• Boron

Boron is an interstitial element and a strong nitride former. Boron has an atomic radius

of 0.97A0 , compared with carbon 0.77A0 and nitrogen 0.71A0 <61 >. Figure 4.20 is a

portion of the Fe-B equilibrium phase diagram. Boron is soluble in austenite to the

order of 0.001 % at 915°C and 0.02% at 1165°C, while the maximum solid solubility of

boron in ferrite is 0.002%<82>.

The diffusion coefficients of boron in austenite and ferrite are given by the following

equations< 83 >:

Austenite, D = 2*10"3 e·210001RT _ _ _ _ _ _ _ _ ,_____4_J5

. D-
Ferrite, _ 106 e -62000/RT ----------------------------------,.
"]6

The above equations suggest that boron forms an interstitial solid solution in austenite,

while in ferrite, boron forms a substitutional solid solution. This was confirmed by

McBride on the basis of relative solubilities, atom diameter and the size of the

interstitial hole in austenite and ferrite< 84>_ The diffusion rate of boron in austenite is ten

times faster than aluminium and similar to that of nitrogen as shown in Figure 4.21 <67 >.

Boron is adsorbed in austenite and tends to segregate at the austenite grain boundaries.

It also has a positive temperature coefficient meaning boron segregates in increasing

amount with increasing temperatures<67>. Boron as already stated, is a strong nitride

former. The effect of boron on the solubility of nitrogen in iron was investigated by

Fountain and Chipman<S 5) and Maitrepierre cited by Franks and Kirkcaldy<86>.


92
Chapter 4 Strain ageing of low carbon steels

ATOMIC PER CENT BORON x 10-'


0 25 50 7 100 125 150

8
1400
-=--------- ---~-=:::.=.
8+y
1300

Y+L

I
I
800o~-5~-,~o-~,s--2..._0
I
_ _2_._s_ __,_30_ __,35
WEIGHT PER CENT BORON x 10 3

Figure 4.20 Fe- B Phase diagram (M.E Nicholson<82>).

The solubility of boron nitride in austenite is given by the equation<82>

Log [B] [N] = -13970ff + 5.24---------------------------------4.J 7

where [B] and [N] are the boron and nitrogen concentrations in steel.

It was reported by Franks and Kirkcaldy <86> that in a steel containing 0.009% N and

0.007%B, at a billet reheating temperature of 1100°C, 0.004%B and 0.0052% N will

redissolve and the balance remains as BN in the austenite. The redissolved nitride

should precipitate out on cooling after rod rolling, however this process will depend

entirely on the kinetic factors like time<86' 87>. Boron has a strong affinity for nitrogen

in the temperature range of 1100°C -850 °c<61 >. In the Fe-N-B system, there are only

93
Chapter 4 Strain ageing of low carbon steels

two precipitates of boron, BN and Fe2B. The identity of the precipitate depends on the

relative amounts of boron and nitrogen in the steel.

I
I
I
I
I
I
I
I
I
I
I
- 15-0
I
I
I
I
I

- 15·5

- ll•O ,_ _.....,_ __.._ _""'---"'---~---


3 •3 3·1 2·9 2•7
1,000/7' (T tn ° R.)

Figure 4.21 Diffusion coefficient of boron (firm line) as compared to carbon


(dotted line) (Busby cs3>).

Based on the above equation, Sugimoto derived an equilibrium phase diagram as shown

in Figure 4.2i 67 >. This diagram shows that at 850°C and 4 ppm boron, the free nitrogen

is reduced to 2 ppm and in a steel containing 80 ppm boron, when cooled from 1000°C

to 850°C, the free nitrogen content decreases from 100 ppm to about 5ppm.

94
Chapter 4 Strain ageing oflow carbon steels

100 i:::- ·· · ·- - ·-

~ I-
I
IZ
11r
,-
I-
y
0.5
0
5 10 SO 100
B • [BJ (wt.ppm)

figure 4.22 Equilibrium phase diagram calculated for Fe-B-N system (Sugimoto-
faulrig<67~.

rhe solubility product ofBN in ferrite is given by<86>

log [B] [N] = -13680/f + 4.63-------------------------- 4.18

Watanabe and co-workers have calculated based on the solubility product data that in

alloys containing a total of 20 ppm boron, 0.07% soluble aluminium and 30 ppm total

nitrogen, 40% of boron (i.e 8 ppm) precipitates as BN at 1000°C<88• 89>. And in an

alloy containing a total boron of 20 ppm, soluble aluminium 0.07%, I 00 ppm

nitrogen, 60% of the total boron, (12 ppm) precipitates as BN at 1000°C. Table 4.3

shows the results of distribution of boron atoms obtained under various heat

treatments. It has been summarised that the segregation behaviour of boron atoms is

determined by

(a) body diffusion of boron atoms

(b) grain boundary diffusion of boron atoms, and

(c) the sweeping velocity of recrystallising austenite grain boundaries.

95
Chapter 4 Strain ageing of low carbon steels

Table 4.3 Distribution of boron atoms as observed by Watanabe et al <88 • 89>

Hot rolling and heat treatment Distribution of B atoms

I 300 °C xI h WQ I. Segregate to r grain boundaries.

I. Segregate to r grain boundaries.


I000°Cx I h WQ
2. Random distribution of precipitates.

I. Precipitate on prior ,- grain (at 1300 °C) boundaries. The intensity of fission
tracks of these precipitates in high Nsteel is stronger than that in low Nsteel.
! 300°Cx I h WQt930°Cx20min WQ
. 2. Segregate to newly formed r grain (at 930 °C) boundaries in low Nsteel while
in high N steel they do not segregate.

I. Segregate to newly formed r grain (at 930 cc) boundaries.


I 000°Cxl h WQt930°Cx20min WQ
2. Random distribution of precipitates.

I. Reduction of the intensity and density of fission tracks of precipitates on prior


1300°Cxlh WQtl000°Cx20min WQ grain boundaries.
2. Segregate to newly formed r grain (at I 000 cc) boundaries.

96
Chapter 4 Strain ageing of low carbon steels

Experiments by Watanabe and Ohtani on steels containing 0.0107 - 0.0326% N and

0.0025 - 0.0096% B revealed hexagonal BN precipitates as shown in Figure 4.23<S9)_

During cooling from 1300°C, boron and nitrogen atoms segregate to the austenite

grain boundaries and concentration exceeds the solubility product causing precipitation

of BN. While aluminium atoms distribute homogenously and do not affect the

precipitation process at this stage. However, when the cooling rate is very low like in

furnace cooling, Al atoms migrate to the grain boundaries and precipitate on the

surface of the BN particles in a dendritic form. When the alloy is heated at 1000°C,

AlN also is found to precipitate and grow acicularly on BN. Based on only solubility

data, BN should precipitate in preference to AlN as shown in figure 4.24, since BN is

100 times less soluble in austenite than AlN<61>. However, thermodynamically BN has

been shown to be less stable than AlN. This discrepancy between solubility and

thermodynamic data is due to the difference in diffusion rates. BN precipitates rapidly

due to the small size of the boron atom, and its interstitial location in the austenite

lattice and therefore rapid diffusion rate. In contrast aluminium diffuses faster in

ferrite than in austenite. However even in ferrite the diffusion rate of aluminium is

quite slow as compared to that of boron<61 >. Figure 4.25 shows the type of precipitate

formed under steelmaking conditions. Boron also precipitates as M23 (CB)6. However,

it has been found that BN is more stable that the M23(CB)6 especially above the Ac3

temperatures. Figure 4.26 is a schematic illustration of the precipitation behaviour of

boron in the manufacturing process.

97
Chapter 4 Strain ageing of low carbon steels

Figure 4.23 Typical spherulite morphology of BN in steel (Watanabe and


Ohtani<89>).

Steels with boron contents in the range of 0.007 to 0.02% showed a reduced ageing at

both room temperature and at 100°C after a box annealing treatmentcss, 89>. With

0.025% or more, strain ageing begins to increase, probably because of the formation of

boron oxides. Sufficient boron can be added to steel so as to prevent strain ageing and

at the same time not destroy the rimming action in steel. Boron being an interstitial

solute diffuses quickly at low temperatures and precipitates nitride under most cooling

conditions. Boron is also capable of combining with oxygen, which renders part of the

boron addition ineffective in reducing free nitrogen.

98
Chapter 4 Strain ageing of low carbon steels

..
'""":1/~~ ..,•.,..
. . -~ -.. ·~... -:·; ~f
A .· ,. ' .
:<,~ ":1.:..~· ·. ·. e~--:~.·· ·.·-

··.. .

a Al
b SN

Figure 4.24 Precipitates in a furnace cooled steel from 800°C after heating at
1300°C for one hour and hot rolling. AlN grows radially on BN (Watanabe and
Ohtani(S9>).

99
Chapter 4 Strain ageing of low carbon steels

o'tx1hr heating rate: 15°c/min

Figure 4.25 Schematic illustration of precipitation behaviour during the


manufacturing process(Watanabe and Ohtani(S9>).

600 700 BOO 900 1000 1100 c•


1110 1290 1470 1650 1830 2000 f•
PRECIPITATION TEMPERATURE

Figure 4.26 Type of precipitate formed under steelmaking conditions


(Faulring(67>).

100
Chapter 4 Strain ageing of/ow carbon steels

According to data published by Derge, a boron content of 0.02% could exist in

equilibrium in molten steel with only 0.022% oxygen<90). Unfortunately this

composition steel does not rim very well. A low carbon steel (0.06%) will rim well

when it contains an oxygen content of 0.045% free oxygen <91 ). This amount of oxygen

should exist in equilibrium with only 0.007% boron<90 l_ This means that additions of

0.007% boron would not combine with oxygen affecting the rimming action. Normally,

to fix all the nitrogen available the amount of boron required will be equal to 0.7 - 0.85

nitrogen<91 )_ A study by Franks and Kirkcaldy revealed that on reheating of billets to

around 1100°C, a significant proportion of boron nitride dissolves back into

austenite<86>.

An advantage of boron is the lack of precipitate strengthening <92>. The major reason

for this is the size and shape of the boron nitride particles formed in steel i.e spherulites

of about 2000AO in diamete/67 ). However it was noted that the ductility of boron steels

dropped with an increase in the amount of free boron remaining after nitride formation,

especially at high nitrogen leve1s<92). Free boron segregates at the austenite grain

boundaries affecting the ductility <33 )_ Further it was also shown by Nicholson et al that

boron and boron nitride have a tendency to dissolve in iron carbide<82>. This

phenomenon accounts for the fact that those steels that contained boron exhibited a

greater amount of carbide and no precipitates of iron boride.

Morgan and Shyne, as early as 1957 have evaluated the effect of boron on the strain

ageing properties of steels< 93 , 94 >_ Their experimental alloys encompassed steels of 0.06-

0.07% carbon, nitrogen contents ranging from 0.003%-0.007% and boron from zero to

101
Chapter 4 Strain ageing of/ow carbon steels

0.03%. Table 4.4 lists the details of compositions used for test. The alloys were chill

cast and then hot rolled at 982°C. Subsequently they were machined to size and then

box annealed at 705°C for 15 hours. Samples were then prestrained to 7 .5% and then

tensile tested after ageing at various temperatures as indicated in Table 4.5, which also

shows the mechanical properties obtained in these alloys after being subjected to

different ageing temperatures. It is evident that boron additions have reduced the

ageing index. The boron containing steels did not exhibit any yield point, and on this

basis the researchers proposed that steels containing between 0.007% and 0.02% boron

be termed as 'non-ageing' grades, when used in drawing operations after prolonged

storage. Boron was also found to increase ductility in these steels. It was therefore

concluded based on this set of experiments that from the viewpoint of ageing alone,

boron additions should optimally be in the range of 0.007-0.02%. However, the

distinct tendency of boron to segregate, its natural affinity for oxygen and the economy

of the process will dictate the actual additions to be made.

Studies by Yamanaka and Ohmofi<95 ) on boron containing steels revealed an increase in

the grain size with an increase in the boron content from 5 to 95 ppm and an increase in

the austenitising temperature from 900 to 1200°c<91 ). This increased grain size coupled

with the fixing of free nitrogen has resulted in a decrease in the yield and tensile

strengths and improvements in ductility. The effect of BIN ratio on elongation is

illustrated in Figure 4.27.

102
Chapter 4 Strain ageing of low carbon steels

Table 4.4 Typical steel compositions studied by Morgan and Shyne <93>_

Allo.y Pet·c Peto Pot.N ...


Otller.

C
CN
CHAI.
0.08 ·
0.0'1
0.08
0.001•
0.0088
0.0088
0.0001
o.oou
O.OMO
-
-
Al, O.ON
CNV 0.08
0.07
0.0013 0.0038
. 0.0057
v. 0.039
CNBi • 0.0008 B. 0.0:,0
CNBt O.O'l 0.0105 o.oou B. o.oao
CNTi 0.06 0.008• · 0.006• Ti; 0.'8
CB 0.06 0.0026 0.0001 .B. o.ooa

• Assumed values; not actually analyzecl.

103
Chapter 4 Strain ageing of low carbon steels

Table 4.5 Mechanical properties of the steels listed in Table 4.4 (Morgan and Shyne<94>) •

Allor ll-rop•r&l,- .AAQeal•• a&SPO


"a'.Da79· aODa7•
af.aa•o
......
---··..-o_
..
......
100-C
~

C %30 8,220 8320 . 15"1'SO &&SO


CN 200 8100 9230 8840 8800
CNA1 A.gin&
230 1090 1'720 1310 11530
CNV
CNBi
CNBs
CNTt
J.ndex,
ps:I.
0
180
160
0 580
1500
1490
490
200
2930
2950
500
-
480
2530
340
-60
1750
240
CB
C. 36930
o· 15100
38050
6580
39370
-
39470
-
39400
CN 3'7760 44290 4.4540. 43740 43100
CNAl 38360 38780 .. · 3'1800 38480 38000
Ui-tun.a-t;e 40290 39860 40800- 40140 40030
CNV
CNB1
CNs.
a-t.re:n~.
psi 37400
37270
38810
36710
3'1-400
38840 3~0
-
38830
•1seo
CNTl·
. ea
C
CN
4::LBOO

52.1 - .
.42080

{ """:_..'. --:=.o ,:ta;•


420G().
30200
'8.4
38.8
-
41700
48.8. ·
40.8 .·
48.1
43:. .
CNA.I. ,.53. 7 .. &52.8 53-0 . 54..2 53.!'"
CKV TotaJ.- 51.2 .48.'1. 50.4 ·s1.o 151;
Cl'f"B.1
CND.,;
eloncation.
pct~:J..tn.
6e.O
55.0
155.IJ
~-3
G5.1
55.6
-52.9 -56.7
~Ti. 52.2. 5 .2 ~2.3. 152.1 52.'1
CB 55.2 50.8 4,8.6

104
Chapter 4 Strain Ageing of low carbon steels

0 1 2 3 4
BIN (WT. % RATIO)

Figure 4.27 Influence of BIN ratio on the elongation (Faulring<67>).

iii) Carbide formers

Molybdenum is not a strong nitride former, however it does form carbides in steel.

Molybdenum contents if the range of 0.3 - 1.98% have had no effect on strain ageing at

250 °C <33 >. Further studies have indicated that elements that form only carbides but not

nitrides will not reduce either low temperature or high temperature ageing.

iv) Nitride and Carbide formers

Chromium, vanadium, niobium, titanium and zirconium are elements which can form

nitrides, carbides or carbonitrides. Thermodynamically nitride formation is more

favourable. Where a separate nitride forms, all nitrogen present has to be fixed before

any carbon is removed from the solution. But elements like vanadium, niobium and

titanium, which result in the formation of a carbonitride may contain carbon at

equilibrium.

105
Chapter 4 Strain Ageing of low carbon steels

• Chromium

Chromium behaves in a similar manner to manganese in that it is the weakest nitride

and carbide former amongst the above stated elements. Edwards and co-workers

studied the effect of various chromium contents on the strain ageing characteristics<96).

It was found that low carbon steels containing 6.10% Chromium do not show any strain

ageing, indicative of fixation of all nitrogen and carbon by chromium. Several others

have investigated effect of upto 0.44 % Chromium in low carbon stee1s<97>_ Increase in

Cr contents apparently decreased strain ageing at 100 °C, and at 0.44% indicate that Cr

fixes nitrogen in solution with a resulting decrease in the migration of nitrogen to

dislocations<97>.

• Vanadium

Vanadium is a stronger nitride and carbide former than chromium, however its

reactions are kinetically retarded resulting in the formation of non-stoichiometric

products<33 >_ As a result there is removal of only one-fourth of available nitrogen from

the matrix. It has been shown that 0.26% vanadium decreases strain ageing at 250 °C

while 0.69 % completely eliminated it<97 - 101 >. Considerably less vanadium (0.02 -

0.03%) is required to prevent low temperature strain ageing. A V/N ratio of 7 to 9 has

been found essential to combine all free nitrogen to form VN<97 - 101 >. Vanadium in

excess of this combines with carbon to form vanadium carbide resulting in undesirable

precipitation hardening effects. Another advantage of using vanadium as a nitrogen

fixer is that it is not a strong deoxidiser, therefore the steel need not be fully killed

before nitrogen can be fixed by vanadium<9&J_ Experiments by Epstein et al have

shown that non-ageing rimming steels can be made by adding vanadium<97 >_ The

amount required depends on the nitrogen content. In steels with high nitrogen content,

106
Chapter 4 Strain Ageing of low carbon steels

it could affect the rimming action seriously. However in lower nitrogen containing

steels the amount of vanadium needed would be 0.03%. Vanadium nitride that forms

takes away available nitrogen from the matrix preventing strain ageing. This nitride is

unstable at temperatures above 700° C and steels annealed above these temperatures

would be susceptible to strain ageing again. Figure 4.28 shows the strain ageing

behaviour of a 0.21% carbon steel containing 0.006% nitrogen, (a) and the same grade

also containing 0.06% vanadium (b). It is evident that strain ageing is almost

eliminated(IO))_ Beyond 0.06% vanadium, there is a possibility of some precipitation

hardening effects. The degree of extra strength provided by a given amount of

vanadium becomes greater as the carbon content is increased. In order to avoid

unacceptably high levels of strength it is necessary to control the composition to

comply with the limit( 33 >(also shown in Figure 4.28)

[V] *[CJ< 0.0026-------------------4.19

The effectiveness of vanadium in stabilising the nitrogen appears to be independent of

the aluminium content( 101 >. There also appears to be only a minimal refinement in the

grain size with vanadium additions of upto 0.1 %. It is therefore logical to conclude

that vanadium additions are less effective that titanium in eliminating strain ageing

together with achieving optimum mechanical properties. However one advantage is

that vanadium can be added to semikilled steels or steel grades to be continuously cast

while titanium has to be added only to killed grades.

107
Chapter 4 Strain Ageing of/ow carbon steels

500 lb)
(al

400
.........
Ii
:z
:It 300
.,;
"'
!!!
ij;
200

100
pre-strain

0 0 10 20 30
Strain, •;.
fa)

SOO

400

""e 300
%
z
::r~ 200
in (al lbl

100
pre-strain

0 L....1'-------"1------,10~------,2:!:o:----='30::-
straain • •.,.
(b)

Figure 4.28 Stress strain curves for (a) low carbon steel 0.005% nitrogen and (b)
addition of 0.06% vanadium (Erasmus et al <101 >).

• Niobium

Niobium appears to form more stable carbides and nitrides than vanadium<33 >_ 0.28%

niobium reduces strain ageing at 250 °C while 0.36% completely eliminated both the

yield point and strain ageing<97 )_ However, the extent of strain ageing depends largely

on the balance between carbide and nitride formation, which is particularly of

importance in niobium bearing steels. It has also been observed by Mandry that the

strain ageing index of niobium containing steels is sensitive to austenitising time and

temperature, first rising to a maximum as the temperature is increased and falling as the

carbonitride is taken into solution and precipitated out on cooling as a nitride rich

108
Chapter 4 Strain Ageing of low carbon steels

carbonitride<33 >. The general trend is that the lower the temperature at which

carbonitride is formed, the higher is the proportion of nitride it contains.

• Titanium

Titanium is one of the strongest carbide and nitride forming elementscso, 102>. Because

of the greater stability of the nitride, there is a tendency for the nitride to first

precipitate from steel <102, 103 ' 104>. This precipitation takes place at temperatures at

which the elements can diffuse quite rapidly and therefore stabilising effect of titanium

is not dependent on the prior heat treatment. At high temperatures, titanium reacts

easily with oxygen, sulphur, nitrogen and carbon to form their compoundsoos, 106>. The

precipitation temperature ranges for these compounds are shown in Figure 4.29. The

temperatures of oxide, sulphide and nitride formation are relatively high and

accordingly these compounds precipitate in the casting and billet making stage.

However, the precipitation temperature of the carbide is low and depends strongly on

the rod rolling/cooling conditionsC 105 ' 106>. Although titanium is one of the strongest

carbide formers, the optimum amount of titanium that can completely fix the solute

carbon in ferrite is yet to be determined. The amount of titanium that is available to

combine with carbon to form TiC is given by the following equation< 105 , 106>.

