Download as pdf or txt
Download as pdf or txt
You are on page 1of 299

An Investigation of Surface Hot Shortness

in Low Carbon Steel

D. S. O’Neill

A thesis submitted in fulfilment


of the requirements for the degree of

Doctor of Philosophy

School of Materials Science and Engineering


Faculty of Science
The University of New South Wales
Sydney, New South Wales, 2052, Australia

March 2002
DECLARATION

I hereby declare that this submission is my own work and to the best of my knowledge it
contains no material previously published or written by another person, nor material
which to a substantial extent has been accepted for the award of any other 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 explicitly 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.

Daniel Scott O’Neill

13th March 2002

ii
ABSTRACT

Surface hot shortness is a processing problem that can affect steels containing only a
few tenths of a percent of copper. Recycled steel is particularly susceptible, since
copper becomes mixed with scrap and is very difficult to remove during steelmaking.
High temperature oxidation of the copper-bearing steel leads to enrichment of copper at
the steel surface and the formation of a copper-rich phase. During continuous casting or
hot rolling, the liquid copper-rich phase weakens grain boundaries and leads to surface
cracking. These cracks are generally not eliminated during subsequent processing and
cause surface defects in the final product. In some instances, complete cracking of the
slab or billet occurs as a consequence of hot shortness. In both cases, the result is a loss
of money due to product downgrading or scrapping.

The purpose of this study was to develop a more comprehensive fundamental


understanding of the mechanism of surface hot shortness and means to overcome the
problem. Laboratory conditions were chosen that simulate a processing environment
that leads to surface hot shortness: reheating and hot rolling of copper-bearing steels for
secondary processing. A commercially-produced low carbon steel grade (A1006) was
used as the basis for a set of model steels. These contained copper levels up to 0.48wt%,
nickel up to 0.22wt% and silicon levels up to 0.52wt%. The model steels, together with
A1006 and iron, were oxidised in air at 1050 and 1150°C, and in a CO2-N2 mixture at
1250°C for times of up to 3 hours. The copper steels oxidised at a similar rate to iron.
The copper/nickel steels and the copper/silicon steel oxidised more slowly than iron,
however the decrease was more pronounced in the latter. This was attributed to the solid
subscale formed at 1050°C, which offered a barrier to diffusion and thus lowered the
oxidation rate. At 1150 and 1250°C, the formation of a liquid subscale promoted gap
formation between the metal surface and the scale, thus lowering the overall oxidation
rate.

The extent and behaviour of copper enrichment during oxidation was determined using
optical and electron microscopy and image analysis techniques. Oxidation conditions

iii
Abstract

and alloying additions that reduced or completely suppressed the formation of a copper-
rich phase at the scale/metal interface were identified. When oxidised in air at 1050 and
1150°C, significant quantities of copper-rich phase were observed for most model
steels. The overriding factor under these conditions was the relatively high oxidation
rate. This led to the rapid development of a copper-rich layer with little copper being
accommodated within the metal via the process of back-diffusion. However, when
oxidation took place in CO2-N2 at 1250°C, the copper-rich phase did not form for a
significant amount of time; and for some model steels, not at all. This was attributed to
the considerably lower oxidation rate in the CO2-N2 atmosphere and the fact that more
copper was found diffused into the steel.

Alloying additions of nickel and silicon were found to be beneficial by reducing the
amount of copper-rich phase measured at the scale/metal interface under the conditions
investigated at 1150°C and 1250°C. This occurred because nickel and silicon addition
promoted the occlusion of copper-rich phase into the scale. In the case of nickel
addition, occlusion was promoted since these steels internally oxidised more readily.
Occlusion occurs when internal oxides coalesce and join with the external scale, leading
to engulfment of the enriched metal surface and copper-rich phase. A different
mechanism of occlusion was observed in the copper/silicon steel. Occluded particles
were often observed associated with the liquid subscale that formed at 1150 and
1250°C. If the liquid subscale reduces the interfacial energy between the external scale
and the copper-rich phase, penetration of the copper-rich phase into the scale would be
expected.