[Ti available] = [Ti] - (2*48/32)[S] - 48/14 [N]--------------4.20

where [Ti], [S] and [N] are the weight percent of each element in the steel. The ratio

of the stoichiometric amount of titanium required to fix carbon as carbide to the amount

of titanium available is given by

Ti ratio= [Ti available]/ (48/12 *[C])---------------4.21

where [C] denotes the carbon content of the stee1< 105 , 106>.

109
Chapter 4 Strain Ageing of low carbon steels

2000

1800 mp Oxides ITixOyJ


Nitride {TiNJ

!
1600
____s_ul-fid_e_s_(T-ix_s_v1_ _ _ _ r
;:!
I! 1400
I

~
Carbide (TiCJ in austenite
1200 A2 ......,_ _ _ _ _ __,,......,.,.,.,.,.,.,...,..

1000

800
Ea~ili 1~:~~i
Figure 4.29 Precipitation temperatures of titanium compounds (Ochiai<105~.

Studies on two batches of steels, one with a carbon content of 0.005% and the other

with a carbon content of 0.02% with additions of titanium in the range of 0.03 to 0.19

% revealed that the mechanical properties of the rods varied as shown in Figure 4.30.

Figure 4.31 shows the influence of titanium additions on the ageing index. The ageing

index was found to decrease with an increase in the titanium ratio. However the

decreasing tendency was found to be strongly dependent on the carbon content of the

rod. The low carbon grades showed a gradual decrease in the ageing index with

increasing Ti ratios and even at a very high ratio of 2.91, the steels exhibited an ageing

index of 8 Mpa. While the high carbon grades showed an abrupt decrease in the ageing

index when the Ti ratio exceeds 0.96 and at a ratio of 2.07, the index falls to a value of

2 Mpa. Titanium carbide was found to precipitate below the A3 temperature in the low

carbon grade steels while the precipitation commenced in the austenite stage itself for

the high carbon grade steels. These precipitates form nuclei for further precipitation in

the ferrite phase.

110
Chapter 4 Strain Ageing of low carbon steels

400,----------~

TS

150
1001-------------J

C (mass%)
0 0.0050--0.0057
ii. 0.017 -0.022

800:::----+,----,!2,---.....J
Ti-ratio 3

Figure 4.30 Mechanical properties of rods containing titanium (Ochiai< 105>).

C lmass 'lit/
O 0.0050-0.0057
e. 0.017 --0.022
)(
.g 40
-=Q
C
"cii20
c(

1 2 3
Ti-ratio

Figure 4.31 Effect of titanium additions on the ageing index of the wire
(Ochiai< 105>).

It has been stated by Ochiai and co-workers that to make a 0.005% carbon steel non-

ageing, the titanium ratio should be increased to a value over three< 105, 106>. However in

order to obtain a non-ageing steel with a lower Ti ratio, the cooling rate should be

decreased further. This becomes industrially impossible in low carbon grades where

the cooling rate has to be very low to achieve the best possible decrease in the ageing

index. It is therefore necessary to increase the carbon and titanium contents so as to

introduce precipitation nuclei during rolling in the austenite region or to accelerate


111
Chapter 4 Strain Ageing of low carbon steels

strain induced precipitation by rolling at a temperature very near the nose of the

isothermal precipitation curve for TiC.

In a steel with carbon content of 0.025%, 0.21 % Ti is enough to eliminate both the

initial yield point and strain ageing at 250°C, while the same level of titanium

associated with higher carbon i.e 0.085% gave rise to strain ageing. Increasing the

titanium content to beyond 1% did not eliminate strain ageing. Titanium or any other

alloying element must be added in such a way that precipitation hardening is avoided
(33)

The maximum titanium content recommended in order to avoid precipitation hardening

is given by the following equation <33 , 105>

% Ti - 3.4% N (total)< 0.07---------4,22

The effect of excess titanium on tensile strength is shown in figure 4.32.

Titanium has an additional advantage of reducing the soluble carbon to extremely low

levels and is used to produce "Interstitial Free" steels. Titanium, however is a very

strong deoxidiser, and therefore the steel has to be fully killed before it can be ensured

that nitrogen can be fixed by titanium< 102, 105 - 106>. Smail, Keown and Erasmus have

confirmed that additions of 0.02-0.03% titanium to a 0.005-0.006% nitrogen steel have

completely eliminated strain ageing< 102>.

112
Chapter 4 Strain Ageing of/ow carbon steels

470

N
'e
420 I
X

I
~ 370

320-._ __..._ _ __.__ _ _........__ _ _.,_,


0 0-025 0·050 0·075 0-01
•1. Ti
-3·4•/.Nr

Figure 4.32 Effect of excess titanium on the tensile strength of low carbon steel
(Erasmus 102>).

However it is clear that the effectiveness in preventing strain ageing increases with an

affinity for carbon and nitrogen in an increasing order of manganese, chromium,

vanadium, niobium and titanium <33 >_ A larger amount of any of these carbide /nitride

formers is required for eliminating strain ageing at 250°C than that needed to offset the

process below 100°C. But because most of these elements combine with both carbon

and nitrogen, a larger amount is required to eliminate lo" temperature ageing than that

needed for a strong nitride former alone.

The exact mechanisms of the effects of carbide and nitride formers is not well known in

many cases. In some studies of internal friction have revealed interactions between

nitrogen and molybdenum, vanadium, chromium and manganese in solid solution,

while this is not true with carbon.

113
Chapter 4 Strain Ageing of low carbon steels

4.2.5.2 Effect of Prestrain

If the directions of straining are different before and after ageing, or when the

prestraining is by rolling, the discontinuous yield behaviour returns slowly<30). The

presence of a drop in the yield strength and subsequent yielding by Luders band

propagation begins at a low density of free / mobile dislocations which are at a

sufficient stress to move dislocations over large distances. Quick multiplication of

these mobile dislocations, which have formed in the prestrain stage, takes place. As a

result, dislocations will have to move at a lower velocity to give the applied strain rate.

Since the dislocation velocity is stress sensitive, the yield strength drops on initial

yielding which then propagates at the lower yield stress by the movement of Luder

band fronts< 30). New dislocations are created at these bands by the stress concentration

existing there.

Stretching in tension to a strain less than the lower yield point elongation divides the

specimen into regions yielded and unyielded separated by Luders bands formed at

approximately the lower yield stress. Until all unyielded regions have been consumed

deformation will continue by the propagation of Luders bands.

A reduction of less than 3% rolling also divides the specimen into yielded and

unyielded zones. But as observed by Hundy (4), Butler 173 ) and Wilson< 107• IOS) , the

zones alternate over short distances so that the availabk strain is widely distributed

throughout the specimen. This type of deformation gives rise to alternate bands of

deformed and undeformed material meeting the surface at perpendicular angles to the

114
Chapter 4 Strain Ageing of low carbon steels

rolling direction and running into the centre of planes of maximum shear stress. Butler

and Wilson have shown that after ageing, yielding begins at boundaries between the

yielded and unyielded bands. This suggests that yielding may be initiated with the help

of microstresses existing at these boundaries. After light strain ageing, yielding begins

at several points within the specimen but on prolonged ageing it gets confined to the

Luders bands. Butler and Wilson were able to show that the directionality of the yield

point return was as a result of directionality of the deformation introduced by rolling.

At greater than 3% temper rolling reductions, this directionality is reversed. This

suggests that for initial suppression of the yield point, the available strain should be

dispersed as evenly as possible throughout the specimen, such that no region is far from

a source of fresh dislocations(3o, 109).

Studies by Wilson and Ogram (l IO) on tubular specimens tested in torsion, aged and

restrained in same or reverse directions revealed that the return of the yield point was

not due to residual microstresses but because of strain ageing. It was observed that in a

specimen strained in tension and restrained in the same direction, most of the mobile

dislocations will not be able to move far since they will be pressed against obstacles

like grain boundaries, precipitates etc. So only a few isolated dislocations need to be

locked up to produce a return of the sharp yield point. While if the restraining is done

in any other direction other than the prestrain, the stress \\ill act to move dislocations

away from the obstacles. Only isolated dislocations may be locked up before those

stopped by obstacles. This is especially true at low solute contents. This explains the

different rates of the yield point return.

115
Chapter 4 Strain Ageing of low carbon steels

Observations by others have revealed that the rate of increase of the upper yield point in

specimens strained in the same direction before and after ageing is greater, the higher

the forward stress that is maintained during ageing. This stress keeps dislocations in

regions of unfavourable internal stress field and helps delay unlocking and subsequent

restraining. If a compression prestrain is applied on then material then it was found that

the yield point returned slowly as compared to when the prestrain was applied in

tension() 11 >. This was made use of commercially by using rolling as a method of

compression prestrain that would retard the return of the lower yield point <11 n_ Similar

results are achieved when a tensile prestrain is employed in a transverse direction. The

difference on behaviour with a change in the prestrain modes was said to have been

caused by residual microstresses.

The change in properties like ultimate tensile strength, elongation and hardness are not

affected by the mode of prestrain, while a higher amount of prestrain does not result in

a greater lower yield point elongation and a greater increase in the ultimate tensile

strength and an increase in the rate of work hardening<30l.

It has also been noted that while an increase in the amount of rolling reduction does

tend to delay the return of the yield point, the ductility of the material is severely

reduced after rolling and ageing.

Hundy et al have also studied that the cumulative effect of straining in increments and

ageing between stages of straining is much more than when the strain is applied in a

single step<6l. Experiments by Nakada et al revealed that as long as the ageing period
116
Chapter 4 Strain Ageing of low carbon steels

did not extend beyond the first stage of ageing, the ageing behaviour was insensitive to

the amount of prestrain.

4.2.5.3 Effect of Temperature

There is considerable evidence that in steels containing nitride formers the ageing

below 100 °C, which is believed to be as a result of nitrogen is negligible<29>_ Above

100 °C, solubility of carbon is sufficient enough to cause ageing. The predominant

effect of increasing the ageing temperature is to accelerate the ageing process. The

activation energies for strain ageing at a particular temperature are found to be

comparable to those needed for the diffusion of these interstitials at that temperature<29>

4.2.5.4 Effect of Prior Heat Treatment

It is obvious that effects of various heat treatments would have to be correlated with

various compositions of the steels <30 >. However a generalisation can be made on the

basis that any heat treatment that results in an increase in the carbon and nitrogen in the

matrix also increases strain ageing susceptibility.

It was earlier believed that removal of free carbon, nitrogen atoms in the matrix by

precipitation occurs during slow cooling. However, work done by Cottrell <112>, Morgan

<113> and co-workers have indicated the precipitates formed during slow cooling rates

would be too widely spaced for carbon and nitrogen atoms to be removed by diffusion.

It was suggested that quench ageing would be a better alternative to fix carbon and

nitrogen. This process gives rise to a higher density of nuclei at temperatures of 100 -
117
Chapter 4 Strain Ageing of low carbon steels

200°C and thus decreases the amount of carbon in solution(30 >. At temperatures below

these, the precipitates that form would be too small to be physically stable. These can

easily redissolve on restraining.

Further, it was also proved by Butler that the carbide distribution that forms during

cooling through the transition range affects the degree of carbon strain ageing<73 >_

Faster cooling rates in quenching or normalising would produce a ferritic structure

supersaturated with carbon. Heating to the annealing tern perature would produce a

fine, uniform dispersion of spheroidal carbides, while lower cooling rates like annealing

would result in the formation of pearlite colonies or grain boundary carbide films. The

grain size and hence the grain boundary carbide separation increases with decreasing

cooling rate. Slow cooling from the annealing temperature would allow carbon to

diffuse to existing carbide particles. As the temperature decreases further diffusion of

carbon becomes sluggish and carbon atom movement is limited to short distances. In

this respect steels containing a fine and closely distributed carbides would prove more

effective in removing carbon from the solution than steels containing a coarse

distribution of carbides.

With a steel containing nitride or carbide formers, an optimum heat treatment is one

that employs temperatures sufficiently low to give a low equilibrium solubility of

carbon and nitrogen and sufficiently high to allow precipitation of these elements. The

effectiveness of a given alloying element in inhibiting strain ageing will depend on how

closely this condition is approached.

118
Chapter 4 Strain Ageing of low carbon steels

As already mentioned earlier, strain ageing can be of two types : static strain ageing and

dynamic strain ageing. The next section contains an account of the process of dynamic

strain ageing and the factors that affect this process.

4.2.6 Dynamic Strain Ageing

This denotes the ageing processes, which occur simultaneously with plastic strain <3549>_

As a result, the effective strain rate is the most important parameter governing the

extent of ageing in steel. The ageing process can occur at room temperature in slow

tensile deformation (10 "6 sec · 1) and may also be effective in raising the room

temperature fatigue strength in many cases <35 -49 >_ At normal tensile deformation rates

i.e 10 4 sec · 1, dynamic strain ageing occurs in the range of 100 - 300 ° C <37>_ If the

interstitial solute levels are very high then effects of the ageing process may extend to

room temperature. At high strain rates - 10 1 to 300 sec -I (impact deformation), the

effects occur at 400 - 670 ° C. In the temperature range or 125 - 300 ° C, most low
carbon steel grades show a pronounced increase in work hardening rate and a decrease

in the ductility. The stress - strain curve changes from a smooth one to a serrated

shape. This blue brittleness phenomenon has been attributed to nitrogen and carbon.

Experiments by Baird and J~eson(I 10> have shown that in steels free of nitrogen and

carbon, the tensile properties exhibited a smooth behaviour. While in steels containing

carbon and nitrogen, serrated yielding was found in the range 175 - 225 ° C and jerky

flow in the range of 125 - 250 ° C. The appearance of serrated/ jerky flow may occur

during tensile, compressive or torsional loading conditions and is characteristic of

deformation during a specific regime of temperature, strain and strain rate<9&-I09>_ This

119
Chapter 4 Strain Ageing of low carbon steels

flow was first observed by Le - Chatelier in mild steels and was further investigated in

Aluminium alloy by both Portevin and Le Chatelier. It is also called the Portevin- Le

Chatelier or the PLC effect. These curves depict dynamic strain ageing which occur

when the rate of straining is such that the interstitials can diffuse and pin the mobile

dislocations and serrations occur because of a rapid generation of new dislocations

which are needed to sustain flow< 104 >_ In this region the probability of freeing

immobilised dislocations is very low. In the process of generating new dislocations the

stress increases.

While serrated flow curves are the most commonly observed macroscopic

manifestations of dynamic strain ageing, there are other phenomena. Bands of localised

deformation, known as PLC bands are also observed on the surface of a deforming

specimen producing a serrated curve< 100- 109 l_ The occurrence of intensified acoustic

emissions concurrent with the process has also been noted. However two important

features are associated with the process i.e. the peak in the flow stress variation with

temperature and the negative strain rate sensitivity of the flow stress< 111 >. The peak in

the flow stress is generally observed to occur at temperatures of - 0.3 Tm· Figure 4.33

shows the general cr - E curve for a pure metal and corresponding curves for an alloy

exhibiting dynamic strain ageing as a function of temperature. Strain localisation may

occur spontaneously as a result of the negative strain rate sensitivity arising as a result

of the process.

120
Chapter 4 Strain Ageing of/ow carbon steels

~-...,,,,,,--
,,,, I
\
i

1'. : ,............~
,••,,,- l
..
/ '

.,
<C
~
2
2
/ I
,// I I
I NO JERKS

I
0 I
z I

,so" 200°

Figure 4.33 A normal stress - strain curve (20°C) and corresponding curves
illustrating dynamic strain ageing as a function of temperature.

The solute responsible for dynamic strain ageing is considered to diffuse sufficiently

fast to slow down freely moving dislocations by forming an atmosphere around

them<93>, The solute atoms exert maximum drag on the dislocations when they move at

some critical speed v c, given by

vc - 4D I h -------------------------------------------4.23

where Dis the diffusion coefficient of the solute, his the effective radius of the solute

atmosphere.

The critical strain rate, Ee is given by

Ee = b p V = {4 p b/h} D------------ ---------4.24

where p is the mobile dislocation density.

D -0.12 Cv exp (-EIRT)-----------------------------------4.25

where E is the effective activation energy for the exchange of vacancies and

Cv is the vacancy concentration which increases with the amo unt of strain.

Cv - B & 111 -------------------------------4.26

121
Chapter 4 Strain Ageing of/ow carbon steels

Therefore
&c = 10 B exp (-E /RT) &cm _________________________________ 4.27

where Ee is the strain rate to the first serration. Further it was proposed by Ham and

Jaffrey that the variation in the dislocation density be taken into account<9 S) •

p- N !3 ----------------------------------------4.28
and therefore

- /. ,1 (m+/i) ,I 29
& - 1consta11t1 &c -----------------------------,.

Figure 4.34 shows the force-velocity diagram for a mobile dislocation. The motion of

dislocations through a metal in the absence of a solute atmosphere is retarded by lattice

friction< 9S)_ The retarding force increases as the velocity of dislocations increases (line

1)° 7). When the velocity of the dislocations is compatible with the diffusion rate of the

solute atoms, a cloud of these atoms will be attracted to the strain field of the

dislocations. As dislocations move, they effectively tow a cloud of solute atoms

through the lattice. The force required to drag the solute atoms increases rapidly with

velocity (Curve 2). As the velocity increases, the concentration of the solute atoms

around the dislocations is reduced, (curve 3). The solute concentration drop has the

effect of reducing the drag force.

Provided that the velocity of a given dislocation does not exceed a critical value Vm ,

the solute atmosphere will remain in the stress field of the dislocation as it moves

through the lattice. At dislocation velocities beyond V111 , the atmosphere can no longer

keep with the moving dislocation. This threshold velocity is a function of temperature,

species of solute and the bulk solute concernration. When a dislocation is suddenly

relieved of the drag force created by the solute atmosphere, it will accelerate up to the

122
Chapter 4 Strain Ageing of/ow carbon steels

next stable velocity V3• At this velocity, the dislocation encounters obstacles that will

slow its speed i.e precipitates, grain boundaries, dislocation substructures caused by

prior cold work. The dislocation is slowed to a point that the solute atmosphere can

once again keep up with the dislocation. The cloud of solute atoms once more

surrounds the dislocation and the cycle continuesC 17 >_

Force
,+-- (2) Solute drag force
I

1/
\ I (3) Concentration of
dift"using solut.e atoms

I ~
I \
I
I ~'~--~----------:i-r
:•

I
I
I • ··~
::
:
··~ '
'- •••
I
I

v. Velocity
SOLUTE i INSTABILITY1 LATTICE ....
DRAG i l FRICTION

Figure 4.34 Force-velocity diagram for a mobile dislocation (McCallum 17).

4.2.6.1 Effect of Carbon and nitrogen

Nitrogen produces a greater effect than carbon as in tl1-: case of static strain ageing,

due to its higher solubilityc 37 , 98 ) . A number of work1:L, hc1, e shown that as nitrogen

levels increase, the blue brittleness strengthening effects increase. Increasing the

123
Chapter 4 Strain. lgeing of low carbon steels

nitrogen levels raises the rate up to 10% strain and also extends the range of rapid

strain hardening induced by dynamic strain ageing. This occurs by increase in the

strain at which the supply of nitrogen to pin freshly generated dislocations becomes

exhausted. Local exhaustion manifests itself as coarse, localised bands of

deformation in which nitrogen is exhausted( 37 , 112 ) . If the material in these bands

cannot work harden sufficiently to compensate for the reduction in area of the test

piece in the band, the specimen breaks ,vithin the band with a low elongation.

Increasing the nitrogen level increases the strain at which 1his occurs, thus increasing

the elongation to failure< 37 >_ However at sufficienlly high nitrogen levels the

elongation to fracture returns to the same value as when the steel was free of nitrogen

i.e embrittlement effects are zero<37).

4.2.6.2 Effect of substitutional alloying elements

The effect of substitutional alloying elements in solid snlution on static strain ageing

normally is to slow down the rate of ageing. In the case uf dynamic strain ageing,

elements like manganese, chromium, molybdenum, tungsten and copper give rise to

strain hardening during tensile deformation of steel at temperature considerably above

the normal blue brittleness range< 93>_ Several of these elements also tended to decrease

the intensity of carbon- nitrogen effect at around 200 ° C. Presence of substitutional

solutes with an affinity for carbon or nitrogen extends tb\..· blue brittleness peaks to

high temperatures <37 , 98 >. Chromium produces a similar effect to manganese and

molybdenum extends this blue brittleness to higher temperatures.

-------------------- - - - ------------
124
Chapter -I- Strain Ageing of/ow carbon steels

Similar to the effects on static strain ageing, substitutionai alloying elements like nickel

which have no affinity for carbon and nitrogen appear to havl' little or no effect on the

dynamic strain ageing. The effects of carbide and nitride formers appear to be

qualitatively similar to those obtained in static strain ageini 981 .

4.2.7 Strain dependence of Static and Dynamic ageing

Both, static strain ageing and dynamic strain ageing are strain dependent. The physical

basis of dynamic ageing is an increase in the effective dislocation glide resistance due
'l' of so lute atoms above a certam
tothe mo b11ty . temperatur~ (I0"-104)
- . Stat1c
. stram
.agemg
.

occurs as a result of the segregation of solute atoms at dislocations at room

temperatures or slightly above room temperature. There have been several models that

predict the strain dependence of static and dynamic strain ageing on strain. In earlier

models it was assumed that mobile solutes vverc interacting with moving

dislocations<93)_ However, it was later realised that the dislocation movement was

always jerky and the effect of solute mobility on the glide resistance will not take place

during the actual movement of the dislocations but when the dislocations are waiting at

obstacles. It was postulated that the strain rate sensitivit) index of the flow stress

affected the behaviour of the materia1° 02 - 1041 . It was also assumed that any solute

mobility that affects the total strain rate sensiti\·ity negati\ ely and that this contribution

increases with strain. When the total strain rate scnsiti\ 1ty becomes negative, then

plastic flow becomes jerky or unstable. This is supported by the experimental

observations that the point of time when the llmv becornl's _jerky after a certain amount

125
Chapter 4 Strom Ageing of low carbon steels

of critical prestrain, occurs at approximately at the same point at which the strain rate

sensitivity becomes negative<9 S)_

Different models have however explained the dependence of the rate sensitivity

differently. Van den Beukel believed that the diffusion coefficient of the solute is strain

dependent which follows from the fact that there is a known increase in the vacancy

concentration with deformation< 106' 107 )_ This applies to substitutional alloys only. The

mobile dislocation density was also found to depend on the strain as indicated from the

relation between the strain rate and the time. This applies tu all the alloys. Another

model by Mulford and Kocks proposed that the strain rate sensitivity in normal alloys is

better described in terms of the flow stress than the strai1/! 8). It was also stated that this

model was only applicable to the dynamic strain ageing process. They concluded that

the solute mobility affects the strain hardening component of the flow stress rather than

the friction stress.

Both models are incomplete in that they do not explain strain (kpendent diffusion or the

dependence of mobile dislocation density on strain.

This comprehensive literature review as presented in Chapters 2, 3 and 4 is followed by

Chapters 5, 6, 7, 8 and 9 which contain information pcrt:1ini11g to the current research

study. The next chapter outlines the main objectives of the current research.

126
Chapter 5 Research Objectives

CHAPTERS

RESEARCH OBJECTIVES

127
Chapter 5 Research Objectives

5. RESEARCH OBJECTIVE

BHP Long Product Division at Newcastle have been supplying low carbon steel rod

grades for wire drawing applications. Several instances of difficulties in drawing these

rods to wires were encountered and detailed investigations of these cases pinpointed the

cause to strain ageing. It was therefore felt necessary that a detailed study of this

phenomenon of strain ageing be carried out in the low carbon steel grades under various

conditions of testing.

The primary aim of the project was to evaluate the strain ageing behaviour of low

carbon steel wire grades used for structural applications and produced at BHP,

Newcastle. The specific objectives were:

• To evaluate the strain ageing behaviour through the use of hot torsion tests

under specific parameters:

Initially hot tension tests were attempted for this purpose. However these

proved less suitable to simulate industrial drawing conditions when

compared to hot torsion, as explained in section 7.5. Hot torsion tests have

been used as the closest possible simulation of the hot rolling/drawing

processes, which allow the material to be formed under a triaxial state of

stress. Other researchers< 27 > have used the hot torsion test to study behaviour

of materials under rolling/drawing conditions.

• Study the strain ageing behaviour as a function of steel nitrogen content: As

indicated extensively in literature, nitrogen is found to be the main cause for

strain ageing. It was therefore considered appropriate to study the impact of


128
Chapter 5 Research Objectives

vanous nitrogen contents on the strain ageing behaviour of low carbon

steels.

• Study the impact of boron addition on the strain ageing behaviour of the

above grades of steel:

Literature has described the various remedial methods that have been used

to alleviate strain ageing. One of the more common and successful ways

have been to fix free nitrogen and carbon in the steel with the deliberate

addition of strong nitride/carbide formers to the steel. As already

summarised in the literature in section 4.1.4.1, initial studies on boron have

revealed an appreciable effect of boron on strain ageing. However, as

mentioned earlier, the exact amount of boron that needs to be added and the

nature and kinetics of precipitation of the boron nitride needs to be studied.

• Study the effect of drawing temperature and time held in between drawing

stages on the strain ageing behaviour:

Temperature and time have been proven to have a significant effect on the

strain ageing phenomenon. During actual manufacturing processes, often,

there are delays between stages, as a result of which the material may be

held soaking at a particular temperature for a certain period of time. This

could have deleterious effect on the mechanical properties in terms of strain

ageing. It was therefore considered necessary to study the effect of these

parameters on various steel compositions.

• Correlate the above hot torsion results with steel microstructures using the

transmission electron microscope:

129
Chapter 5 Research Objectives

As mentioned earlier, one of the more popular methods of combating strain

ageing is to fix the free nitrogen/carbon with the addition of suitable alloying

elements to form nitrides or carbides, which would be distributed in the steel

microstructure. Therefore characterisation of the steel microstructures was

of paramount importance as this would yield useful information that could

explain the trend in mechanical properties.

The next Chapter describes the experimental procedure that was adopted in the current

research project.

130
Chapter 6 Experimental Methods

CHAPTER6

EXPERIMENTAL METHODS

131
Chapter 6 Experimental Methods

EXPERIMENTAL METHODS

6.1 STEEL COMPOSITION AND PROCESSING:

Four low carbon steel grades used for structural applications were chosen for study.

These steels were produced at the BHP steel Newcastle works. Table 6.1 lists the

chemical compositions as determined by conventional wet analysis and atomic

spectrometer methods, of these four steels. Two of these steels were production steels

produced by the conventional "Basic Oxygen Steelmaking" (BOS) process, where

oxygen is blown through the steel. One of the BOS steels had a low nitrogen content

(0.002 wt%), while the other contained more nitrogen, 0.008 wt%. The third grade, an

experimental steel chosen for the current study was produced by the electric arc (EAF)

process and had 0.008 wt% nitrogen with deliberate additions of 0.001 % boron. The

other steel was also an experimental steel, prepared by the Basic Oxygen process and

had 0.006 wt% nitrogen and 0.004% boron, added with an intention of scavenging the

nitrogen.

6.2 ROOM TEMPERATURE MECHANICAL PROPERTIES:

The steels were tension tested for their mechanical properties at room temperature using

an Instron machine (model 1185 manufactured in U.K) using a conventional strain rate

of 10 -I /sec to Australian Standard AS 1391 using a 5mm diameter test piece.

6.3 MICROSTRUCTURAL CHARACTERISATION:

6.3.1 Optical microscope:

Samples cut from as-rolled rods of the four steel grades were metallographically

polished using different grades of silicon carbide papers and diamond paste polishing

cloth. These were then etched with a solution of 2% Nital. The microstructures were

132
Chapter 6 Experimental Methods

examined by bright field optical microscopy with the objective of characterising the

different phases present on an Olympus PMG3 microscope.

Table 6.1 Chemical compositions of the four grades of plain carbon steels used for
test.

MATERIAL C Mn Si Ti V Cr Al Mo N B
BOSLowN 0.07 0.32 0.158 0.001 0.001 0.03 0.004 0.01 0.002 0

BOS HighN 0.08 0.33 0.17 0.001 0.002 0.02 0.003 0.01 0.008 0

EAF 0.05 0.49 0.144 0.001 0.003 0.05 0.003 0.01 0.008 0.0012
(LOWB)
BOS (N+B) 0.08 0.36 0.174 0.02 0.006 0.01 0.006 0.0038
(HIGH B)

MATERIAL TOTAL SOLUTE TOTAL INTERSTITIAL


CONTENT CONTENT
BOS LOWN 0.596 0.072
BOSHIGHN 0.624 0.088
EAFLOWB 0.7602 0.0592
BOSHIGHB 0.6598 0.0918

6.3.2 Transmission electron microscope characterisation (TEM)

Sections of as-rolled rod samples of the four steels were sliced, electro-discharge

machined to the required 3mm disc size and mechanically polished to 0.1mm thickness.

These thin sections were then electropolished using a solution of 5% perchloric acid in

acetic acid to electron transparent thickness suitable for transmission electron

microscopy. The samples were then examined under a Phillips 200 transmission

electron microscope.

133
Chapter 6 Experimental Methods

6.4 ELECTRICAL RESISTIVITY

Initially it was intended that the sequence of the strain ageing behaviour be studied by

electrical resistivity measurements. This meant that a thermal simulation of rod rolling

and wire drawing steps be carried out coupled with sensitive electrical resistance

measurements. Initial tests on studying quench ageing behaviour on these low carbon

steels did not provide very consistent data, as a result of the low sensitivity of the

experimental apparatus used to measure the resistivity. Therefore this technique was

discarded.

6.5 HOT TENSION TESTS

Initially, hot tension tests were carried out on a limited number of samples in an attempt

to evaluate the strain ageing behaviour. The materials, which were received as 5 mm

rods were tested in the as received condition for this test. A fixed length of 40 mm

from the as received rods served as the test piece. A small resistance-heating furnace

was loaded onto an Instron tensile testing machine. The sample holders, together with

the sample, were fixed in such a way that the sample was held in the hot zone of the

furnace throughout the test duration. The material rods were heated to a temperature of

either 130°C or 200°C and soaked for a period of 5 minutes for temperature uniformity.

Uniaxial loads were then applied on the samples at a strain rate of 0.3/second. After an

initial prestrains varying between 0.1 - 0.2, the load was removed and the sample

soaked at this temperature for either 30 seconds or 600 seconds. Subsequently, the

samples were tested to failure at this temperature.

134
Chapter 6 Experimental Methods

6.6 HOT TORSION TESTS

The torsion test has been used over a long period of time to assess the hot workability of

.
vanous . 1sc212s111-119)Th
matena ' ' . e advantages ofhh · test me
t e ot torsion · Id
u e

• in-situ simulation of strain ageing which incorporates simulation of rod rolling and

wire drawing stages.

• capability of generating very high strains and strain rates which would be a closer

simulation of the actual wire drawing process<27)_

A hot torsion facility was designed and constructed by BHP Melbourne Research

Laboratories. This equipment previously had been extensively used for this type of

testing in other studies in which the test method and results were validated. All hot

torsion tests mentioned in the current study were carried out initially at BHP Research

Laboratories, Melbourne. Before completion of this research work, this laboratory had

closed and the hot torsion equipment was moved to Deakin University, where the

remaining experiments were completed.

The steels were received in bar form, 20 mm in diameter. To facilitate hot torsion tests,

these bars were machined to fabricate specimens of dimensions as shown in Figure 6.1.

The hot torsion rig used for the current study was capable of working up to temperatures

of 1000 °c and at strain rates as high as 350/sec. The rig was horizontal type with the

specimen bolted onto its heads, one movable and the other fixed. The brake was a

pneumatically operated clutch brake, which was capable of stopping the twisting in a

very short time. The efficiency and consistency of this rig was extensively evaluated

before actual tests 011 -119 )_

135
Chapter 6 Experimental Methods

Prior to testing, the steels were austenitised at 1100°C for 30 minutes and then cooled to

950 °C at the rate of 2°C/sec. The steel specimens were then torqued to an initial strain

of 1.05, in an attempt to simulate the rod-rolling step. The specimens were then cooled

to 600°C at a rate of 3°C/sec and then to room temperature at the rate of 1-2°C/sec and

baked at 170°C for 40 minutes. This cooling sequence was an exact replication of the

cooling sequence used in the actual production sequence. Following this, the specimens

were heated to the wire drawing temperature of either 130° or 200°C and then torqued

to a strain of 0.33, the nearest experimentally possible simulation of the wire drawing

step. The temperatures chosen are again the temperatures commonly used during wire

drawing. The specimens were held after the initial strain for a period of either 30

seconds or 600 seconds and then torqued to failure. The delay time chosen represent the

extremes that would probably be encountered in the actual production sequence. This

procedure is illustrated in Figure 6.2. The stress-strain data was then determined and

plotted from the hot torsion tests through a software package, which converted surface

strain and torsion stress to true stress and strain.

Results of the tests that are described in this Chapter are outlined in the next Chapter.

136
Chapter 6 Experimental lvlethods

k<----------1~0 .. ~ - - - - - - - - - - - 1
?.
t kl,..,.. i-----4•- ~3.,.,,r, ~
•'II
.,. ~ ~
?
'"°' ,.... l

."
·----- - - - --lM y ••1
--=t---l----·-·--
'Ji: ... h --
.,:,

"'&...
r"
~
~
~
'Q. 'Q.
(S ~ ~
~
,,c
1
l
0
1-1
.)
"Q.
0

-~ 3 3 )
I
1
·!
' It

in
i
I :S>
~
%.

Figure 6.1 Typical specimen geometry of a hot torsion test sample.

137
<.__-:hapter 6 Experimental lvlethod.r;;

t .30 rn.i0u.-le.s

1-oi.bal s-lt-a..i..n I· 05
9BOc

u
"'
1-!J
et.
::)
I-
~
C::
\JJ
Cl.
t
l-~ 0c/sQ.c

I. \
~co c .sfro..in 0·33
0

~ ~ ' hald ~ ro rnc._n or


0

l':70 C
o-~ nun
40 .
~ - - - \ 130°c. slra.in 0·33
ha.ld ~ to min. m-
e, s ,.,..,'-".

t 0,9,UJl.d U> f'o.i.lurQ. .

Figure 6.2 Test schedule used for bot torsion tests

138
Chapter7 Results

CHAPTER 7

RESULTS

139
Chapter 7 Results

This chapter describes the results obtained in various tests on the four steels. Section

7.1 shows the results of the optical microscopic examination of sections cut from the as-

rolled rods of the four steels. Section 7.2 describes the results of the measurements of

free nitrogen content in the four grades of steels. Results of room temperature

mechanical tests on the steels are described in Section 7.3, while section 7.4 shows the

TEM photographs of the sections of as-rolled steels. Sections 7.5 and 7.6 show the

results obtained from hot tension and hot torsion tests respectively.

7.1 OPTICAL MICROSCOPIC EXAMINATION:

Figures 7 .1 - 7.4 show the microstructures of the four grades of steel m the "as

received" condition i.e " as rolled" condition.

Figure 7.1 Optical microstructure of BOS low nitrogen steel

140
Chapter7 Results

Figure 7.2 Optical microstructure of BOS high nitrogen steel

Figure 7.3 Optical microstructure of EAF low boron steel

141
Chapter7 Results

Figure 7.4 Optical microstructure of BOS high boron steel

Table 7.1 Image analysis

PROPERTY BOS LOWN BOSHIGHN EAFLOWB EAF HIGH

GRAIN SIZE ASTM8 ASTM8 ASTM8 ASTM6-7


( 18-~~ ,Ull') (i8-Q~_?l'I) G s - oRQ µm) ( Qq-uoµm
Volume 6.7 7.1 8.1 8.8

%PEARLITE

The microstructures were characterised using an image analysis system (Quantimet).


Table 7.1 summarises the image analysis results of these microstructures. It is obvious
from the above microstructures and the image analysis results that the BOS low and
high nitrogen grades and the EAF low boron grade appear to show a similar grain size.
However, the microstructure in the BOS high boron steel grade shows a slight coarser
grain size. The amount of pearlite seems marginally greater in the boron containing
grades. It is to be noted that the dark spots that appear on these micrographs are not

142
Chapter7 Results

precipitates or second phase particles, but are polishing and etching artefacts that have
resulted due to the softness of the steel grades.

7.2 DETERMINATION OF "FREE NITROGEN"CONTENT IN THE STEEL GRADES

The hydrogen hot extraction method was used to determine the amount of free nitrogen

in the samples, according to the procedure outlined in reference (120). The steel

samples were cut into pieces of an approximate size of2-3 mm which were used for this

experiment. These chips were placed in a silica crucible and annealed in a stream of

99.9 % pure hydrogen at 450°C under a gas flow rate of 200 cc/minute for a period of

one hour. The samples were then cooled and analysed for nitrogen content using

LECO TC 436 instrument. Samples of the original material were also analysed using

the LECO analyser for total N content. Results are tabulated below. All values are in

wt%.

Table 7.2 N·t


1 roeen anaw . of steel era des aft er hot ex tracf10n w1·th
l sis hlYd roeen.
Material Total Total Total N N content Difference
N B (LECO) after hot (A-B)
(A) extraction attributed to
(B) free nitroeen
BOSLowN 0.002 - 0.003 nd nd
BOSHiehN 0.008 - 0.008 nd nd
EAFLowB 0.008 0.0012 0.0078 0.0024 0.0054
BOS Hieb B 0.006 0.0038 0.0069 0.0026 0.0043
Nd- not determined

7.3 ROOM TEMPERATURE MECHANICAL PROPERTIES:

Rods from the four steel grades were tested for room temperature mechanical properties

using an Instron machine. These results are tabulated in the following table7.3.

143
Chapter7 Results

Table 7.3 R oom t empera t ure mec h amca


. I properties of stee s used"m current stu dy
MATERIAL 0.2% ULTIMATE %

PROOF TENSILE ELONGATION

STRESS (MPa) STRENGTH (MPa)

BOS LOWN 30.9 40.3 53.9

BOSHIGHN 32.1 42.9 53.4

BOSHIGHB 22.8 34.5 61%

EAFLOWB 32.3 42.6 50.9

7.4 TRANSMISSION ELECTRON MICROSCOPIC EXAMINATION:

Some typical features in the four steel compositions in the as-rolled condition as

observed under the TEM are shown below. No distinct differences were observed in

these steel grades.

Figure 7.5 (a) TEM micrograph of a BOS grade high nitrogen steel
144
Chapter? Results

Figure 7.5 (b) TEM micrograph of a BOS high boron steel

7.5 HOT TENSILE TESTS:

Figure 7.6 shows the stress strain curves obtained from hot tension for the BOS low N

grade steel. Similar results are shown in Appendix A. Problems encountered during the

test were

• the inability to apply a consistent prestrain with the present set-up.

• Temperature control was within± 15 °C and

• Total prestrain that could be applied was very small as compared to actual

strains achieved during wire drawing. The order of strains used in drawing operations

are of the order of a minimum of 0.5 and above. However in a hot tension test a

maximum strain of approximately 0.06 can be applied. Moreover, the strain rates used

in hot tension are also only a fraction of the actual strain rate used in wire drawing.

Therefore hot tension was not a close simulation of the drawing process and this testing

was not pursued further.

145
Chapter7 Results

Stress strain curve at 130 C - 47757(1) - increasing


prestrain - 10 minutes
(BOS Low N)
50
45 0.14 prestrain
40
Stl35
es30 0.1 prestrain
~i?5
a)20
15
10
5
0
0 0.2 0.4 0.6
Strain

Figure 7.6 Consistency in hot tension tests

7.6 HOT TORSION TESTS

Figure 7.7 shows a typical true stress strain curve derived from engineering stress- strain

values obtained from a hot torsion test. X-axis values are true normal strain values and

the true normal stress values in megapascals are plotted on the y-axis. As already stated

earlier, the two different parts of the curve are labelled in the Figure 7.7. The initial part

of the curve labelled as A is the prestrain and the subsequent part is the stress-strain

behaviour when tested to failure after holding for a period of time following initial

prestrain for a period of either 30 seconds or 600 seconds.

146
700

600

-
CU
D.
500

-
:al:: 400
en
en
W 300
a::
I-
en
200

100

0
0 0.5 1 1.5 2 2.5 3
STRAIN

Figure 7.7 A typical true stress-strain curve derived from the hot torsion data.

147
Chapter 7 Results

7.6.1 Consistency of hot torsion tests

The consistency of the stress-strain data is illustrated in Figure 7.8. The figure shows

the stress-strain curves obtained on repeat runs on a specific steel composition under the

same parameters of temperature, prestrain and time of holding after prestrain. An

important point to note is that the level of prestrain is consistent in all tests. The values

of flow stress appear fairly consistent with a scatter of± 10 Mpa (a percentage error of

5%). However the strains to failure show a larger scatter in the range of 0.2(8% error).

The magnitude of these errors were taken into account when interpreting test results.

Figure 7.9 shows the stress strain curves obtained on repeat runs on the same

composition of steel tested under the same parameters with a superimposed error band.

The stress-strain values have been compared only when the difference in the absolute

readings consistently exceed the error limits specified.

7.6.2 Hot torsion test results:

The true stress - strain curves of samples tested under hot torsion are shown in

Appendix B. These curves show the variation of true stress as a function of true strain.

148
Chapter 7 Results

This page has been intentionally left blank

149
700

600

_ 500
CU
a.
6 400
en
en
w 300
0:::
..,_
en 200

100

0
0 0.5 1 1.5 2
STRAIN

Figure 7.8 Consistency of hot torsion results on EAF steel containing low boron and high nitrogen, when tested at 200°C and a
delay time of 30 seconds after prestrain.
150
700 , - - - - - - - - - - - - - - - - - - - - - - - - - - -

600

-ea 500
D.
~ 400
en
en
W 300
0::
I-
en 200
100

o---------------.-------- -.--.....a.---------------
0 0.5 1 1.5 2 2.5 3

STRAIN

Figure 7.9 (a) Error band calculated for the stress values.