Copper enrichment during oxidation was modelled using a numerical description of the
diffusion processes involved. Given values for appropriate parameters (oxidation rate,
concentration of copper occluded, original copper concentration etc.) the model
calculated the time for commencement of copper-rich phase formation, copper-rich
phase thickness and copper concentration profile within the metal. Predictions of the
time for commencement of copper-rich phase formation for oxidation in CO2-N2 at
1250°C were in close agreement with observation. Predictions indicated that copper-
rich phase formation was almost instantaneous for oxidation in air at 1050/1150°C.
Agreement between the predicted and observed copper-rich layer thickness varied.
When the observed copper enrichment behaviour was simple; i.e. rapid formation, little

iv
Abstract

or no occlusion, and uniform distribution at the scale/metal interface, agreement was


excellent. However, this was not the case when occlusion was significant and the
copper-rich phase distribution was non-uniform. When this occurred, the measured
thickness often varied non-uniformly with time and therefore was difficult to predict
from a steady-state model. In addition, results at 1250°C indicated that the level of
occlusion decreased at longer oxidation times. The model treated occlusion in a
simplified manner, assuming that the concentration of copper occluded was fixed and
independent of time. More research is required to quantify occlusion behaviour so that
the treatment in the current model can be modified.

The cracking susceptibility of the model steels was examined by oxidising in situ and
compressing at a high strain rate using a Gleeble 3500. Oxidation was performed in air
at 1050, 1150 and 1250°C and most specimens were compressed at 1050°C. Cracks
were found to develop in the following manner: copper-rich phase precipitation at the
metal surface is followed by some penetration along grain boundaries during oxidation.
The penetration depth is small and was found not to significantly influence cracking.
The amount and depth of penetration greatly increased during compression. Cracks
develop since the presence of a copper-rich phase along grain boundaries considerably
lowers the energy required for crack formation. The amount of cracking was found to
increase with the amount of copper-rich phase precipitated at the scale/metal interface
during oxidation.

The amounts of copper, nickel and silicon in the steels were found to affect cracking
behaviour. Significant cracking was measured in a steel containing 0.47%Cu at all
oxidation temperatures. In contrast, only small amounts were detected at 1050 and
1150°C and none at 1250°C in a steel containing 0.19%Cu. This was attributed to there
being significantly more copper-rich phase measured in the 0.47%Cu steel than the
0.19%Cu steel. In general, nickel addition reduced the amount of cracking at all
temperatures; and under some conditions prevented cracking altogether. Silicon reduced
or completely suppressed cracking when the subscale formed was liquid. The beneficial
effects of nickel and silicon addition were attributed to their effect of promoting
occlusion and thus reducing the amount of copper-rich phase available to form cracks.

v
Abstract

A number of strategies to overcome or minimise the effects of surface hot shortness


were identified. Preventing copper-rich phase formation would completely overcome
the problem and this can be achieved by oxidising in low oxidation potential
atmospheres. When the copper enrichment rate is rapid and copper-rich phase formation
cannot be avoided, occlusion is the only mechanism that can alleviate the problem. Both
nickel and silicon addition were found to promote occlusion. However, increased
occlusion is likely to adversely affect descalability and more research is required to
identify what levels can be tolerated.

vi
ACKNOWLEDGMENTS
I am greatly indebted to my thesis supervisor, Prof. D.J. Young, for his excellent
guidance, advice and support. I would like to thank Dr. B. Gleeson who was mainly
responsible for initiating the project and for his early guidance.

This research was supported by funding from the Australian Research Council and BHP
Steel. I also gratefully acknowledge the receipt of an Australian Postgraduate Award
during the course of this research.

I would like to thank Dr. R. Chen and Dr. R. Smith of BHP Steel Research Laboratories
for their time and advice. Thanks are also extended to Dr. P. Manohar of the University
of Wollongong (UOW), and Mr. S. Laird of BHP Flat Products, for their help in the use
of the UOW Gleeble testing facility.

I wish to thank the staff and technical officers of the School of Materials Science and
Engineering and The Electron Microscope Unit of the University of New South Wales
for their helpful instruction and advice. Thanks are also extended to my peers for their
advice and interest; in particular Mr. N. Xu, Mr. J. Moon, Dr. R. Durham, Dr. E.
Copland, Mr. T. Munro and Mr. W. McMillan.