151
700 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

600

ci 500
D.
~ 400
en
en
W 300
0:::
I-
en 200

100

0 -L----.....--------.-----~~~5'-----,--------,----~
0 0.5 1 1.5 2 2.5 3

STRAIN

Figure 7.9 (b) Error band calculated for the strain values.

152
Chapter 7 Results

7.6.2.1 Influence of temperature on the stress-strain behaviour of the four steel

compositions:

Figures 7 .10 to 7.17 compare the stress strain curves of the four steels as a function of

test temperature. It is evident that the nature of the curves in the region of the initial

prestrain are different. At the higher test temperature of 200°C, the initial part of the

curve (during prestrain) appears to have a higher slope. After the initial prestrain, the

curves show a similar behaviour in the strain to failure region. This trend is evident in

all the four steel compositions and for both the delay times. The flow curves at 200°C

appear more serrated than those at 130°C, probably indicative of more dynamic strain

ageing at 200°C than at 130°C.

153
700 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - -....

600 0
@130 C Aging index= 71.SMPa

@200 C Aging index =24.6 Mpa
500

'ii
D. 400
:a:
ui
rn
w
~
I- 300
m

200

100

o ....
0
-----...-------...-------...---------...-------...-----------1
0.5 1.5 2 2.5 3
STRAIN

Figure 7.10 Stress strain curves of BOS low N steel as a function of test temperature tested after a delay time of 30 seconds after
prestrain.
154
700

600 0
@130 C Aging index= 83.1 Mpa
0
@ 200 C Aging index =38.2 Mpa
500

'j 400
:iii:
'ii,
Ill

,,,i 300

200

100

0
0 0.5 1 1.5.
S tram 2 2.5 3

Figure 7.11 Stress strain curves of BOS high N tested as a function of temperature tested after a delay time of 30 seconds after
prestrain.

155
700 - - . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
Cl>

600 @130 C Aging index =21.4 Mpa


C>
@ 200 C Aging index =85.2 MPa

500

-CIS
400
-
CL
:E
0

-
0
f!! 300
en

200

100

0+---------,------,------~ -------,.------,---------1
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.12 Stress strain curves of EAF low B steel as a function of temperature tested after a delay time of 30 seconds after
prestrain.

156
700
@ 130'\: Aging index = 69.9 MPa
0
@ 200 C Aging index = 54.2 MPa
600
200C

500

l::E: 400
'in
II)

i
en
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.13 Stress strain curves of BOS high N+B steel as a function of temperature tested after a delay time of 30 seconds after
prestrain.

157
700
0
@ 130 C Aging index= 67.7 MPa

600 •
@ 200 C Aging index= 20.3 MPa

500
-C'II
o.. 400
-
:::!!:
U)

-
U)
f 300
tn

200

100

0
0 0.5 1 1.5 2 2.5 3

Strain

Figure 7.14 Stress strain curves of BOS low N steel as function of temperature tested after a delay time of 600 seconds after
prestrain.

158
700 - . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
0
@ 130 C Aging index =21.4 Mpa
600 ~
@ 200 C Aging index =85.2 MPa
200 C

500 130 C

-"'
a. 400
-
:E
I ll

-
Ill
~ 300
u,

200

100

0 +-------...--------,------~--------,--------,----------'I
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.15 Stress strain curves of EAF low B steel as a function of temperature tested after a delay of 600 seconds after
prestrain.

159
700 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
0
@ 130 C Aging index =85.5 Mpa
600 0
@200 C Aging index =31.8 Mpa

500

'i:!!: 400
U)
en

-
! 300
u,

200

100

0 - - - - - - - - - - - -.........- - - - - - - - - - - - - - . . . . - - - - - -
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.16 Stress strain curves of BOS high N steel as a function of temperature tested after a delay time of 600 seconds after
prestrain.

160
700 - . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

@ 130 C Ageing index =116.4 MPa
600 •
@ 200 C Aging index= 84.2 MPa

500

'ii 200C
ll. 400
:E
I2!
-
en
300

200

100

0 -!----------,---------,------ --r----------r------.......,:i --....- - ~


0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.17 Stress strain curves of BOS high N+B steel as a function of temperature tested after a delay of 600 seconds after
prestrain.

161
Chapter 7 Results

7.6.2.2 Influence of delay time after prestrain on the stress strain behaviour of the

four steel compositions:

Figures 7.18 to 7.25 compare the stress strain curves of the four different steels as a

function of delay times after prestrain. There seems to be no major difference in the

behaviour of the materials at different times of delay except for a slight difference in the

total strain to failure in several samples. These variations in strain to failure are

believed to be due to random experimental errors, and they are not considered to be

significant. There seems to be no appreciable difference in the stress levels or the

nature of the curves at the two delay times.

162
700
@30 SECONDS Aging index= 71.SMPa

600 @ 600 SECONDS Aging index= 67.7 MPa

600SECONDS
500

'ii 30SECONDS
a.. 400
!!.
Ill
Ill

-~
u,
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.18 Stress strain curves of BOS low N steel as a function of delay time after prestrain tested at 130°C.

163
700 @ 30 SECONDS Aging index=83.1 MPa

600 @ 600 SECONDS Aging index =85.5 MPa

500
-
ns
a. 400
600SECONDS

-
:iE
,n
,n

-
f 300
en
200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.19 Stress strain curves of BOS high N steel as a function of delay time tested at 130°C.

164
700
@ 30 SECONDS Aging index =85.2 MPa
600
@ 600 SECONDS Aging index= 149.9 Mpa

500

-
CU
400
-
0.
:iE
,n
,n

-
f 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.20 Stress strain curves of EAF low B steel as a function of delay time tested at 130°C.

165
700 • @ 30 SECONDS Aging index = 69.9 MPa 1

@600 SECONDS Aging index=116.4 MPa


600
600 seconds
500

-CG
a. 400
30 seconds

-
:::E:
,n
,n

-! 300
u,

200

100

-----,-------,-------,-------,------~-----!
0 .....
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.21 Stress strain curves of BOS high N+B steel as a function of delay time after prestrain tested at 130°C.

166
700
@ 30 SECONDS Aging index= 24.63 MPa
600 @ 600 SECONDS Aging index= 20.3 MPa

500

-
CU
a. 400
30 SECONDS

-
:E
U)

-
U)
f 300
U)

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.22 Stress strain curves of BOS low N steel as a function of delay time after prestrain tested at 200°C.

167
700
600SECONDS
@ 30 SECONDS Aging index =38.2 MPa
600 @ 600 SECONDS Aging index =31.8 MPa

-I
ea
500 30SECONDS

-i
en
400

-
...
en
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.23 Stress strain curves of BOS high N steel as a function of delay time after prestrain tested at 200°C.

168
700
@ 30 SECONDS Aging index =21.4 Mpa
600
@ 600 SECONDS Aging index =26.5 Mpa

500 I
-ns
400
4 30SECONDS

-!
ll.
:ii!:
I I)

-
en
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.24 Stress strain curves of EAF low B steel as a function of delay time after prestrain tested at 200°C.

169
700
@ 30 SECONDS Aging index = 54.2 Mpa

600 @ 600 SECONDS Aging index= 84.2 MPa

30 SECONDS
500
600 SECONDS

-
:._ 400
-t
:ii:
,n
,n
300
u,

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.25 Stress strain curves of BOS high N+B steel as a function of delay time after prestrain tested at 200°C.

170
Chapter 7 Results

7.6.2.3 Influence of material composition on the stress strain behaviour of the

four steel compositions:

Figures 7.26 to 7.29 depict the stress strain curves of the four different material

compositions at different test temperatures and delay times after prestrain. It is

important to note that in general, a very similar trend is exhibited by the materials at

various test parameters. There is no significant difference in the behaviour of the BOS

high N and the EAF low boron grade, as observed in Figures 7.26, 7.27 and 7.28. The

highest value of total strain to failure is exhibited by BOS high N+B material, followed

by BOS low N, EAF low Band lastly BOS high N. This is true for all combinations of

test temperatures and delay times as observed in Figures 7.26-7.29. In addition, BOS

high N+ B grade appears to show an appreciable decrease in the stress values as

compared to the other steel grades as observed in Figures 7.26-7.29.

In general as observed from figures 7.26-7 .29, the level of the stress-strain curves

follow a definite order - BOS high N+ B is the lowest followed by BOS low N and then

by BOS high N and EAF low B.

The slope of the stress-strain curves in the prestrain region, which is indicative of work

hardening appears to be approximately the same and relatively a higher value for BOS

high N and EAF low B steels at both the test temperatures and delay times. This is

followed by the BOS Low N steel and then the BOS high N+ B steel. A similar trend is

noticed in the strain to failure region for the four grades at all combinations of test

temperatures and times.

171
650

600

550

500
ii
a.
450
:ii: BOS High N
ui 400 SLowN
tfj
~
-
UJ 350
BOS Hidh N+B
300

250

200

150
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.26 Stress strain curves of the four material compositions tested at 130°C and 30 seconds delay time after prestrain.

172
650

600 l __ t\..z) eat EAFLowB


550

500

'i_ 450 i 1.........-- BOS High NI I I BOS High NtB


:E
wtn
400
Cl)
a..
I NII L I IBOS Low N
u, 350

300

250

200

150
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.27 Stress strain curves of the four material compositions tested at 130°C and a delay time of 600 seconds after prestrain.

173
650

600

550

500
-ns 450
0..
:E: BOS Hidh N + B
BOS High N
;; 400
U)
G)

b 350
Cl)
BOS Low N
300

250

200

150
0 0.5 1 1.5 2 2.5 3
Strain

Figure 7.28 Stress strain curves of the four material compositions tested at 200°C and a delay time of 30 seconds after prestrain.

174
650 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
EAF LowB
600

550

500

ci 450
D. BOS High N+B
:il:
"; 400 BOS High N
VI
f
u; 350

300 BOS LowN

250

200

150 __.,_-----+------.--------,.--+----.....--+----.1.------,-------+---------l

0 0.5 1 1.5 2 2.5 3


Strain

Figure 7.29 Stress strain curves of four material compositions tested at 200°C and a delay time of 600 seconds after prestrain.

175
Chapter 7 Results

7.6.3 Results of transmission electron microscopic examination:

Sections parallel to the fracture surface of the hot torsion samples were prepared for

examination under the transmission electron microscope according to the procedure

described in section 7.4. Several micrographs are shown in Figures 7.30. to 7.35.

From the micrographs it is evident that there are no precipitates, small or large, that

are present in any of the samples examined of the four different material compositions

or under any combination of test parameters. The ferrite matrix does not contain any

evidence of any BN or any other nitride or carbide precipitates. Selected area

diffraction patterns were observed and confirmed the above findings. They were

however not included in the thesis as they were not of a very high photographic

standard.

Figure 7.30 Transmission electron micrograph showing a very high density of


dislocations and a matrix devoid of precipitates in BOS low nitrogen grade of steel
tested at 200°C after a delay of 600 seconds.

176
Chapter 7 Results

Figure 7.31 Transmission electron micrograph showing a very high density of


dislocations and a matrix devoid of precipitates in BOS high nitrogen grade of steel
tested at 200°C after a delay of 600 seconds.

0.2µm

-
Figure 7.32 Transmission electron micrograph showing a very high density of
dislocations and a matrix devoid of precipitates in EAF low boron grade of steel
tested at 130°C after a delay of 600 seconds.

177
Chapter 7 Results

Figure 7.33 Transmission electron micrograph showing a very high density of


dislocations and a matrix devoid of precipitates in BOS high nitrogen + boron
grade of steel tested at 130°C after a delay of 600 seconds.

Figure 7.34 Transmission electron micrograph showing a very high density


of dislocations and a matrix devoid of precipitates in BOS high nitrogen +
boron grade of steel tested at 200°C after a delay of 30 seconds.

178
Chapter 7 Results

Figure 7.35 Transmission electron micrograph showing a very high density of


dislocations and a matrix devoid of precipitates in BOS high nitrogen + boron
grade of steel tested at 200°C after a delay of 600 seconds.

179
Chapter 8 Discussion

CHAPTERS

DISCUSSION

180
Chapter 8 Discussion

8.1 OPTICAL MICROSCOPY EXAMINATION

The optical microstructures of the BOS low and high grades of steel and the EAF low

boron steel grade do not show much difference in the grain size. This is most probably

due to the very minor difference in the compositions of the grain refiners like Al. The

BOS low N and the EAF low B grades contain 0.003% Al as compared to 0.004% in

BOS high N grade. Although the EAF grade contains an addition of boron which can

form fine precipitates that can restrict grain growth [Franks and Kirkaldy< 73 >; Clark and

Vamey< 8>], the amount added is so small as to not make any appreciable effect. In

contrast, the BOS high boron steel shows a marginally coarser grain size, although it

contains a marginally higher amount of Al (0.006%). This is attributed to the fact that

boron has not been able to form grain growth restricting precipitates like BN. This

conclusion is supported by evidence from transmission electron microscopic

observations as shown in section 7.6.2. However it must be noted that the increase in

grain size is only marginal and is discussed here on the basis of the observations made

by Takahashi and Shibata-Faulring<54>_

8.2 ROOM TEMPERATURE MECHANICAL PROPERTIES

The room temperature mechanical properties of the BOS low and high nitrogen grades

and the EAF low boron grade appear similar. This is despite the fact that two of these

grades contain high nitrogen, which is known to increase work hardening (there is only

a marginal increase in the strength values). However the BOS high nitrogen and high

boron grade do show a decrease in the strength values (both yield and ultimate tensile

stress) and an increase in the elongation. This could be attributed in part to the increase

in grain size observed in the high boron grade steel.

181
Chapter 8 Discussion

8.3 HOT TORSION PROPERTIES

8.3.1 Influence of temperature on the stress - strain behaviour of the four steel

compositions:

8.3.1.1 Work hardening exponent

The nature of the stress strain curve at 200°C is very different from that exhibited at the

lower temperature of 130°C. In the prestrain region, the 200°C curve appears to have a

higher slope. The slope of a true stress-strain curve indicates the work hardening

exponent. Therefore, the work hardening exponents have been computed, as shown

Figures C-1 to C-31, Appendix C and the exponents are listed in Table 8.1. It may be

noted however that the current study is under triaxial stress states and values of work

hardening exponents computed may not necessarily be the same as in uniaxial tension

state. The work hardening exponent has been computed for both the prestrain region

and the strain to failure region separately. In order to minimise errors in comparison,

stress values at fixed values of strain (0.35 and 1.08) have been used to calculate the

exponent for all materials and test parameters. Figures C-1 to C-15 show the calculation

of the exponent in the prestrain region. Figures C-16 to C-31 show the calculation of

the exponent in the strain to failure region. For ease of comparison, these values have

been shown on a bar chart in Figures 8.1 and 8.2. It is evident from the bar charts that

at both 30 and 600 seconds, the work hardening exponents at higher temperature of

200°C are higher as compared to the exponent values at the lower temperature of 130°C.

This is attributed to the occurrence of dynamic strain ageing at higher temperatures

where both interstitial carbon and nitrogen have diffusion coefficients sufficiently high

in order to pin the dislocations dynamically during straining itself. A comparison of the

diffusion coefficients at l 30°C and 200°C indicate that the coefficient at 200°C is

approximately 125 times higher than at the lower temperature (De 20ocl De 130c = 30, as
182
Chapter 8 Discussion

per equation 4.14). This trend is generally observed in all the four steel compositions.

However the actual values of the exponents differ from material to material. This

observation is explained in section 8.3.3.

Table 8.1 Values of work hardening exponents of the four different steel
compositions under various test parameters.

Material Work hardenin2 exponent


130°C 30 seconds 130°C 600 seconds
Prestrain Strain to Prestrain Strain to
failure failure
BOSLowN 0.28 0.17 0.27 0.17
BOSHi2hN 0.30 0.15 0.24 0.15
EAFLowB 0.24 0.15 0.25 0.16
BOS High 0.23 0.07 0.17 0.06
N+B

Material Work hardenin2 exponent


200°C 30 seconds 200°C 600 seconds
Prestrain Strain to Prestrain Strain to
failure failure
BOSLowN 0.30 0.12 0.30 0.12
BOSHi2hN 0.31 0.13 0.34 0.13
EAFLowB 0.35 0.12 0.35 0.13
BOS High 0.30 0.03 0.30 0.01
N+B

183
0.400 -~ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

D130C
0.350
•2ooc
~ 0.300
a,
C
0
a. 0.250
X
(l)
Cl
-~ 0.200
C
a,
...
"O
1 0.150
...
.:,i:

0
~ 0.100

0.050

0.000
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.1 Comparison of work hardening exponents in the prestrain region as a function of temperature tested at 30 seconds delay after
prestrain.

184
0.400

0.350 0130 C
• 600 seconds
cC1)
o.300
C:
0
~ 0.250
Cl)
i.J1
C:
"i: 0.200
Cl)
.::,
,._
~ 0.1 50
.:,::
,._
0
S: 0.100

0.050 _,

0.000
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.2 Comparison of work hardening exponents obtained in the prestrain region as a function of temperature after 600 seconds
delay after prestrain.

185
Chapter 8 Discussions

8.3.1.2 Maximum stress m/ues

Figures 8.3 to 8.10 compare the stress strain behaviour of the four materials as a

function of temperature. Figures 8.11 to 8. 13 show the maximum stress values obtained

after prestrain alone and after 30 and 600 seconds delay after prestrain respectively. It

is observed in Figure 8.11 that the values of maximum stress obtained at 200°C are

higher than obtained at 130''C. This is due to the occurrence of dynamic strain ageing at

200°C (Figures 8.1 and 8.2).

In the strain to failure region (Figures 8.12 and 8.13), there does not seem to be a

significant difference in the maximum stress values. Generally the curve at 200°C is

marginally higher than that exhibited at 130°C.

It appears that the test temperature has a greater effect in the prestrain region than in

the strain to failure region.

186
700
0
@ 130 C Aging index= 149.8 Mpa

600 •
@ 200 C Aging index = 26.5 Mpa

500

-ea
D. 400
-
:!:
I ll
Ill

-~ 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.3 Stress strain behaviour as a function of temperature BOS Low N tested at 30 seconds delay after prestrain.

187
700
@ 13o"C Aging index= 83.1 Mpa
@200°C Aging index =38.2 Mpa
600

500

-C'tl
a. 400
-
~
f/J

-
f/J
f 300
V,

200

100

0 -,
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.4 Stress strain behaviour as a function of temperature of BOS High N steel tested at 30 seconds delay after prestrain.

188
700
0
@ 130 C Aging index =21.4 Mpa
600 0
@ 200 C Aging index =85.2 MPa

500

- 400 -
ns
a.
-
::i!:
1/)
1/)

-
f 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.5 Stress strain behaviour as a function of temperature of EAF low B steel tested at 30 seconds delay after prestrain.

189
700 . . . . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
0
@ 130 C Aging index= 69.9 MPa
600
@ 200•C Ag!!!g index = 54.2 MPa
200C --__L=

500
130 C
-
~ 400
-"'
:iE
I I)

~ 300
en

200

100

0 + - - - - - - - - - , - - - - - - - - - . - - - - - - , - - - - - - - , - - - - - - - - , -.......- - - - - - - 1
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.6 Stress strain behaviour as a function of temperature of BOS high N+B steel tested at 30 seconds delay after prestrain.

190
700
0
@ 130 C Aging index= 67.7 MPa

600 @ 200°c Aging index= 20.3 MPa

500

-
~ 400
-
~
I ll
Ill

-~ 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3

Strain

Figure 8.6 Stress strain behaviour as a function of temperature of BOS Low N steel tested at 600 seconds delay after prestrain.

191
700
0
@ 130 C Aging index =85.5 Mpa
600 0
@ 200 C Aging index =31.8 Mpa

500
-C'G

~ 400
-en
en
~ 300
I-
en
200

100

0
0 0.5 1 1.5 2 2.5 3
STRAIN

Figure 8.8 Stress strain behaviour as a function of temperature of BOS high N steel tested at 600 seconds delay after prestrain.

192
700
0
@ 130 C Aging index= 149.8 Mpa

600 0
@ 200 C Aging index = 26.5 Mpa

500

-
:_ 400
-
:::i!1:
I I)
II)

-~ 300
Cl)

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.9 Stress strain behaviour as a function of temperature of EAF Low B steel tested at 600 seconds delay after prestrain.

193
700
@ 13o•c Ageing index =116.4 MPa

600 @ 200•C Aging index= 84.2 MPa

500

-ea
a. 400
-
:E
I ll
Ill
....~
u,
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.10 Stress strain behaviour as a function of temperature of BOS high N+B steel tested at 600 seconds delay after
prestrain.

194
700 -
D 130°C
0

600 •2ooc
500 ~

-n:s
a. 400 -
-
:E
1/J
1/J

....~
Cl)
300

200

100

0 ~

BOS Low N BOS High N EAF Low B BOS High B


Material

Figure 8.llComparison of maximum stress values obtained in the prestrain region as a function of temperature.

195
700 D 130°C

•2ooc
600 -

500 -

-
I ll
a. 400 -
-e
~
I ll
Ill

...
Cl)
300

200 -

100

0 _____.__ _

BOS LowN BOS High N EAF LowB BOS High B


Material

Figure 8.12 Comparison of maximum stress values in the strain to failure region as a function of temperature after a delay of 30 seconds
after prestrain.

196
700
013o·c
•200°c
600

500 -

-
~ 400
-
:E
1 /)
1/)

...e 300 -
Cl)

200

100

0
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.13 Comparison of maximum stress values as a function of temperature after a delay of 600 seconds after prestrain.

197
Table 8.2 Values of maximum stress obtained in the four different compositions in the prestrain and strain to failure regions
under various test parameters

Material Maximum Maximum stress Maximum stress Maximum stress Maximum stress
stress at at 130°C and 30 at 130°C and 600 at 200°C and 30 at 200°C and 600
prestrain seconds delay seconds delay seconds delay seconds delay
(MPa)
BOSLowN 373.48 565.1 550.97 572.95 575.65
BOS High N 399.46 590.81 582.15 594.87 609.75
EAFLowN 406.5 586.48 616.25 615.44 632.22
BOS High N+B 345.5 552.86 563.74 554.21 557.45

198
Chapter 8 Discussions

8.3.1.3 Total strain to failure

Table 8.3, Figures 8.13 and 8.14 compare the values of total strain as a function of

temperature. It is evident from these figures that there appears to be a very marginal

difference in the strain values at the two temperatures. Most of the differences in the

strain values between the two test temperatures fall within the random experimental

limits specified in section 7.6.1. Therefore it is seen that temperature appears to have

little or no effect on total strain.

199
Table 8.3 Values of total strain to failure of different steel compositions under various test parameters.

Material Total strain to failure


13o·c 30 seconds 13o•c 600 seconds 200•C 30 seconds 2oo·c 600 seconds
delay delay delay delay
BOSLowN 1.68 1.79 1.75 1.79
BOSHiehN 1.32 1.35 1.32 1.14
EAFLowB 1.48 1.47 1.43 1.64
BOS Hieb N+B 2.6 2.68 2.58 2.55

200
;,s~--------------------------------------
ID13<Tcl
l•200°c j

2.5

2
C:

...
ns
.....
Cl) 1.5
.....ns
0
I-
1 _,

0.5

0 _ ___.__ _

BOS Low N BOS High N EAF Low B BOS High N+B


Material

Figure 8.13 Comparison of total strains to failure as a function of temperature tested at 30 seconds delay after prestrain.

201
3---.----------------------------------,
013o·c
•2oo·c
2.5

2
C

...nl
-I I)
1.5
-nl
0
I-
1

0.5

0 ·-··
BOS Low N BOS High N EAF Low B BOS High N+B
Material

Figure 8.14 Comparison of total strains to failure as a function of temperature tested at 600 seconds delay after prestrain.

202
Chapter 8 Discussions

8.3.1.4 Ageing indices

Ageing index is defined as the difference in the value of stress before and after

prestrain. In order to provide a consistent measure of the difference in stress and

minimise errors in calculation of the ageing index, values of stress at a constant value of

prestrain (0.3) and stress at a constant value of strain after prestrain (0.35) have been

noted. The difference in these values has been assigned as the ageing index for the

material tested under a certain parameter.

Tables 8.4 and 8.5 list the values of ageing indices and Figures 8.15 and 8.16 compare

the indices on a bar chart. It is evident from the table and figures that the ageing index

is higher at 130°C than at 200°C. This is due to the occurrence of more dynamic strain

ageing at 200°C, which raises the level of the stress - strain curve in the prestrain

region. This in turn results in a lower ageing index. This behaviour is observed in

results obtained at both 30 and 600 seconds delay time after prestrain.

203
Table 8.4 Values of strain ageing indices in samples tested at 130 and 200°C after a time delay of 30 seconds after prestrain

TEMPERATURE STRESS AT INITIAL STRESS AFTER A STRAIN


MATERIAL OF TEST PRESTRAIN (MPa) DELAY TIME OF 30 AGEING
S1 SECONDS AFTER INDEX=
PRESTRAIN(MPa) S2 S2 - S1
BOS LOWN 130 C 373.48 444.93 71.45
BOS LOWN 200C 438.98 463.61 24.63
BOS HIGHN 130 C 399.46 482.55 83.09
BOSHIGHN 200C 471.73 509.81 38.16
EAF LOW BORON 130 C 406.3 491.48 85.18
EAF LOW BORON 200C 488.78 510.16 21.38
BOSHIGHN+B 130 C 345.64 415.52 69.88
BOSHIGHN+B 200C 423.89 478.12 54.23

204
Table 8.5 Values of strain ageing indices in samples tested at 130 and 200°C after a time delay of 600 seconds after prestrain.

MATERIAL TEMPERATURE STRESS AT INITIAL STRESS AFTER A STRAIN


OF TEST PRESTRAIN (MPa) DELAY TIME OF 600 AGEING
S1 SECONDS AFTER INDEX=
PRESTRAIN(MPa) S2 S2 - S1
BOS LOWN 130 C 361.85 421.51 67.66
BOS LOWN 200C 430.86 451.16 20.3
BOSHIGHN 130 C 410.29 483.36 85.52
BOS HIGHN 200C 486.88 518.68 31.8
EAF LOW BORON 130 C 320.26 470.1 149.84
EAF LOW BORON 200C 474.43 500.96 26.53
BOSHIGHN+B 130 C 360.21 476.62 116.41
BOSHIGHN+B 200C 394.48 478.66 84.18

205
90
13o·c o
80 · 2oo·c •

70

60

f
~
50
1/)
1/)

e 40
..,
Cf)

30

20

10

0
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.15 Comparison of ageing indices as a function of temperature tested at 30 seconds delay after prestrain.

206
160 -

130•C 0
140 200°c •

120

-
nl
c..
~
100

VI
VI
80

-...
Q)

Cl)
60

40

20

0
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.16 Comparison of ageing indices as a function of temperature at 600 seconds delay after prestrain.

207
Chapter 8 Discussions

8.3.2 Influence of delay time after prestrain on the stress - strain behaviour of

the four steel compositions:

8.3.2.1 Work hardening exponents:

Table 8.6 lists the values of the work hardening exponents and Figures 8.17 and 8.18

compare the values of the work hardening exponent obtained in the four steel

compositions as a function of delay time. These exponents have been computed in the

portion of the stress-strain curve after the time delay has occurred. There seems to be

no significant change in the value of the exponent with an increase in the delay time

after prestrain. The only significant variation is in the high boron + high nitrogen grade

which shows a marked decrease at the higher temperature of 200°C as the delay time is

increased from 30 seconds to 600 seconds.

Table 8.6 Work hardening exponents of the four grades of steels as a function of

delay time.

MATERIAL WORK HARDENING EXPONENT

130 °C 130 °C 200 °C 200°c

30 seconds 600 seconds 30 seconds 600 seconds

BOS LOWN 0.173 0.173 0.118 0.117

BOS HIGH 0.145 0.149 0.127 0.132

EAFLOWB 0.153 0.16 0.116 0.128

BOS HIGH 0.068 0.06 0.028 0.005


N+B

208
0.2

C30 seconds
0.18
•soo seconds
0.16 -

i: 0.14
Cl>
C
0
~ 0.12
Cl>
C")
C
C 0.1
a,
,._
-0

1 0.08
,._
~

0
~ 0.06

0.04

0.02

0 7

BOS Low N BOS high N EAF low B BOS high N+B


Material

Figure 8.17 Comparison of work hardening exponents in the prestrain region as a function of delay time after prestrain at 130°C.

209
0.14 ~ - - - -- - - - - - - - - - - -- - - - - - - -- - - - - - ~

030 seconds
• 600 seconds
0.12

0.1
i:
Q)
C:
0
c.
~ 0.08
Cl
.:C:
Q)
"O
~ 0 .06
.s::
-"'-
0
s:
0.04

0.02

O +-~ - ~
BOS Low N BOS high N EAF low B BOS high N+B
Material

Figure 8.18 Comparison of work hardening exponents in the prestrain region as a function of delay time after prestrain at 200°C.

210
Chapter 8 Discussions

8.3.2.2 Maximum stress values:

Figures 8.19 to 8.26 show the stress strain behaviour of the four steel compositions as

a function of delay time. Figures 8.27 and 8.28 compare the maximum stress values

obtained at both delay times. There appears to be no significant difference in the

maximum stress values obtained in the strain to failure region as a function of delay

time. This proves that all the solute has segregated to the dislocations within the first 30

seconds and no further addition is made to the amount of solute segregated at times

greater than 30 seconds.

8.3.2.3 Total strain to failure:

Figures 8.30 and 8.31 compare the values of total strain to failure as a function of delay

times. It is very evident that there is no, or very little difference in the strain values with

a variation in the delay time. This further strengthens the interpretation that there is no

addition in the amount of solute that has segregated to the dislocations with an increase

in the delay time.

211
700
@ 30 SECONDS Aging index= 71.SMPa
600
@ 600 SECONDS Aging index= 67.7 MPa

500

-
:_ 400
-
:!E
en
en

-e
Cl)
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.19 Stress strain behaviour of BOS low N as a function of delay time after prestrain tested at 130°C.

212
700 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

@ 30 SECONDS Aging index=83.1 MPa


600

@ 600 SECONDS Aging index =85.5 MPa


500 600 SECONDS

'i_ 400
-
::E
fl)

-
fl)

~ 300
en

200

100 ~

0 -------------,-----.....---------,--------,------
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.20 Stress strain behaviour of BOS high N as a function of delay time after prestrain tested at 130°C.

213
700
@ 30 SECONDS Aging index =85.2 MPa

600
30 SECONDS @ 600 SECONDS Aging index= 149.8 Mpa

500

-
C'CI
c.. 400
600 SECONDS

-
~
,n
,n

-
e
en
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.21 Stress strain behaviour of EAF low B steel as a function of delay time tested at 130°C.

214
700
@ 30 SECONDS Aging index = 69.9 MPa

600 600 seconds @ 600 SECONDS Aging index=116.4 MPa

500
-CU
o.. 400 30 seconds
-...,,,
:E

-: 300
u,
200

100

0
0 0.5 1 1.5 2 2.5 3

Strain

Figure 8.22 Stress strain behaviour of BOS High N+B steel as a function of delay time after prestrain tested at 130°C.

215
700
@ 30 SECONDS Aging index = 24.63 MPa

600 @ 600 SECONDS Aging index= 20.3 MPa

500

-
n,
a.. 400
-
:E
u,
u,

-
f 300
en

200 -

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.23 Stress strain behaviour of BOS Low N as a function of delay time after prestrain tested at 200°C.

216
700
@ 30 SECONDS Aging index =38.2 MPa
600 @ 600 SECONDS Aging index =31.8 MPa

-ea
~ 400
500
30 SECONDS

-u,
~ 300
.......
"' 200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.24 Stress strain behaviour of BOS High N as a function of delay time after prestrain tested at 200°C.

217
700
@ 30 SECONDS Aging index =21.4 Mpa
600
@ 600 SECONDS Aging index =26.5 Mpa

500

-"'
c.. 400 -
-"'
~
30 SECONDS
600 SECONDS

-"'
~
tn
300

200 -

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.25 Stress strain behaviour of EAF low B steel as a function of delay time after prestrain tested at 200°C.

218
700
@ 30 SECONDS Aging index = 54.23 Mpa

600 @600 SECONDS Aging index= 84.18 MPa


30 SECONDS

500
600 SECONDS

-
CU
a. 400
-
~
I ll
Ill

-
f 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure 8.26 Stress strain behaviour of BOS high N+B steel as a function of delay time after prestrain tested at 200°C.

219
700
rD 30 seconds-
600 • 600 seconds

500 -

-113
n.. 400
~
en
en
~ 300
....,
(/)

00

100

0
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.27 Comparison of maximum stress values in the strain to failure region as a function of delay time after prestrain tested
at 130°C.

220
700
a30 seconds
• 600 seconds
600

500 _,

('0
a.. 400
2

"'"'e 300
~

200

100

0 r

BOS Low N BOS High N EAF Low B BOS High B


Material

Figure 8.28 Comparison of maximum stress values in the strain to failure region as a function of delay time after prestrain tested at
200°C.

221
3
030 seconds

2.5

2 -
C

....ro
....,
(/)
1.5
ro
....,
0
I-
1

0.5

0
BOS Low N BOS High N EAF Low B BOS High N+B
Material

Figure 8.29 Comparison of values of total strain to failure as a function of delay time after prestrain tested at 130°C.

222
3 ---.---------------------
_ -_ -_ -_- - - - - - - - - ,
ID 30 seconds
~ 600 seconds
2.5

2
C:

...."'...
1/) 1.5
...."'0
I-

0.5

0
BOS Low N BOS High N EAF Low B BOS High N+B
Material

Figure 8.30Comparison of values of total strain to failure as a function of delay time after prestrain tested at 200°C

223
Chapter 8 Discussions

Figure 8.31 and 8.32 compare the ageing indices obtained at 30 and 600 seconds. A

significant difference in the index is observed in EAF low B and BOS high N+B with

an increase in the delay time after prestrain. However there is no significant difference

in the index with an increase in the delay time in BOS low and high N steels. This

could be attributed to the possibility of boron adding to strain ageing with the entire

amount boron not being fixed as BN. This is corroborated by the absence of BN

precipitates observed in the BOS high N+B steel under the transmission electron

microscope.

224
160
030 seconds
140 • 600 seconds

120

100
ro
a..
2
1/)
1/)
80

-
Q)
I,..

CJ)

60

40

20

0
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.31 Comparison of ageing indices as a function of delay time after prestrain tested at 130°C.

225
160
030 seconds
140 • 600 seconds

120
-
n:s
a.. 100
--
~
><
Q)
"O
C
80
-
0)
C
Q) 60
0)
<(

40

20

0
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.32 Comparison of ageing indices as a function of delay time after prestrain tested at 200°C.

226
Chapter 8 Discussions

8.3.3 Influence of material composition:

Figures 8.33 to 8.36 compare the stress strain behaviour of the four different steel

compositions under various test parameters. It is important to note that in general, a

very similar trend is exhibited by the materials at various test parameters. There is no

significant difference in the behaviour of the BOS high N and the EAF low boron grade,

as observed in Figures 8.34 - 8.35. The highest value of total strain to failure is

exhibited by BOS high N+B material, followed by BOS low N, EAF low B and lastly

BOS high N. This is true for all combinations of test temperatures and delay times as

observed in Figures 8.33 - 8.36. In addition, BOS high N+ B grade appears to

show an appreciable decrease in the stress values as compared to the other steel grades

as observed in Figures 8.34 - 8.37.

In general as observed from figures 8.33 - 8.36, the level of the stress-strain curves

follow a definite order- BOS high N+ B is the lowest followed by BOS low N and then

by BOS high N and EAF low B.

The slope of the stress-strain curves in the prestrain region, which is indicative of work

hardening appears to be approximately the same and relatively a higher value for BOS

high N and EAF low B steels at both the test temperatures and delay times. This is

then followed by the BOS Low N steel and then the BOS high N+ B steel. A similar

trend is noticed in the strain to failure region for the four grades at all combinations of

test temperatures and times

227
700 - --· ~-~- --- ---·~ -.----· --~~ ----

600

500

ii
c. 400
:ii:
en
en

~ 300 t.F t BOS High N I


l ,OS Low N

200 -- BOS Hiilh N+B

100

0
0 0.5 1 1.5 2 2.5 3
STRAIN

Figure 8.33 Stress strain behaviour of the four different steel compositions tested at 130°C and 30 seconds delay after prestrain.

228
700 - --· -· - - -=···-·----- -- - - - - - - - ---··-- ---··-·- --------------·

600

500

-
C'IS
a. 400
EAF LowB

-
:!:
en
en
w
BOS high N+B

o::: 300
I-
en BOS High

200 BOS LowN

100

0 + - - - - - - - - , - - - - - - - - - - . . - - - - - - - , - - - - - - - - - -...- - - - - J
0 0.5 1 1.5 2 2.5 3
STRAIN
Figure 8.34 Stress strain behaviour of the four steel compositions tested at 130°C and 600 seconds delay after prestrain.

229
'
700 r--··--·-·---· ·-------· - •o-- - - - - - •• --·-~-·-----· - - ·--· ·--~-·- ·---~-~,~~----·

600

500

ii BOS High N
a.
~ 400
en B@S High B
en
w Fllow B
a:: 300
I-
C/)

BOS LowN
200

100

0 ....
0 0.5 1 1.5 2 2.5 3
STRAIN

Figure 8.35 Stress strain behaviour of the four steel compositions tested at 200°C and 30 seconds delay after prestrain.

230
700 - - . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

600

500

'i::ii: 400 BOS High N

-
en
en
w
BOS High B
o::: 300 EAFLowB
t-
en
BOS LowN
200

100

0 ~

0 0.5 1 1.5 2 2.5 3


STRAIN

Figure 8.36 Stress strain behaviour of the four steel compositions tested at 200°C and 600 seconds after prestrain.

231
Chapter 8 Discussions

8.3.3.1 Work hardening exponents

Table 8.7 lists the values of the work hardening exponents obtained at vanous

combinations of the test parameters.

Figures 8.37 - 8.40 illustrate the variations in the work hardening exponent which may

be related to a variation in the interstitial content for the four steel compositions as listed

in Table 8.7. The variation in the work hardening exponent for the four steels tested at

130°C and 30 seconds, as shown in Figure 8.38, shows a marginal increase with an

increase in the interstitial content for BOS low and high nitrogen grades and the EAF

low boron grade. This is to be expected since the exponent depends very strongly on

the interstitials, which are very mobile and are capable of pinning the dislocations.

However an exception to this trend is noticed in the BOS high N+B grade wherein,

despite an interstitial content almost equal to the BOS high nitrogen grade, the exponent

is marginally lower (0.23 as compared to 0.30 for the BOS high N steel). The marginal

difference of 0.07 in the value of the work hardening exponent could be attributed to a

slightly lower content of nitrogen and fixing of some nitrogen by boron as indicated in

the "free nitrogen"analysis in Table. 7.2. In this context, the 'fixing' of free nitrogen

could be due to solute interaction between boron and nitrogen in solution, which could

explain the observed absence of precipitates, section 7.6.3. It must be noted that at the

lower temperature of 130°C, diffusion of nitrogen is higher than that of carbon. At

130°C and 600 seconds (Figure 8.39), the exponents for the BOS low N, BOS high N

and the EAF low B appear to be almost at the same value, while that of the BOS high

N+B steel is marginally lower (0.17 as against 0.24 - 0.27 for the other

232
Table 8. 7 Variation in work hardening exponents with the interstitial content for the four steel compositions.

Material Total solute Interstitial Work Hardening exponent


content content (C+N)

130°c 30 130°c 600 200°c 30 200°c 600

SECONDS SECONDS SECONDS SECONDS

BOS LOW 0.596 0.072 0.284 0.269 0.298 0.297


N
BOS HIGH 0.624 0.088 0.303 0.237 0.312 0.340
N
EAF LOW 0.760 0.058 0.244 0.250 0.354 0.353
B
BOS HIGH 0.660 0.086 0.228 0.173 0.295 0.295
B+N

233
0.15 0.400
I- Interstitial content ·!
+ 0.380
0.13 + 1-+-Work Hardening exponent
ca,
j
1
0.360
ca, 0.11 0.340
C
&,
+,I
C ><
0.320 a,
8 0.09 en
-:.::;ns o.3oo ·c
C

a,
:.::; 0.07
...a,
rn 0.280 "E
ns
+,I .c
C 0.05 0.260 ~ ...0
0.240 3:
0.03
0.220
0.01 0.200
BOS LowN BOS High N EAF LowB BOS High B
Material

Figure 8.38 Variation of work hardening exponents (prestrain region) and interstitial content for the four different steel
compositions tested at 130°C and 30 seconds

234
0.15
. -Interstitial content
0.14 I-+- Work Hardening exponent

~
0.13

0.12
0.360
-C
a,
C
0
s§ 0.11
c.
0.310 ><
a,

.
-
CJ
t'G

f? 0.09
0.1
0.260 'E
C)

-
C
C
a,

s t'G
.c
.i.::
C 0.08 I.
0
0.07 0.210 ~

0.06
0.05 - - - - - - - - - - - - - - - - - - - - - - 0.160
BOS LowN BOS High N EAF LowB BOS High B
Material
Figure 8.39 Variation of work hardening exponents (prestrain) and interstitial content for the different steel compositions tested at
130°C and 600 seconds

235
0.15 - Interstitial content
0.14 -+-Work Hardening exponent
0.360 ~
0.13 a,
~ C
0
; 0.12 a.
~
C ><
0.310 a,
o(.) 0.11 C')

ci 0.1 ·-ca,:
C:
;;
...a,
:;; 0.09 0.260 .
"C
ea
.c
~ 0.08
- 0.07 0.210 ~
~
...
0.06
0.05 -+------.---+--------+-----,-------+- 0.160
BOS Low N BOS High N EAF Low B BOS High B
Material
Figure 8.40 Variation of work hardening exponents (prestrain) and interstitial content of four different steel compositions tested
at 200°C 30 seconds.