Finally, I wish to thank my fiancée Grace for her unyielding support and
encouragement.

vii
TABLE OF CONTENTS
1. Introduction _____________________________________________________________ 1
2. Literature Review ________________________________________________________ 4
2.1 Thermodynamic and Kinetic Aspects of Steel Oxidation _____________________ 4
2.1.1 Thermodynamics of Pure Metal Oxidation ____________________________ 4
2.1.2 Oxidation of Pure Metals __________________________________________ 7
2.1.2.1 Mechanism of Scale Formation___________________________________ 7
2.1.2.2 Kinetics of Scale Formation _____________________________________ 9
2.1.2.3 Wagner Theory of Oxidation____________________________________ 10
2.1.2.4 The Oxidation of Iron in Air or Oxygen ___________________________ 14
2.1.3 The Oxidation of Alloys _________________________________________ 22
2.1.3.1 External Scale Formation ______________________________________ 23
2.1.3.2 Internal Oxidation beneath an External Scale _______________________ 27
2.1.3.3 Conditions for the Transition from Internal to External Oxidation _______ 28
2.1.3.4 Conditions for the Simultaneous External and Internal Oxidation of B ___ 29
2.1.3.5 The Oxidation of Steel_________________________________________ 31
2.1.3.5.1 Effect of Alloying Elements _________________________________ 31
2.1.3.5.2 Effect of Oxidant Atmosphere _______________________________ 34
2.2 The Phenomenon of Surface Hot Shortness _______________________________ 41
2.2.1 Residual Elements in Steels _______________________________________ 41
2.2.2 Mechanism of Surface Hot Shortness _______________________________ 47
2.2.2.1 Introduction _________________________________________________ 47
2.2.2.2 Copper-rich phase formation ____________________________________ 52
2.2.2.3 Penetration and Cracking ______________________________________ 56
2.2.2.3.1 Liquid Metal Embrittlement and Wettability ____________________ 56
2.2.2.3.2 Proposed Mechanisms of Cracking____________________________ 63
2.2.3 Factors Affecting Surface Hot Shortness_____________________________ 67
2.2.3.1 Steel Chemistry ______________________________________________ 68
2.2.3.1.1 Copper__________________________________________________ 68
2.2.3.1.2 Nickel __________________________________________________ 70
2.2.3.1.3 Tin _____________________________________________________ 76
2.2.3.1.4 Silicon __________________________________________________ 78
2.3 Modelling of Surface Hot Shortness ____________________________________ 80
2.4 Aims and Focus ____________________________________________________ 86

viii
Table of Contents

3. Experimental Procedure __________________________________________________ 87


3.1 Introduction _______________________________________________________ 87
3.2 Oxidation Behaviour of Model Steels ___________________________________ 87
3.3 Enrichment Behaviour of Residual Elements during Oxidation _______________ 93
3.4 Cracking Behaviour of Model Steels ____________________________________ 95
4. Oxidation Behaviour of Model Steels ________________________________________ 99
4.1 Introduction _______________________________________________________ 99
4.2 Results ___________________________________________________________ 99
4.2.1 Initial Oxidation Experiments _____________________________________ 99
4.2.2 Oxidation Behaviour ___________________________________________ 102
4.2.3 Rate of Oxidation______________________________________________ 107
4.2.4 Internal Oxidation _____________________________________________ 111
4.3 Discussion________________________________________________________ 117
4.3.1 Oxidation in Simulated Reheat Atmosphere _________________________ 117
4.3.2 Scale Separation_______________________________________________ 118
4.3.3 Scale Morphology _____________________________________________ 119
4.3.4 Rate of Oxidation______________________________________________ 124
4.3.4.1 Copper Steels_______________________________________________ 124
4.3.4.2 Copper/Nickel Steels _________________________________________ 125
4.3.4.3 Copper/Silicon Steel _________________________________________ 126
4.3.5 Internal Oxidation _____________________________________________ 127
4.4 Summary and Conclusions ___________________________________________ 133
5. Enrichment Behaviour of Residual Elements during Oxidation ___________________ 135
5.1 Introduction ______________________________________________________ 135
5.2 Results __________________________________________________________ 135
5.2.1 Copper-rich Phase Morphology___________________________________ 135
5.2.2 Copper-rich Phase Formation ____________________________________ 140
5.2.2.1 General Remarks ____________________________________________ 140
5.2.2.2 Copper Steels_______________________________________________ 142
5.2.2.3 Copper/Nickel Steels _________________________________________ 146
5.2.2.4 Copper/Silicon Steel _________________________________________ 153
5.2.3 Amount of Copper-rich Phase ____________________________________ 158
5.3 Discussion________________________________________________________ 163
5.3.1 Copper-rich Phase Morphology and Distribution _____________________ 163
5.3.1.1 Copper-rich Phase at Scale/Metal Interface _______________________ 163
5.3.1.2 Copper-rich Phase in Scale ____________________________________ 164
5.3.2 Residual Element Enrichment ____________________________________ 167