236
0.15
- Interstitial content
0.14 -+-Work Hardening exponent
0.360 c
..., 0.13
G)
C
0
...,; 0.12 c.
C
o(.) 0.11 0.310 ><
a,
C,

ci 0.1 ·-cC
E G)
1n
... 0.09 0.260 ...ea
-c
G)
c o.oa .c
...
- 0.07 0.210 ~
~

0.06
0.05 -+------+-------+-----+------+- 0.160
BOS Low N BOS High N EAF Low B BOS High B
Material
Figure 8.41 Variation of work hardening exponents (prestrain) and interstitial content of the four steel compositions tested at
200°C and 600 seconds.

237
Chapter 8 Discussions

At the higher temperature of 200°C, the exponents have similar values at both the times

(Table 8.7). However, the exponent of the BOS high N+B grade appears to be the same

as that for the BOS low N grade. However, in the strain to failure region, there is an

appreciable difference in the exponents among the steel grades. BOS high B+N grade

shows the lowest values of the exponents at both the temperatures and times, possibly

due to the fixing of some amount of nitrogen in the steel.

In general these results agree with the results of Frank and Kirkcaldy< 73 ) who have

reported a lower work hardening rate in boron treated rods. Studies on two 0.06%

carbon EAF grade steels, one without boron and the other treated with 0.007 % boron

revealed a lower rate of work hardening below 50% reduction in area in the boron

treated steel.

Gladman et a1< 119> have derived the following equation on the influence of alloying

elements on the flow stress at room temperature.

a = 355 + 601(%C)+ 86(%Mn)+ 201(%Si)-741(%S) +880(%P)+200 (%Sn)

+114(%Cr)+4478(%N)+27.8d-112------------------------8.J

where d = grain size.

Using this equation, values of flow stress for the four grades of steel was calculated at a

true strain of 0.8. Table 8.8 compares the values obtained by calculation. Since the

actual hot torsion tests have been carried out at slightly elevated temperatures, the

values of flow stress at these temperatures, is expected to be lower. These calculations

also do not take into account the effect of Sn and the grain size, therefore adding to the

238
Chapter 8 Discussions

difference in the values of the calculated and measured flow stress. It is also noted that

the flow stress values obtained from the hot torsion data show similar results for the

BOS low N and BOS high N+ B steels, while those for the BOS high N and EAF low B

grades are about 60 MPa higher. This trend is in accordance with the results of Franks

and Kirkcaldy< 73 ) who have reported that the difference between the boron treated and

non-boron treated grades can be as high as 200 MPa. This phenomenon has been

attributed to the combination of nitrogen with boron restricting dynamic strain ageing.

However in the current investigation, the smaller difference in the flow stress values

between the boron and non-boron grades would be due to the incomplete fixing of all

nitrogen by boron as confirmed by the "free nitrogen" analysis. In this context, the

'fixing'of free nitrogen could be due to solute interaction between boron and nitrogen

in solution, which could explain the observed absence of precipitates, section 7.6.3.

Another observation to be made is the dependence of the Gladman's equation on the

grain size. As reported earlier, the grain size in the BOS boron treated high nitrogen

grade steel was found slightly higher (ASTM 6-7 as against ASTM 8 in the other

grades). This could also have resulted in a higher flow stress in the BOS boron treated

high nitrogen grade steel.

239
Table 8.8 Comparison of calculated values (using Gladman's equation) of flow stress and measured values of flow stress at a true
strain of 0.8

MATERIAL FLOW STRESS USING FLOW STRESS FROM


GLAD MAN'S HOT TORSION
EQUATION(MPa) CURVE(MPa)
BOS LOWN 474.30 496.6
BOSHIGHN 509.36 549.94
EAFLOWB 503.26 549.13
BOSHIGHN+B 503.76 487.84

240
Chapter 8 Discussions

8.3.3.2 Maximum stress

The maximum stress levels obtained at constant strain of 1.08, as shown in Figures 5.42

to 5.45 and Table 8.9, indicate high values for BOS high N and EAF low B. This is

attributable to the high interstitial content and total solute content (corroborated by high

values of work hardening exponent-Table 5.6). BOS Low N and BOS High N+B show

almost the same levels of stress, although the work hardening exponent for the former is

much higher than that for the latter (0.17 against 0.07 for BOS high N+B steel). These

results indicate the incomplete action of the boron additions in fixing all free nitrogen

and this is corroborated by the results of the free nitrogen analysis in these samples

(Table 7.2). This behaviour is exhibited at all combinations of parameters.

241
Table 8.9 Values of maximum stress obtained for the four different steel compositions under various test parameters.

MATERIAL MAXIMUM STRESS


130°C 30 130°C 600 200°c 30 200°c 600
SECONDS SECONDS SECONDS SECONDS
BOS LOWN 565.1 550.97 572.95 575.65
BOSHIGHN 590.81 582.15 . 594.87 609.75
EAFLOWB 586.48 616.25 615.44 632.22
BOS HIGH 552.86 563.74 554.21 557.45
N+B

242
600

590

580

-
ea
a. 570
-
~
( /)

-
(/)

f 560
en

550

540

530
BOS LowN BOS High N EAF LowB BOS High B
Material

Figure 8.42 Maximum stress at constant strain (1.08) for the four steels at 130°C and 30 seconds.

243
640 ~ - -- -- - - - - - - - - - -- - - - - -- ----,

620

600

-~ 580
:!!:
u,
1/)

-f! 560
( /')

540

520

500 -+-- - ' - - - - - - - ' - - ~ - _ _ . __ _...___~_ _.___ __.__---,-- ~ - - - ' - - -

BOS Low N BOS High N EAF Low B BOS High B


Material

Figure 8.43 Maximum stress at constant strain (1.08) for the four steels at 130°C and 600 seconds.

244
600 . . . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

580

560 -
cu
a.
~
in 540
Ill

-
~
en
520

500

480 --------------........-------.__-....,...._ ___._____..__---,-_ _.________


BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.44 Maximum stress at constant strain (1.08) for the four steels at 200°C and 30 seconds

245
600 - . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

580

560

i:i: 540
u,
If)

t
Cl)
520

500

480

460 • -
BOS Low N BOS High N EAF Low B BOS High B
Material

Figure 8.45 Maximum stress at constant strain (1.08) for the four steels at 200°C and 600 seconds

246
Chapter 8 Discussions

8.3.3.3 Total strain to failure

Table 8.10 shows that variation in the values of total strain to failure for the four

different compositions. BOS Low N and EAF low B steels show similar values, while

BOS high N shows a marginally lower value of strain to failure- (higher N and C

content). BOS high N+ B steel shows a distinct increase in the total strain to failure

(corroborates with the reduced values of the work hardening exponent, thus resulting in

a larger total strain to failure). This agrees with results obtained by Franks and

Kirkaldy< 73 ) who have reported a lower work hardening rate and higher drawability in

boron treated rods. A similar trend is observed at all test parameters. Although this

phenomenon has been atrributed to the fixing of nitrogen in boron<73 ), in the current

study due to clear evidence that boron has not completely fixed free nitrogen present,

this phenomenon could not be explained under the constraints of the research project.

However a slightly coarser grain size could be one of the factors influencing the total

strain to failure.

247
Table 8.10 Variation of total strain to failure as a function of steel composition.

Material Total strain to failure


130°C 30 seconds 130°C 600 seconds 200°C 30 seconds 200°C 600 seconds
delay delay delay delay
BOSLowN 1.68 1.79 1.75 1.79
BOSHi2h N 1.32 1.35 1.32 1.14
EAFLowB 1.48 1.47 1.43 1.64
BOS High N+B 2.6 2.68 2.58 ·2.ss

248
Chapter 8 Discussions

8.3.3.4 Ageing index

Table 8.11 lists the values of ageing index obtained for different test parameters.

Figures 8.46 to 8.49 indicate the variation in the ageing index and the interstitial

content for the steel compositions. The values of the ageing index for the four

compositions show similar values at 130°C and 30 seconds (variation is within

experimental error). This indicates the ineffectiveness of boron additions. This is

further confirmed by the index results exhibited by the four steels tested at the

remaining three combinations of parameters. The index values for the EAF low B

steel and the BOS high N+B steel are definitely higher than that obtained in BOS low

N. At the higher temperature of 200°C, due to dynamic strain ageing, the indices are

lower as compared to those at 130°C. However, the same trend of a marginal

difference in the index as a function of composition is maintained. BOS high N+B

steel still shows a high ageing index, indicating ineffective role of boron additions to

these steels.

249
Table 8.11 Variation in the ageing index, interstitial content (C+N) and the boron content, in the four steel compositions.

Material Interstitial Boron content Ageing index (MPa)


C+N content (wt%)
(wt%)
130°c 30 130°c 600 200°c 30 200°c 600
seconds seconds seconds seconds
BOS LOWN 0.072 0 71.45 67.66 24.63 20.3
BOSHIGHN 0.088 0 83.09 85.52 38.16 31.8
EAFLOWB 0.058 0.0012 85.18 149.84 21.38 26.53
BOS HIGH 0.086 0.0038 69.88 116.41 54.23 84.18
N+B

250
0.12 190

0.11
-interstitial content I
-+- Ageing index
+170
I
150

-f
C
0.1

0.09
130 ><
G)
0
u 110 ~
ea o.os
.::;
C,
C
.::; 90 G)

--
fG) 0.07 C,
<(
C 70
0.06 50
0.05 30

0.04 10
BOS LowN BOS High N EAF LowB BOS High B
Material

Figure 8.46 Variation of interstitial content and ageing index for the four steel compositions tested at 130°C and 30 seconds.

251
0.12 190

..,
0.11
0.1
t
I
~
- Interstitial content
I -+-Ageing index t170
150
C
..,Cl) ><
130 Cl)
5 0.09 "CS
CJ 110 .5
ea o.os C)
~
~
90 ·-C
Cl)
fCl) 0.07 C)
.., 70 <C
C 0.06 • 50
0.05 30
0.04 10
BOS LowN BOS High N EAF LowB BOS High B
Material
Figure 8.47 Variation of interstitial content and ageing index for the four steel compositions tested at 130°C and 600 seconds

252
0.12 190
- Interstitial content I
0.11 + I -+-Ageing index
+ 170
I

.... 150
C 0.1
....
Cl)
C
><
130 Cl)
"CS
8 0.09 110 .5
-CU C,

~ 0.08 90 ·-
C
Cl)
C,
fCl) 70 <(
c: 0.07 '
I
- 50
0.06 + _______.____ v~ I 30
0.05 10
BOS LowN BOS High N EAF LowB BOS High B
Material

Figure 8.48 Variation of interstitial content and ageing index for the four steel compositions tested at 200°C and 30 seconds.

253
0.12 190
- interstitial content 1 + 170
0.11 t I
-+- A gemg
. .m d ex
..., 0.1 I
+ 150
C
...,Q) ><
130 Cl)
§ 0.09 "C
c., 110 .5
ea o.oa C)
+i
+i 90 ·-C
/
Cl)
r? 0.07 C)
...,Q) 70 <(
C 0.06 50
0.05 30

0.04 10
BOS LowN BOS High N EAF LowB BOS High B
Material
Figure 8.49 Variation in the interstitial content and ageing index for the four steel compositions tested at 200°C and 600 seconds.

254
Chapter 8 Discussions

8.4 TRANSMISSION ELECTRON MICROSCOPIC EXAMINATION:

Extensive examination under the TEM did not reveal the presence of any BN

precipitates in any of the steel grades under the current experimental conditions.

This would mean that the kinetic factors of temperature and time combinations

were not favourable for BN precipitation in these compositions. Experiments by

Watanabe and Ohtani(S 9) on steels containing 0.0096% boron, 0.0326%nitrogen

and 0.001 % aluminium (BIN ratio= 0.29) revealed BN precipitates when cooled

from a hot rolling temperature of 1300°C at the rate of 300°C /minute. A slower

cooling rate of 0.67°C/minute gave a mix of BN and AlN precipitates. The

cooling rate employed in the current study is of the order of water quenching

followed by a cooling rate of 2°C/sec or 120°C/minute from 900°C, which is in

between the two cooling rates stated in Watanabe's study. For BN to precipitate,

it should do so, either, during watet quenching or when cooled from 900°C at a

slower rate. However, it must be noted that the austenitising temperature used in

Watanabe's study was 1300°C as against 1100°C followed by water quenching to

900°C in the current study from where a cooling rate of 2°C/sec was employed.

More importantly, a boron content of 0.0096% was used by Watanabe, against a

maximum of 0.0038% used in the current study. No precipitation in the matrix

or on the dislocations was noticed., leading to the conclusion that the strain ageing

phenomenon in these steels was due to atmosphere formation around dislocations.

The second stage of precipitation on dislocations was not reached even though

enough number of sites were available for heterogenous nucleation and sufficient

driving force in the form of solute available in excess of the equilibrium solubility

at the lower drawing temperatures of 130°C and 200°C, was available.

255
Chapter 9 Conclusions

CHAPTER9

CONCLUSIONS

256
Chapter 9 Conclusions

This research project has generated varied data on the phenomenon of strain ageing in

low carbon steel wire rods supplied by BHP, Newcastle. The analysis of data has led to

several important conclusions as stated below. It must however be mentioned that the

project has been formulated and researched under several constraints, namely limited

availability of industrial and experimental steel grades and relatively high testing costs.

The conclusions are as follows:

9.1 USE OF HOT TORSION TESTING AS A SIMULATION OF THE WIRE DRAWING

PROCESS.

As mentioned in Chapters 2 and 3, hot torsion has been used successfully by several

researchers to simulate the wire drawing processes. Extensive hot torsion testing on the

BHP-developed hot torsion rig and subsequent analysis of data during the current

research project has confirmed the consistency and suitability of the test method for

such studies.

9.2 INFLUENCE OF VARIOUS PARAMETERS ON THE STRAIN AGEING BEHAVIOUR:

9.2.1 Influence of temperature:

• Work hardening rate

The work hardening exponents (in the prestrain region) at the testing temperature of

200°C are higher as compared to the exponent values at the lower temperature of 130°C.

This is attributed to the occurrence of dynamic strain ageing at higher temperatures

where both interstitial carbon and nitrogen have high diffusion coefficients sufficiently

high in order to pin the dislocations dynamically during straining itself. The diffusion

257
Chapter 9 Conclusions

coefficient of nitrogen and carbon at 200°C is approximately 125 times higher than at

the lower temperature (De 20ocl De 130c = 125). This trend is generally observed in all

the four steel compositions. The work hardening exponents in the strain to failure

region (0.005 - 0.173) of the stress-strain curve are generally lower than the prestrain

values(0.173 - 0.354), indicative of the onset of plastic strains in the strain to failure

region. At a higher temperature of 200°C, the work hardening rates (0.295 - 0.354 in

the prestrain region) show generally higher values as compared to those obtained at

130°C (0.173 - 0.303 in the prestrain region). This is a result of increased interstitial

diffusion rates at 200°C leading to a higher dynamic strain ageing.

• Maximum Stress

The levels of the maximum stress show a very distinct increase at 200°C (a maximum of

25 MPa) as compared to those obtained at 130°C in both the prestrain and the strain to

failure regions of the stress-strain curves. This is attributed to increased dynamic strain

effects in both regions of the stress-strain curve, specifically higher during the prestrain

region. At the higher temperature, there is increased diffusion of the interstitials to the

dislocations causing an increase in the dynamic strain ageing effects. In the strain to

failure region, due to the onset of plastic deformation, the rise in the level of the stress-

strain curve is lower.

• Total strain to failure

Testing temperature does not have a significant effect on the total strain to failure. This

could be possibly due to the smaller difference between the drawing temperatures and

the fact that the two test temperatures are generally in the cold working range for all the

material compositions.

258
Chapter 9 Conclusions

• Ageing index

The ageing index is observed to be higher at 130°C than at 200°C. This is due to the

occurrence of dynamic strain ageing at 200°C, which raises the level of the stress -

strain curve in the prestrain region. This in tum results in a lower ageing index. This

behaviour is observed in results obtained at both 30 and 600 seconds delay time after

prestrain.

9.2.2 Influence of delay time between prestrain and strain to failure:

• Work hardening rates

There seems to be no significant change in the value of the exponent with an increase

in the delay time after prestrain. This is possibly due to the fact that all interstitials

have migrated during prestrain and the first 30 seconds of holding time. No further

diffusion of interstitials has taken place beyond 30 seconds, thus leading to a negligible

effect of holding time on dynamic strain ageing and therefore, on the work hardening

exponent.

• Maximum stress

There appears to be no significant difference in the maximum stress values obtained in

the strain to failure region as a function of delay time. This leads to conclusion that all

the solute has segregated to the dislocations within the first 30 seconds and no further

addition is made to the amount of solute segregated at times greater than 30 seconds.

• Total strain to failure

The delay time between prestrain and strain to failure does not affect the total strain to

failure. This phenomenon is observed at both the testing temperatures.

259
Chapter 9 Conclusions

• Ageing index

A significant difference in the index is observed in EAF low B (149MPa for 600

seconds delay as against 89 MPa for the 30 seconds delay) and BOS high N+B(l 16

MPa at 600 seconds delay as against 70 MPa for the 30 seconds delay) with an increase

in the delay time after prestrain. However there is no significant difference in the index

with an increase in the delay time in BOS low and high N steels. It must also be

mentioned that it may be possible that the delay time of 600 seconds may not be enough

for interstitial diffusion at the low temperatures

9.2.3 Influence of material composition:

• Work hardening rates

Among the four grades of steels, BOS high B+N shows a marginally lower value of

work hardening rate in the prestrain region (0.228 as compared to 0.284 (BOS low N);

0.303 (BOS high N) and 0.244 in the EAF N+B grade at 130°C and 30 seconds; 0.295

as compared to 0.298 (BOS low N); 0.312 (BOS high N) and 0.354 in the EAF N+B

grade at 200°C and 30 seconds). This agrees with the results of Franks and Kirkcaldy< 73 >

who have reported lower work hardening rates in boron steels. The reasons for this

behaviour are the fixing of some of the free N by boron (confirmed by 'free nitrogen

analysis) and a coarser grain size. The values of the work hardening exponent are

definitely low for BOS high B+N grade in the strain to failure region (0.068 as

compared to 0.173 (BOS low N); 0.145 (BOS high N) and 0.153 in the EAF N+B grade

at 200°C and 30 seconds; and 0.028 as compared to 0.118 (BOS low N); 0.127 (BOS

high N) and 0.116 in the EAF N+B grade at 200°C and 30 seconds).

• Maximum stress

260
Chapter 9 Conclusions

The maximum stress levels in the BOS high B +N steel are comparable to the BOS Low

N steel, although the work hardening rates are generally much lower in the former

material. The elongation in the BOS high B+N steel is also much higher as compared to

the BOS low N grade. The effect on maximum stress levels coupled with the high total

strain to failure in the high boron grade leads to a conclusion that some solid solution

strengthening is occurring in the BOS high N steel grade. A comparison of the

compositions indicates an appreciable difference in the boron levels between the two

grades, leading to a possibility of boron strengthening ferrite, having not fully

precipitated as BN. The other elements which could lead to ferrite strengthening are

almost the same level in the two steels.

• Total strain to failure

The value of the total strain to failure for the BOS low N is slightly higher (0.3 units

higher) as compared to the BOS high N and the EAF low B grades. This is expected

due to the higher interstitial content in the latter grades, although it is to be noted that

the work hardening rates are similar or even slightly higher for the BOS low N grade.

The BOS High B+N grade exhibits the highest value of strain. As reported in literature

this is expected to be due to the fixing of all free nitrogen by boron. However, in the

current study it has been observed that all free nitrogen has not been fixed by boron. It

is therefore inferred that the high value of total strain to failure is due to a combination

of the fixation of some free nitrogen as well as a coarser grain size, resulting in higher

drawability.

• Ageing index

The values of ageing index for the four compositions show similar values at l 30°C and

30 seconds (variation is within experimental error). This indicates that boron additions

261
Chapter 9 Conclusions

to the BOS and the EAF grade steels have not resulted in an additional advantage as far

as reducing strain ageing is concerned. This is further confirmed by the ageing index

results exhibited by the four steels tested at the remaining three combinations of

parameters, namely, 130°C-600 seconds, 200°C-30 seconds and 200°C- 600 seconds.

The index values for the EAF low boron steel and the BOS high N+B steel are

definitely higher than that obtained for the BOS low N steel. At the higher temperature

of 200°C, due to dynamic strain ageing, the indices are lower as compared to those at

130°C. However, the same trend of a marginal difference in the index as a function of

composition is maintained. BOS high N+B still shows a high ageing index indicating

incomplete effects of boron additions to these steels.

From the results of hot torsion and free nitrogen analysis studies, it has been observed

that BN has not completely precipitated in either the low boron steel (0.0012%boron

and 0.008% nitrogen, 0.003%aluminium, BIN ratio=0.15) or the high boron steel

(0.0038% boron, 0.006% nitrogen,. 0.006%aluminium, BIN = 0.63). However, both

these steel compositions have exceeded the equilibrium solubility limit of BN in

austenite. Absence of precipitation under these conditions could only mean that the

kinetic factors of temperature and time combinations were not favourable for BN

precipitation in these compositions. It has been reported that during the current

investigation, rods were quenched to the rod rolling temperature (from 1100°C to about

900°C) at BHP Newcastle. Experiments by Watanabe and Ohtani<S9) on steels

containing 0.0096% boron, 0.0326%nitrogen and 0.001% aluminium (BIN ratio= 0.29)

revealed BN precipitates when cooled from a hot rolling temperature of 1300°C at the

rate of 300°C /minute. A slower cooling rate of 0.67°C/minute gave a mix of BN and

262
Chapter 9 Conclusions

AlN precipitates. The cooling rate employed in the current study is of the order of

water quenching followed by a cooling rate of 2°C/sec or 120°C/minute from 900°C,

which is in between the two cooling rates stated in Watanabe's study. For BN to

precipitate, it should do so either, during water quenching or when cooled from 900°C at

a slower rate. However, it must be also be stated that the austenitising temperature used

in Watanabe's study was 1300°C as against 1100 °C from where it was water quenched

and then 900°C in the current study from where a cooling rate of 2°C/sec was employed.

More importantly, a boron content of 0.0096% was used by Watanabe, as against a

maximum of 0.0038% used in the current study.

Besides this, another observation made is that due to a high degree of working of the

materials at 900°C, it is expected that enough number of dislocation sites are available

for heterogenous nucleation of BN in the boron containing grades. However, it has

been confirmed by 'free nitrogen' analysis results that a very small quantity of boron

has fixed nitrogen. This is further corroborated by the results of ageing indices as

discussed in earlier Chapters 7 and 8, where the BOS high N+B grade does not show a

marked and consistent improvement in the value of the ageing index as expected. In

view of the above arguments, it can be concluded that boron additions of 0.0012wt%

and 0.0038wt% to steels of 0.006-0.008 wt% N have not been effective in fixing all free

nitrogen under the cooling conditions of water quenching from 1100°C to 900°C and

subsequent cooling to room temperature at a cooling rate of 2°C/s.

This definitely leads to the conclusion that although thermodynamically, BN has

exceeded the equilibrium solubility· limits at the austenitising temperature, the driving

force due to supersaturation is not sufficient to cause precipitation during quenching

263
Chapter 9 Conclusions

from this temperature. Further this supersaturation is not sufficient to facilitate

heterogenous BN precipitation during rolling.

It is also important to note that extensive transmission electron microscopic examination

of the four grades of steel tested at various combinations of time and temperature, did

not reveal any type of precipitation in the matrix or on the dislocations. This leads to the

conclusion that the strain ageing phenomenon in these steels was due to atmosphere

formation around dislocations. The second stage of precipitation on dislocations was

not reached even though enough number of sites were available for heterogenous

nucleation and sufficient driving force in the form of solute available in excess of the

equilibrium solubility at the lower drawing temperatures of 130°C and 200°C, was

available.

264
References

REFERENCES

265
References

1. G.E. Dieter, "Mechanical Metallurgy", third edition, 1986, published by McGraw


-Hill.

2. R. Louis, "Factors influencing the drawability of wire rod - A Literature review",


Technical report BHPRIOE/N/005.

3. W.C Leslie, "Iron and its dilute substitutional solid solutions", Metallurgical
Transactions, vol. 3, January 1972, p 5 - 26.

4. B.B Hundy, "The strain age hardening of Mild steel", Metallurgia, May 1956, p
203 - 211.

5. A.H Cottrell and G.M Leak, "Effect of quench aging on strain aging in Iron",
Journal of the Iron and Steel Institute, November 1952, p 301 -306.

6. B.B Hundy, "The repeated strain aging of Mild steel", Metallurgia, January
1957, p 27 - 30.

7. G. Krauss, "Principles of Heat treatment of Steels", ASM publication, Metals


Park, Ohio, 1980.

8. D.S Clark and W.R Varney, "Physical Metallurgy for Engineers", Second edition,
1962, published by Litton Educational Publishing, INC.

9. D.A Porter and K.E Easterling, "Phase transformations in metals and alloys,
1988, published by Van Nostrand Reinhold (International) Co. Ltd.

10. R.W.K Honeycombe and H.K.D.H Bhadeshia, "Steels- Microstructure and


Properties", published by Edward Arnold, London, 1995.

11. K.E Thelnig, "Steel and its heat treatment", second edition, published by Boston
Buttersworths, London, 1984.

12. U. F Kocks, "Kinetics of solution hardening", Metallurgical Transactions,


volume 16A, December 1985, p 2109 - 2129.

13. E.C Bain and H.W Paxton, "Alloying elements in Steel", ASM Publication,
second edition, ASM publication, Metals Park, Ohio, 1961.

14. W.C Leslie, "Nitrogen in ferritic steels - a critical survey of the literature",
American Iron and Steel Institute, 1959.

15. P. Shewmon, "Diffusion in solids", second edition, published by The Minerals,


Metals and Materials Society, 1989.

16. H.I Aaronson, Jong K. Lee and K.C Russell, "Diffusional nucleation and
growth", Precipitation processes in solids, Proceedings 1976 TMS Fall Meeting,
New York, September 20- 21, p 31- 86.

266
References

17. R.Simoneau, G.Begin and A.H Marquis, "Progress of NbCN precipitation in


HSLA steels as determined by electrical resistivity measurements", Metal
Science, August 1978, p 381- 386.

18. A.H Cottrell and A.T Churchman, "Change of electrical resistance during the
strain aging of Iron", Journal of the Iron and Steel Institute, July 1949, p 271-
276.

19. R.McCallum, J.J Jonas and Steve Yue, "The drawability of Low carbon steel
wire", Wire Journal International, p 102- 107.

20. A.S Keh and H.A Weirdt, "An electron transmission study of nitride precipitation
in alpha iron", Transactions of the Metallurgical Society of AIME, June 1962,
volume 224, p 560 - 572.

21. W.C Leslie and A.S Keh, "An electron transmission study of the strain aging in
Iron", Journal Iron and Steel Institute, 1962, volume 200, p 722- 728

22. H.H Willard, L.M Merritt and J.A Dean, "Instrumental methods of analysis",
fourth edition, published by D. Van Nostrand Company, Inc, London, August
1965.

23. J .J Jonas and I. Weiss, "Effect of precipitation on recrystallisation in microalloyed


steels", Metal Science, March - April 1979, p 238 - 245.

24. M.G Akben, T.Chandra, P.Plassiard and J.J Jonas, "Dynamic precipitation and
solute hardening in a titanium microalloyed steel containing three levels of
manganese", Acta Metallurgica, 1984, volume 34, No. 4, p 591 - 601.

25. I. Weiss and J .J Jonas, Interaction between recrystallisation and precipitation


during high temperature deformation of HSLA steels", Metallurgical
Transactions, July 1979, volume lOA, p 831 - 840.

26. T. Chandra, I. Weiss and J .J Jonas, "Effect of static recrystallisation of the growth
ofNbCN in a 0.07% Nb HSLA steel", Canadian Metallurgical Quarterly, 1981,
volume 20, No. 4, p 421 - 428.

27. J.P Michel and J.J Jonas, "Precipitation kinetics and solute strengthening in high
temperature austenites containing Al and N", Acta Metallurgica, volume 29, p
513 - 526.

28. T.Chandra, I.Weiss and J.J Jonas, "Influence of dynamic recovery and dynamic
recrystallisation on coarsening of NbCN in Nb microalloyed steel", Metal
Science, February 1982, volume 16, p 97 -104.

29. J.D Baird, "Strain aging of steel - A critical review, Part I, Practical aspects",
Iron and Steel, May 1963, p 186 - 192.

30. J .D Baird, "Strain aging of steel - A critical review, Part I, Practical aspects", Iron
and Steel, June 1963, p 326 - 334.

267
References

31. B.B Hundy, "Accelerated strain aging of Mild steel", Journal of the Iron and
Steel Institute, September 1954, p 34 - 38.

32. W.R Thomas, "The strain aging of Alpha iron", Journal of the Iron and Steel
Institute, June 1955, p 155 - 161.

33. J .D Baird, "The effects of strain aging due to interstitial solutes on the mechanical
properties of metals", Metallurgical reviews, Review 149, 1971, p 1 -18.

34. J.D Fast and M.B Verrijp, "Diffusion of nitrogen in Iron", Journal of the Iron
and Steel Institute, January 1954, p 24 - 28.

35. L.P Kubin, K.Chihab and Y.Estrin, "The rate dependence of the Portevin-
LeChatelier effect", Acta Metallurgica, 1988, volume 36, No. 10, p 2707 -
2718.

36. H.M Zbib and E. C Aifantis, "A gradient dependent model for the Portevin - Le
Chatelier effect", Scripta Metallurgica, 1988, volume 22, p 1331 - 1336.

37. G. Schoeck, "Moving dislocations and solute atoms", Physical Review, June
1956, volume 102, number 6, p 1458-1459.

38. J.D Baird, "Strain aging of steel - A critical review, Part III, Dynamic strain
aging", Iron and Steel, September 1963, g 450 - 457.

39. Shukui Lou and Derek 0. Northwood, "Effect of carbon and nitrogen content on
dynamic strain aging in Low carbon steels", Canadian Metallurgy Quarterly,
1994, volume 33, No. 3, p 243 - 24.

40. W.R Clough, L.A Jackman and Y.G Andreev, "Dynamic strain aging and the
Charpy specimen behaviour of annealed 4340 steel", Journal of Basic
Engineering, Transactions ofASME, March 1968, p 13 - 20.

41. A.V Loginov, A.E Trotsenko, M.Mastakhov and L.P Loshamanov, "Dynamic
strain aging of steels in a wide range of deformation rates", Physics of Metals
and Metallography, volume 68, No. 4, p 7 - 10.

42. J.Schlipf, "On the kinetic~ of static and dynamic strain aging", Scripta
Metallurgica et Materialia, volume 31, No. 7, p 909 - 914.

43. G.H Rubiolo and P.B Bozzano, "Dynamic Interaction of impurity atmospheres
with moving dislocations during stress relaxation", Materials Transactions JIM,
volume 36, No. 9 (1995), p 1124 - 1133.

44. J.M Robinson and M.P Shaw, "Microstructural and mechanical influences on
dynamic strain aging phenomena", International Materials reviews, 1994,
volume 39, No. 3, p 113 - 122.

268
References

45. J.A Yaker, C.C Li and WC Leslie, "The effects of dynamic strain aging on the
mechanical properties of several HSLA steels", SAE, International Congress on
Automotive Engineering, February 26 - March 2, 1979, Cobo Hall, Detroit.

46. A. Van Den Beukel and U .F Kocks, "The strain dependence of static and dynamic
strain aging", Acta Metallurgica, 1982, volume 30, p 1027 - 1034.

47. A.Van Den Beukel, "On the mechanism of serrated yielding and dynamic strain
aging", Acta Metallurgica, 1980, p 965 - 969.

48. A.K Sachdev, "Dynamic strain aging of various steels", Metallurgical


Transactions, 1982, volume 13A, p 1793 - 1797.

49. C. Weidig, M. Espindola, B. M. Gonzalez, P. Rodrigues and M. Andrade,


"Dynamic strain ageing in low carbon steel wire rods", Wire Journal
International, January 1995, p 82-85.

50. J.D Baird, "Strain aging of steel - A critical review, Part II, Theory of strain
aging", Iron and Steel, July 1963, p 368 - 374.

51. J.D Baird, "Strain aging of steel - A critical review, Part II, Theory of strain
aging", Iron and Steel, August 1963, p 400 - 405.

52. A.H Cottrell and B.A Bilby, "Dislocation theory of yielding and strain aging of
Iron", Proceedings. Physics Society, 1949, [A], 62, p 49.

53. S.Harper, Physics Review, 1951, volume 83, p 709.

54. W.S Carswell, "Investigation of strain aging of quench aged alpha Iron
containing nitrogen, using internal friction technique", Acta Metallurgica, July
1961, volume 9, p 670 - 677.

55. T.Mura, E.A Lautenschlager and J.O Brittain, "Segregation of solute atoms
during strain aging", Acta Metallurgica, May 1961,volume 9, p 453 - 458.

56. D.V Wilson and B. Russell, "The contribution of atmosphere locking to the strain
aging of Low carbon steels", Acta Metallurgica, January 1960, volume 8, p 36
- 45.

57. D.V Wilson and B. Russell, "The contribution of precipitation to strain aging in
Low carbon steels", Acta Metallurgica, July 1960, volume 8, p 468 - 479

58. A.S Keh and W.C Leslie, 'Materials Science Research (edited by H.H Stadelmeier
and W.W Austin), 1963, volume 1, p 208, New York (Plenum press).

59. W.C Leslie, "Quench aging of low carbon and Fe-Mn alloys - an electron
transmission study", Acta Metallurgica, 1961, volume 9, p 1004.

269
References

60. A.H Cottrell, NPL Conference on "Relations between the structure and
mechanical properties of metals", published by HM Stationery Office, 1963, p
455.

61. S.J.Kennett and W.S Owen, Iron and Steel Institute, preprints for Conference
on 'Recent developments in annealing.,1963.

62. T.Mura and J.O Brittain, Acta Metallurgica, May 1960,volume 8, p 767.

63. W.C Leslie and A.S Keh, Journal Iron and Steel Institute, 1962, volume 200, p
722.

64. W.A Owen, M.J Roberts, 'Dislocation Dynamics", (edited by A.R Rosenfield et
al) 1968, p 357, New York and London (McGraw Hill).

65. Y.Nakada and A.S Keh, "Kinetics of Snoek ordering and Cottrell atmosphere
formation in Fe- N single crystals",' Acta Metallurgica, 1959, p 879- 883.

66. D.T Llewellyn, "Nitrogen in steels", Ironmaking and Steelmaking, 1993,


volume 20, No.I, p 35 - 41.

67. G.M. Faulring, "Nitrogen scavenging with boron", 1989, Electric Arc Furnace
Proceedings, p 155-161.

68. L.J Dijkstra and R.J Sladek, Transactions A/ME, 1953, volume 197, p 69.

69. D.A Leak, W.M Thomas and G.M Leak, "Diffusion and solubility of nitrogen in
silicon-iron", Acta Metallurgica, September 1955, volume 3, p 501.

70. W.M Thomas and G.M Leak, Proc. Phy. Society, 1955, B 68, p 1001.

71. A.Josefsson and B.Backstrom, Iron and Steel, 1966, volume 39, p 43.

72. A. Karimi Taheri, T.M Maccagno and J.J Jonas, "Effect of cooling rate after Hot
rolling and of multistage strain aging on the drawability of Low carbon steel wire
rod", Metallurgical and Materials Transactions, May 1995, volume 26A, p
1183 -1193.

73. J. F Butler, "The effect of heat treatment and microstructure on carbon strain
aging in Low carbon steels", Transactions of the Metallurgical Society of
A/ME, February 1962, volume 224, p 89 - 96.

74. E.T Stephenson, "Factors affecting the strain aging susceptibility of box annealed
Low carbon steel", Transactions of the ASM, 1969, volume 62, p 490 - 501.

75. S.Bergh, Jernkontorets Ann., 1966, volume 150, p 249.

76. N.P Allen, W.P Rees, B.E Hopkins and H.R Tipler, "Tensile and impact
properties of high purity Fe-C and Fe-C-Mn alloys of low carbon content",
Journal Iron and Steel Institute, 1953, volume 174, p 108.

270
References

77. E.T Stephenson, "Effect of carbon in solution on the properties of an Aluminium


killed Low carbon steel", Transactions of the ASM, 1962, volume 55, p 624 -
639.

78. W.C Leslie and R.L Rickett, "Influence of Al and Si deoxidation on strain aging
of low carbon steels", Transactions AIME, 1953, volume 197, p 1021-1031,
Journal of Metals, 1953.

79. F.Laxar, J.W Frame and D.J Blickwede, "Effect of Al on strain aging and internal
friction in low carbon steel", Transactions ASM, 1961, volume 53, p 683.

80. F .D Richardson, "Thermodynamics of metallurgical carbides and of carbon in


iron", Journal Iron and Steel Institute, September 1953, volume 175, p 33- 51.

81. P. Messien and V. Leroy, "Solute carbon and nitrogen in the ferrite after
continuous annealing of low carbon low aluminium steels", Materials
Technology steel research 56, 1985, No.7, p 385-391.

82. M. E Nicholson, "Constitution of iron-boron alloys in the low boron range",


Journal of Metals, Transactions ofA/ME, February 1954, p 185 - 190 .

83. P.E Busby and C. Wells, "Diffusion of boron in alpha iron", Journal of Metals,
Transactions ofA/ME, September 1954, p 972.

84. C. C. McBride, J. W. Spretnak and R. Speiser, "A study of the Fe-Fe2B system",
Transactions of the ASM, Volume 46, p 499 - 524.

85. R.W Fountain and J. Chipman, "Solubility and precipitation of boron nitride in
iron-boron alloys", Transactions of the Metallurgical Society of A/ME, volume
224, June 1962, p 599-605.

86. A. Franks and A. Kirkcaldy, "The effect of boron on the properties of electric arc-
sourced plain carbon wiredrawing qualities", Wire Journal International, May
1998, p 100 -113.

87. G.F Melloy, P.R Simmon and P.P Podgursky, "Optimising the boron effect",
Metallurgical Transactions, volume 4, October 1973, p 2279-2289.

88. S. Watanabe, H. Ohtani and T. Kunitake, "The influence of Hot rolling and heat
treatments on the distribution of boron in steel", Transactions /SIJ, volume 23,
1983, p 31 - 37.

89. S. Watanabe, H. Ohtani, "Precipitation behaviour of boron in high strength


steel", Transactions ISIJ, volume 23, 1983, p 38 - 42.

90. G. Derge, "The Boron-Oxygen equilibrium in liquid iron", Transactions AIME,


1946, volume 167, p 93 - 110; Metals Technology, August 1946.

91. P. Messien and V. Leroy, "Scavenging additions of boron in low C- low Al


steels", Materials Technology steel research 60, 1989, No.7, p 320 -328.

271
References

92. I.D Mclvor, "Microalloyed very low carbon steel rod", Ironmaking and
Steelmaking, 1989, volume 16, No. 1, p 55 - 62.

93. E.C Morgan and J.C Shyne, "The strain aging of boron treated low carbon steels",
Journal Iron and Steel Institute, February 1957, p 156 -160.

94. E.R Morgan and J.C Shyne, "Preparation and properties of boron treated non
aging open hearth steel", Transactions AIME, 1957, volume 209, p 781.

95. K.Yamanaka and Y. Ohmori, "Effect of boron on the transformation of low


carbon low alloy steels", Transactions ISIJ, volume 17, 1977 (93).

96. C.A Edwards, D.L Phillips and H.N Jones, Journal Iron and Steel Institute, 1940,
volume 142, p 199.

97. W.R.D Jones and G.Coombs, "Strain aging of mild steel- effect of vanadium and
chromium strain aging of rimming steels", Journal Iron and Steel Institute,
1953, volume 174, p 9-15.

98. S.Epstein, H.J Cutler and J.W Frame, "Vanadium treated non aging rimming steel
for deep drawing quality sheet", Journal Mater, 1968, volume 3, p 559.

99. M.S Rashid, "Strain aging kinetics of Vanadium or Titanium strengthened high
strength low alloy steels", Metallurgical Transactions, 1976, volume 7A, p 497-
503.

100. M.G Akben, I Weiss and J.J Jonas, "Dynamic precipitation and solute hardening
in a vanadium microalloyed steel and two niobium steels containing high levels
ofmanganese",Acta Metallurgica, volume 29, p 111-121.

101. L.A Erasmus and L.N Pussegoda, "The strain aging characteristics of reinforcing
steel with a range of vanadium contents", Metallurgical Transactions, February
1980, volume llA, p 231 - 237.

102. J.S Smail, S.R Keown and LA.Erasmus, "Effect of titanium additions on the
strain aging characteristics and mechanical properties of carbon-manganese
reinforcing steels", Metals Technology, April 1976, p 194 - 201.

103. R. Doremus and E.F Koch, "The precipitation of carbon from alpha iron I.
Electron microscopic study", Transactions, AIME, August 1960, volume 218, p
591 - 596.

104. R. Doremus and E.F Koch "The precipitation of carbon from alpha iron II.
Kinetics", Transactions, AIME, August 1960, volume 218, p 596 - 605.

105. I. Ochai, H. Ohba and A. Kawana, "Effect of titanium addition on strain ageing of
low carbon steel wire rod", Wire Journal International, December 1994, p 72-
81.

272
References

106. I. Ochai, H. Ohba and A. Kawana, "Effect of titanium addition on strain ageing of
low carbon steel wire rod", Wire Journal International, December 1994, p 82-
83.

107. D.V Wilson, "Grain size dependence of discontinuous yielding in strain aged
steels", Acta Metallurgica, 1968, volume 16, p 743-753.

108. D.V Wilson, "Effects of plastic deformation on carbide precipitation in steel",


Acta Metallurgica, 1957, volume 5, p 293.

109. H. P Tardif, "The effect of temper rolling on the strain aging of Low carbon
steel", Journal of the Iron and Steel Institute, January 1956, p 9 - 19.