ix
Table of Contents

5.3.3 Amount of Copper-rich Phase ____________________________________ 173


5.3.3.1 Initial Considerations_________________________________________ 173
5.3.3.2 Measured Results____________________________________________ 176
5.4 Summary and Conclusions ___________________________________________ 180
6. Modelling of Residual Element Enrichment __________________________________ 183
6.1 Introduction ______________________________________________________ 183
6.2 Development of an Analytical Model __________________________________ 183
6.2.1 Initial Considerations ___________________________________________ 183
6.2.2 Analytical Model Assuming Immediate Copper-rich Phase Formation ____ 185
6.3 Development of a Numerical Model ___________________________________ 189
6.3.1 Initial Considerations ___________________________________________ 189
6.3.2 Stability Criterion for Convergence of Explicit Finite Difference Method __ 191
6.3.3 The Fixed Grid Method _________________________________________ 192
6.3.4 The Initial and Boundary Conditions_______________________________ 196
6.3.5 Transient Condition Stage _______________________________________ 196
6.3.6 Fixed Boundary Condition Stage__________________________________ 200
6.4 Application of the Analytical Model ___________________________________ 203
6.5 Application of the Numerical Model ___________________________________ 205
6.5.1 Initial Simulations _____________________________________________ 205
6.5.2 Simulations of Time Taken for Precipitation of the Copper-rich Phase ____ 208
6.5.2.1 Simulations at 1250°C________________________________________ 208
6.5.2.2 Simulations at 1050 and 1150°C ________________________________ 213
6.5.3 Simulations of Predicted Copper-rich Layer Thickness ________________ 218
6.5.4 Other Simulations _____________________________________________ 223
6.6 Summary and Conclusions ___________________________________________ 228
7. Cracking Behaviour of Model Steels _______________________________________ 230
7.1 Introduction ______________________________________________________ 230
7.2 Results __________________________________________________________ 230
7.2.1 General Points ________________________________________________ 230
7.2.2 Appearance of Cracks __________________________________________ 232
7.2.3 Measurement of Crack Characteristics _____________________________ 235
7.3 Discussion________________________________________________________ 243
7.3.1 Possible Limitations of the Hot Cracking Test _______________________ 243
7.3.2 Mechanism of Crack Formation __________________________________ 243
7.3.3 Variation of Cracking Behaviour with Testing Conditions ______________ 245
7.3.4 Variation of Cracking Behaviour with Alloy Composition ______________ 246
7.4 Summary and Conclusions ___________________________________________ 251

x
Table of Contents

8. Summary and Significance of Results_______________________________________ 254


9. References ____________________________________________________________ 260
Appendix A _______________________________________________________________ 267
Appendix A.1 Composition in simulated reheat atmosphere _______________________ 267
Appendix A.2 Conversion relationships between rate constants_____________________ 270
Appendix A.3 Mass transfer in viscous flow regime______________________________ 275
Appendix B _______________________________________________________________ 278
Appendix B.1 Maximum copper-rich layer thickness _____________________________ 278
Appendix B.2 Time taken for dissolution of a copper-rich layer of known thickness ____ 281
Appendix C _______________________________________________________________ 283
Appendix C.1 Derivation of an equation describing diffusion within the copper-rich phase283
Appendix C.2 Derivation of an equation describing diffusion within the metal _________ 287
Appendix C.3 Estimation of rate constant using an Arrhenius relationship ____________ 288

xi
CHAPTER
ONE
Introduction

Surface hot shortness is a problem mainly confined to the production of steel via
electric arc furnaces (EAF). The high level of scrap used in this production method
introduces higher proportions of residual elements (e.g. copper, nickel and tin) than are
encountered in steels produced by more traditional methods. High temperature
oxidation of the steel leads to surface hot shortness as a result of the enrichment of
residual elements, forming an embrittling liquid phase. The problem manifests itself as
surface cracking during secondary processing of the steel. These cracks are generally
not eliminated during subsequent processing and cause surface defects in the final
product. In some instances, complete cracking of the slab or billet occurs as a
consequence of hot shortness. In both of these cases, the result is a loss of money due to
product downgrading or scrapping.

Although the amount of EAF-processed steel in Australia is comparatively small, it is


certainly a major processing route in other countries. EAF-produced steel continues to
grow at the expense of that produced from more traditional methods. As a result,
problems associated with residual elements will become more important as the amount
of steel recycling grows. There are also economic gains to be made in the utilisation of
scrap high in residual element content, due to the relatively lower cost of this feed
source.

The current project has been initiated to investigate the mechanisms of surface hot
shortness cracking and means to overcome the problem. This field has been extensively
studied, but there are areas where knowledge is presently lacking. For example, little is
known of the kinetics of copper enrichment and copper-rich phase formation. Little data
has been published regarding the enrichment and segregation of elements to the copper-
rich phase, and the level of iron accommodated in this phase in the presence of other

1
Introduction

elements. The resulting compositions would affect both the solubility of copper in
austenite and the melting point of the copper-rich phase, both obviously important
factors in surface hot shortness. Alloying elements such as nickel and silicon have been
shown to inhibit surface hot shortness cracking and hence knowledge of the enrichment
and effect these have on copper-rich phase formation is important. In addition, there
exists some confusion as to whether a stress is required to cause copper-rich phase
penetration along grain boundaries and subsequent cracking.

The purpose of the present work was to determine experimentally the extent and
behaviour of the copper enrichment process during oxidation, and to model the process.
Details of previous models predicting copper enrichment are either flawed or not
accessible (i.e. proprietary), and so the development of a new model is warranted.
Modelling was intended to allow the effect of various conditions to be predicted and to
provide a useful method for predicting steel composition/mill conditions in the future.

The current project was carried out in collaboration with the Australian Research
Council and BHP Steel Research Laboratories. BHP Steel has a strong interest in scrap-
based electric arc steelmaking. A low carbon steel grade produced by BHP Steel,
A1006, was chosen as the basis for a set of model steels. The model steels contained
copper levels up to 0.48wt%, nickel up to 0.22wt% and silicon levels of 0.52wt%.
Along with A1006 and iron, the steels were oxidised between 1050 and 1250°C for
times of up to 3 hours. Scaling kinetics were determined by measuring the scale
thickness or by thermogravimetric analysis. Copper-rich phase formation was
investigated using optical and scanning electron microscopy techniques. The amount of
copper-rich phase formed was determined using image analysis of back-scattered
electron images. The enrichment behaviour of residual elements during oxidation was
characterised by electron probe micro-analysis. Copper enrichment was modelled using
analytical and numerical descriptions of the diffusion processes involved. Finally, the
cracking susceptibility of the model steels was examined by oxidising in situ and
compressing at a high strain rate using a Gleeble 3500.

The literature review, Chapter 2, is made up of three main sections. Section 2.1
discusses the thermodynamic and kinetic aspects of steel oxidation. Surface hot
shortness requires oxidation to take place so that copper enriches at the metal surface. It

2
Introduction

is therefore pertinent to review issues relating to the oxidation of steels. Section 2.2
deals with issues relating specifically with surface hot shortness: residual elements and
their enrichment during oxidation, liquid metal embrittlement and wettability, proposed
mechanisms of cracking, and the effect of temperature and alloying additions. Lastly,
Section 2.3 reviews models that predict copper enrichment during oxidation.

Chapter 3 reports the experimental procedures used, including sample production,


oxidation experiments, techniques used to investigate the enrichment behaviour of
residual elements, and how the cracking behaviour of the steels were investigated.
Chapter 4 reports the results from the oxidation experiments and discusses how
oxidation conditions and alloying additions affect the oxidation behaviour of the steels.
Residual element enrichment behaviour is reported in Chapter 5 and conditions are
identified where the formation of the copper-rich phase was reduced or inhibited.
Analytical and numerical modelling of copper enrichment during oxidation is reported
in Chapter 6. Chapter 7 reports on the cracking behaviour of the steels and is interpreted
in terms of the amount of copper-rich phase formed during oxidation. Finally, Chapter 8
reviews the findings of the project and discusses their significance.