110. D.V.Wilson and G.R Ogram, Journal Iron and Steel Institute, volume 206, p
911.

111. J.O Britain and S.E. Bronisz, "Stress induced strain aging", Transactions AIME,
1960, volume 218, p 289.

112. A.H Cottrell, "Theory of brittle fracture in steel and similar metals",
Transactions of The Metallurgical Society ofAIME, April 1958, p 192 - 203.

113. Eric R. Morgan and J.C Shyne, "Control of strain aging in Alpha iron",
Transactions AIME, Journal of Metals, January 1957, p 65 - 69.

114. J.D Baird and A.Jameson, NPL Symposium on "The relation between structure
and mechanical properties of Metals", 1963.

115. E.Pink and S.Kumar, "Patterns of serrated flow in a low carbon steel", Materials
Science Engineering A, 1995, volume 202, p 58 - 64.

116. L.J Cuddy and W.C Leslie, "Some aspects of serrated yielding in substitutional
solid solutions of Iron", Acta Metallurgica, October 1972, volume 20, p 1157-
1167.

117. R.Louis, P.D Hodgson and K.Little, "Rod drawability - Part 3, A new technique
to simulate wire production", BHPRIOEIN/034.

118. R.Louis and K.Little, "Rod drawability - Part 4, Simulation of wire drawing of
low carbon steels", BHPRIOEIN/033.

119. R.Louis and K.Little, "Part 4, Simulation of wire drawing of low carbon steels",
BHPRIOE/N/048

120. S.M Stevens, "Derivation of a procedure for determining free nitrogen in ferritic
steels", TWI Report 406/1989.

121. L.Roesch, "Effect of strain aging on wire drawing and stress relaxation behaviour
of carbon steels",

273
References

122. K. Aoki, S.Sekino and T.Fujishima, "On the quantitative measurement of


interstitially dissolved N and C in commercial steel by the internal friction
method", Transactions, Japan Institute of Metals, 1963, volume 4, p 151.

123. P.N Richards and K.V Barratt, "Dissolved carbon and nitrogen at low
concentrations in commercial steel determined by the internal friction", Journal
of the Iron and Steel Institute, April 1966, p 380 - 384 ..

124. T.Gladman, B. Holmes and F.B Pickering, "Work hardening of Low carbon
steels", Journal of the Iron and Steel Institute, February 1970, p 172 -183.

274
AppendixA

APPENDIX A

275
AppendixA

500
450
400

-
~ 300
350

-
:il:
tn
tn
250
i
en
200
150
100
50
0
0 0.2 0.4 0.6 0.8 1 1.2
Strain

Figure A-1 Hot tensile curves ofEAF Low B steel tested at 130°C after a delay of 600 seconds.

276
AppendixA

500
450
400
350
-
~
::E
300
";; 250
U)
Cl) 0.27 prestrain
b 200
en
150
100
50
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Strain

Figure A-2 Hot tensile curves of EAF low B steel tested at 130°C after a delay of 30 minutes.

277
AppendixA

500
450
400
0.18 prestrain
350
-
~ 300
:E
-; 250
u,
a,
.b 200
u,
150
100
50
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Strain

Figure A-3 Hot tensile curves of BOS high N tested at 130°C after a delay of 30 minutes.
278
AppendixB

APPENDIXB

HOT TORSION - TRUE STRESS STRAIN

CURVES

279
Appendix B

The true stress-strain plots for all the four grades of steel tested under varied

combinations of temperature and delay time are included in this Appendix B. The

various parts of the curve are explained in the first figure, Figure B-1. The initial part of

the curve labelled as A is the prestrain and the subsequent part is the stress-strain

behaviour when tested to failure after holding for a period of time following initial

prestrain for a period of either 30 seconds or 600 seconds. A point to note is that the

prestrain is consistent in all the stress-strain plots, indicative of the consistency and

reproducibility of the hot torsion test rig.

280
700

600

500

f:!!: 400
-U)
U)

"'-
l!! 300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-1 Stress-strain curve of BOS Low N tested at 130°C for a delay time of 30 seconds after prestrain.

281
700

600

500

-
~ 400
-
:::i:
I I)

-
II)

! 300
U)

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-2 Stress strain curve of BOS Low N steel tested at 130°C after a delay time of 600 seconds after prestrain

282
700

600

500

-
C,_ 400
-
:!!:
1/)
1/)

....f
u,
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-3 Stress strain curve for a BOS Low N steel tested at 200°C and 30 seconds after prestrain.

283
700

600

500

ci
400
-
D.
:iii:
I ll
Ill

- f 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-4 Stress strain curve of a BOS Low N steel tested at 200°C after a delay time of 600 seconds after prestrain.

284
700

600

500

-
CU
400
-e
D.
:I:
t /)
t/)

.... 300
u,

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-5 Stress strain curve of a BOS High N steel tested at 130°C after a time delay of 30 seconds after
prestrain

285
700 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

600

500

l:::.ii: 400
'iii'
Ill
i 300
"'
200

100

0-------------.----------,---------,--------,-------
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-6 Stress strain curve of a BOS high N steel tested at 130°C after a time delay of 600 seconds after prestrain.

286
700 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

600

500 ·

"i:iii 400
-e
u,
u,

-
en
300

200

100

o-------------------------------
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-7 Stress strain curve of a BOS high N steel tested at 200°C after a delay time of 30 seconds after prestrain

287
700 - - . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,

600

500

-
~ 400
-
:!!:
I ll
Ill

-f 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-8 Stress strain curve of a BOS high N steel tested at 200°C after 600 seconds delay after prestrain

288
700

600

500

-
~ 400
-
:E
0

-
0
2! 300
C/J

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-9 Stress strain curve of an EAF low B steel tested at 130°C after a time delay of 30 seconds after prestrain.

289
700

600

500

-ea
a. 400
-
:ii:
t l)

-
tl)
f 300
u,

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-10 Stress strain curve of an EAF low B steel tested at 130°C after a time delay of 600 seconds after prestrain.

290
700

600

500

'ii
o.. 400
-
::iil:
u,
u,
w
0:: 300
I-
u,

200

100

0
0 0.5 1 1.5 2 2.5 3
STRAIN

Figure B-12 Stress strain curve of an EAF low N+B steel tested at 200°C after a time delay of 30 seconds after prestrain.

291
700

600

500

-
I'll
a. 400
-
:ii:
f l)
fl)

...f
u,
300

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-12 Stress strain curve of an EAF low N+B steel tested at 200°C after a time delay of 600 seconds after prestrain.

292
700

600

500

f 400
-
:ii:
1/j

-
1/j
f 300
U)

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-13 Stress strain curve of a BOS high N+B steel tested at 130°C after a delay time of 30 seconds after prestrain.

293
700

600

500

-ea
a.. 400
-
:E
,n
,n

-! 300
u,

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-14 Stress strain curve of a BOS high N+B steel tested at 130°C after a delay of 600 seconds after prestrain.

294
700

600

500

-
3!_ 400
-
:ii:
in

-
in
! 300
en

200

100

0
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-14 Stress strain curve of a BOS high N+B steel tested at 200°C after a delay of 30 seconds after prestrain.

295
700 --r----------------- -------------------- ,
600

500

':: 400
-
=5
I I)
II)
~ 300
tn
200

100

0 -1-------.....-------......------. .....------......------"T""------ -1
0 0.5 1 1.5 2 2.5 3
Strain

Figure B-15 Stress strain curve of a BOS high N+B steel tested at 200°C after a delay of 600 seconds after prestrain.

296
Appendix C

APPENDIXC

COMPUTATION OF WORK HARDENING

EXPONENTS

297
Appendix C

The following figures numbered C-1 to C-33 are plots of log stress versus log strain.

Figures C-1 to C-16 are plots of log stress-log strain for values computed from the

initial prestrain part of the stress-strain curve. While figures C-16 to C-33 represent log

stress-log strain plots for the strain to failure region. The method of calculating the

exponents is described in section 8.3.1.1.

- - - - - - - - - - - - - - - - - - - - - - - - - - - - 2.65

Log stress = 0.3033*1og strain + 2. 754 2.6

2.55

2.5

2.45

2.4

2.35

2.3

- - - . . - - - - . . . . . . - - - - , - - - -........- -........- - - . - - - - - - - - - - - 2.25


-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-1 Work hardening exponent in the prestrain region of the stress strain
curve of BOS low N tested at 130°C and 30 seconds delay time after prestrain.

298
Appendix C

2.6
log stress= 0.28421og strain+ 2.7176
2.55

2.5

2.45

/.
2.4

• 2.35

2.3

2.25
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-2 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N steel tested at 130°C after 30 seconds delay after prestrain.

- - - - - - - - - - - - - - - - - - - - - - - - - - - - - 2.65

y = 0.2441x + 2.7258 2.6

2.55

Ill
2.5
-
Ill
I!!
Cl)
CJ)
0
...J
2.45

2.4

2.35

- - - . . . . . - - -.........- -.......- - - - . - - - - - , . - - - - , . - - - . - - - - . . - - - - 2.3


-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-3 Work hardening exp(ment in the prestrain region of strain curve of
EAF low B tested at 130°C and 30 seconds delay time after prestrain.

299
Appendix C

-----------------------------,3
...... , ...
log stress= 0.2283109 strain+ 2.6293

........ ,1111111111 I

1
2.5

Ill

-!
I I)
Cl
0
1.5

..I

0.5

~ - -.......- - - . - - - - - - , - - - - - - , - - - - - - - . - - - , - - - - - , - - - - - , - - - ~ 0
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-4 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N+ B tested at 130°C and 30 seconds delay time after prestrain .

. - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - , 2.6
log stress = 0.2686109 strain + 2.6951
2.55

2.5

,,,
2.45
-
e!
u,
Cl
0 2.4
...J

2.35

2.3

- - - - . - - - -........- - - - - - . - - - - - - - - - - - - - - - - - - - - - - 2.25
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-5 Work hardening exponent in the prestrain region of the stress strain
curve of BOS low N tested at 130°C and 600 seconds delay time after prestrain.

300
Appendix C

. . . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - , 2.65

log stress= 0.2374109 strain+ 2.729


2.6

2.55

2.5

2.45

2.4

2.35

2.3

- - - - - - - . - - - - . - - - - - , - - - - - - , - - - . . . . . , . . . - -........- -.......- - - . - - - 2.25


-2 -1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-6 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N tested at 130°C and a time delay of 600 seconds after
prestrain.

. . . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - . 2.65

log stress= 0.2502109 strain+ 2.7182 2.6

2.55

Ill
Ill 2.5
...
I!?
U)

g> 2.45
.J

2.4

2.35

1----...,....---,-----,.---...----...,....----,-----,,---....----1 2.3
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-7 Work hardening exponent in the prestrain region of the stress strain
curve of EAF low B tested at 130°C after a time delay of 600 seconds after
prestrain.

301
AppendixC

.------------------------------3
log stress= 0.1728Iog strain+ 2.5941
2.5

0
1.5

0.5

1----...-----,------,----.....--------,------,----.....-------! 0
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-8 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N+B tested at 130°C and a time delay of 600 seconds after
prestrain.

. - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - 2.7

2.65
log stress= 0.2979Iog strain+ 2.7912

2.6

2.55

2.5

2.45

2.4

2.35

1----------.------------------------
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
2.3

Log Strain

Figure C-9 Work hardening exponent in the prestrain region of the stress strain
curve of BOS low N tested at 200°C after a 30 seconds delay after prestrain.

302
Appendix C

. - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - , 2.7
log stress= 0.31181og strain+ 2.8219
2.65

2.6

2.55

2.5

2.45

2.4

2.35

2.3

- - - - - - - - - - - - - - - - . - - -........- - - - . - - - - - - - , - - - - - - 2.25
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-10 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N tested at 200°C after 30 seconds delay after prestrain.

2.75

log stress = 0.3539109 strain + 2.8783 2.7

2.65

2.6

2.55

2.5

2.45

2.4

2.35

2.3
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-11 Work hardening exponent in the prestrain region of the stress strain
curve of EAF low B tested at 200°C after 30 seconds delay after prestrain.

303
AppendixC

...-------------------------------.3

.. 0 ,. co ...
log stress= 0.29451og strain+ 2.7467

' I tllfil S2 it11 I 1 2.5

jCl)
1.5
0
...J

0.5

----,------,---.....,.----,---,-----,------,------,---- 0
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-12 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N+B tested at 200°C after 30 seconds delay after prestrain.

. - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - . 2.7
log stress = 0.3221og strain + 2.8366
2.65

2.6

2.55

2.5

2.45

2.4

2.35

2.3

2.25

1------,-----,.----.---........- -.......--....-----,,-----.-----,----'1 2.2


-2 -1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-13 Work hardening exponent in the prestrain region of the stress strain
curve of BOS low B tested at 200°C after 600 seconds delay after prestrain.

304
Appendix C

- - - - - - - - - - - - - - - - - - - - - - - - - - - - . 2.7

2.65
log stress= 0.2967Iog strain+ 2.7831
2.6

2.55

2.5

2.45

2.4

2.35

2.3

- - - . . . - - - . . . . . - - - - , - - - - - , - - - - . . . . . . . - - -.......- - . . . - - -.......- -.... 2.25


-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

FigureC-14 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N tested at 200°C after 600 seconds delay after prestrain

.------------------------------,3
• 0 Ct T
' ...
log stress= 0.2947Iog strain+ 2.7318 ,
t U uet II $11 1111
11111

'1 2.5

1.5

0.5

- - - - , - - - - . . . . - - - - - , - - - - - , - - - - - , . - - - . - - - - . . . - - -.......- - - ! 0
-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-15 Work hardening exponent in the prestrain region of the stress strain
curve of EAF low B tested at 200°C after 600 seconds delay after prestrain.

305
Appendix C

- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - , 2.75

2.7
log stress = 0.34641og strain + 2.8589
2.65

2.6

2.55

2.5

2.45

2.4

2.35

2.3

- - - - - - - . - - - - - - , . - - - . - - - -.......- -.......- - - - - - - - - 2.25


-1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0
Log Strain

Figure C-16 Work hardening exponent in the prestrain region of the stress strain
curve of BOS high N+B tested at 200°C after 600 seconds delay after prestrain.

- - - - - - - - - - - - - - - - - - - - - - - - - - - - - 2.77

log stress= 0.14481og strain+ 2.7506 2.76

2.75

2.74

2.73

2.72

2.71

2.7

---.. . . --.. . . --.. . . .-----.---..--------------1-------1 2.69


-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
log strain

Figure C-17 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of BOS low N tested at 130°C and 30 seconds delay after
prestrain.

306
Appendix C

. . . . - - - - - - - - - - - - - - - - - - - - - - - - - , - - - - - - - . 2.7

log stress = 0.0679Iog strain + 2.6924 2.695

2.69

Ill
2.685
ICl)
0 2.68
...J

2.675

2.67

1------,----.------,----,-----,------,,-----+------I 2.665
-0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-18 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of BOS high N tested at 130°C and 30 seconds delay after
prestrain.

. . . . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - . 2.77

2.76
log stress= 0.1529Iog strain+ 2.7534

2.75

2.74

2.73

2.72

2.71

2.7

- - - . . . . - - - - . - - - - , - - - - - - - - - - - - - - - - - - - " " " ' 2.69


-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-19 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of EAF low B tested at 130°C and 30 seconds delay after
prestrain.

307
AppendixC

- - - - - - - - - - - - - - - - - - - - - - - - , - - - - - , 2.7

log stress = 0.06791og strain + 2.6924 2.695

2.69

-i
2.685
Ill
C>
.3 2.68

2.675

2.67

1-----,-----.---...-----,----....---- -,----1-----1 2.665


-0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-20 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of BOS high N+B tested at 130°C and 30 seconds delay after
prestrain

. - - - - - - - - - - - - - - - - - - - - - - - - - - . . . . - - - - - . 2.73

2.72
log stress= 0.17281og strain+ 2.7075
2.71

2.7

2.69

2.68

2.67

2.66

2.65

- - - - - - - - - - - - - - - . - - - - - - - - - - - - - - - - - 2.64
-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

FigureC-22 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of BOS low N tested at 130°C and 600 seconds delay after
prestrain.

308
AppendixC

- - - - - - - - - - - - - - - - - - - - - - - - - - - - - . - - - - , 2.78

2.77
log stress= 0.14911og strain+ 2.7552
2.76

2.75

2.74

2.73

2.72

2.71

- - - - - - - . - - - - - - , . - - - . . . - - - . . . . - - - . . . . - - - . . . , . . . - - - - - - 2.7
-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-23 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of BOS high N tested at 130°C and 600 seconds delay after
prestrain.

- - - - - - - - - - - - - - - - - - - - - - - - - - . . . - - - - 2.78

2.77

2.76

2.75

2.74
log stress = 0.05991og strain + 2. 7043
2.73

2.72

2.71

....
-- 2.7
.... .
2.69
-0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-24 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of EAF low B tested at 130°C and 600 seconds delay after
prestrain.

309
Appendix C

. - - - - - - - - - - - - - - - - - - - - - - - - - - , - - - - - - , 2.77

2.76
log stress= 0.16021og strain+ 2.7454
2.75

2.74

2.73

2.72

2.71

2.7

2.69

1----..---..---..----..----..-----,----- ,-----1------1 2.68


-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log Strain

Figure C-25 Work hardening exponent in a portion of 'strain to failure' region of


the stress-strain curve of BOS high N+B tested at 130°C and 600 seconds delay
after prestrain.

2.78

2.77
log stress= 0.1271og strain+ 2.7637
2.76

2.75

2.74

2.73

2.72

2.71

2.7

2.69
-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1
LOG STRAIN

Figure C-26 Work hardening exponent in 'strain to failure region of the stress-
strain curve of BOS low N tested at 200°C and 30 seconds delay after prestrain.

310
Appendix C

- - - - - - - - - - - - - - - - - - - - - - - - - - - . - - - - . 2.75

2.74
log stress= 0.11781og strain+ 2.7266

2.73

2.72

2.71

2.7

2.69

1-----,-----,-----.-----,,----,---..----........---+------'I 2.68
-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log Strain

Figure C-27 Work hardening exponent in the 'strain to failure'region of the stress-
strain curve of BOS high N tested at 200°C and 30 seconds delay after prestrain.

. . - - - - - - - - - - - - - - - - - - - - - - - - - - . - - - - . 2.79

2.78
log stress= 0.11631og strain+ 2. 7661

2.77

ICII
0
..J
2.76

2.75

2.74

2.73

1----..----..----..-----,----...,....---,-----,----+----'I 2.72
-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-28 Work hardening exponent in the 'strain to failure'region of the stress-
strain curve of EAF low B tested at 200°C and 30 seconds delay after prestrain.

311
AppendixC

. - - - - - - - - - - - - - - - - - - - - - - - - - - . , . . . . . . - - - - , 2.722

2.72
log stress= 0.02751og strain+ 2.719
2.718

2.716

2.714

2.712

2.71

2.708

2.706

.,__----,-----,------,.---......----,,----...,...---+-----'I 2.704
-0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-29 Work hardening exponent in the 'strain to failure'region of the stress-
strain curve of BOS high N+B tested at 200°C and 30 seconds delay after prestrain.

. - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - , 2.79

2.78

log stress= 0.13211og strain+ 2.7737 2.77

2.76

2.75

2.74

2.73

2.72

2.71

2.7

- - - - , - - - - - , . - - - . . - - - -.......- - - - . - - -.......- -......- - - 2.69


-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0
Log Strain

Figure C-30 Work hardening exponent in the 'strain to failure'region of the stress-
strain curve of BOS low N tested at 200°C and 600 seconds delay after prestrain.

312
Appendix C

- - - - - - - - - - - - - - - - - - - - - - - - -........- - - . 2.74

2.73
log stress= 0.1174109 strain+ 2.7192
2.72

2.71

2.7

2.69

2.68

1----....---........ ---.-----.----------.....----+-----I 2.67


-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-31 Work hardening exponent in the 'strain to failure'region of the stress-
strain curve of BOS high N tested at 200°C and 600 seconds delay after prestrain.

- - - - - - - - - - - - - - - - - - - - - - - - - - - - - 2.79

2.78

log stress= 0.1278109 strain+ 2.7633


2.77

2.76

2.75

2.74

2.73

2.72

1 - - - -........----,-----,,---.-----,----,-----,,----+-------! 2.71
-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-32 Work hardening exponent in the 'strain to failure'region of the stress-
strain curve of EAF low B tested at 200°C and 600 seconds delay after prestrain.

313
Appendix C

- - - - - - - - - - - - - - - - - - - - - - - - - - - -.... 2.78

2.77

2.76

2.75
log stress= 0.00461og strain+ 2.7037
II)

-
II)
2.74
I!!
I I)
CD
0 2.73
...J

2.72

2.71

2.7

- - - . . . . . . - - - - - - - - - - - . - - - - . . . . - - - - - - - - - - - - - - . . . , , 2.69
-0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0.05
Log strain

Figure C-33 Work hardening exponent in the 'strain to failure'region of the stress-
strain curve of BOS high N+ B tested at 200°C and 600 seconds delay after
prestrain.

314

You might also like