3
CHAPTER
TWO
Literature Review

2.1 Thermodynamic and Kinetic Aspects of Steel Oxidation

There are a number of factors that need to be satisfied for a steel to suffer surface hot
shortness. Foremost, the steel must contain copper and undergo oxidation to enable its
enrichment beneath the scale layer. As oxidation is a prerequisite for surface hot
shortness, it is pertinent to discuss issues relating to the oxidation of pure metals and
alloys, with particular reference to iron and steel.

2.1.1 Thermodynamics of Pure Metal Oxidation

Thermodynamics can be used to determine whether a metal is likely to form an oxide


when exposed to oxidising conditions. It is not the only criterion that needs to be
considered in determining whether an oxide will form. An understanding of the kinetics
of the reaction and those of less stable oxides may be required to establish what oxide(s)
form under particular conditions. Thermodynamics does however determine the most
stable oxide that could form.

When two substances are combined at constant temperature and pressure (e.g. a metal
and oxygen), the Gibbs free energy of the system will continue to change until
equilibrium is established. The Gibbs free energy change, ∆G can thus be thought of as
the driving force for the reaction. At constant temperature and pressure, ∆G is defined
as:

∆G = ∆H - T∆S (2.1)

4
Literature Review

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


entropy change. A reaction will not proceed unless ∆G < 0, is at equilibrium when ∆G =
0, and is thermodynamically unfavourable if ∆G > 0.

The reaction of a metal with oxygen to form one mole of its oxide can be expressed as:

y
xM + O2 = MxOy (2.2)
2

where M is the reacting metal and MxOy is the oxide. The driving force for Eqn. (2.2)
can be expressed in terms of the standard Gibbs free energy change, ∆G°:

 aM xO y 
∆G = ∆G° + RT ln   (2.3)
x
( )
 (a M ) ⋅ a O 2
y 2


where ai is the chemical activity of component i and R the gas constant. Some
simplifications can be made to Eqn. (2.3) if the solids are assumed to be pure and the
gas ideal. The activity of a pure solid is unity, while the activity of an ideal gas is equal
to its pressure. At equilibrium (∆G = 0), Eqn. (2.3) reduces to:

y
∆G° = RT ln p Odiss (2.4)
2 2

where p Odiss
2
is the partial pressure of oxygen controlled by the M/MxOy equilibrium,

known as the equilibrium dissociation pressure of MxOy. As a result, MxOy will not form
unless the ambient oxygen pressure exceeds p Odiss
2
. Standard Gibbs free energy change

data and the dissociation pressure of metal oxides as a function of temperature are
conveniently summarised in an Ellingham diagram (Figure 2-1).

5
Literature Review

Figure 2-1: Ellingham diagram for various oxides [1].

Gas phase reactions are often used to achieve low oxygen partial pressures. The two
most important are H2O/ H2 and CO2/CO:

xM + yH2O = MxOy + yH2 (2.5)

xM + yCO2 = MxOy + yCO (2.6)

and particular H2O/ H2 or CO2/CO ratios correspond to p Odiss


2
for any given oxide.

6
Literature Review

The methods discussed above can be used to determine the thermodynamically most
stable oxide under particular conditions. Metastable oxides (i.e. those which are
thermodynamically unstable) often remain after corrosion in gases [2]. For example,
FeO is unstable below 570°C. From a thermodynamic standpoint, it would therefore be
assumed that the scale formed on steel after secondary working be Fe3O4 or Fe2O3. In
practice, the major component of mill scale at room temperature is FeO [2]. This is a
result of the slow decomposition kinetics of FeO. Hence, an understanding of both
thermodynamic and kinetic aspects for oxidation reactions is required to identify
products. Kinetic aspects will be discussed in Section 2.1.2.2.

2.1.2 Oxidation of Pure Metals

2.1.2.1 Mechanism of Scale Formation

The reaction of an oxidant and metal may depend on a number of factors, and the
reaction mechanism can be quite complex [3]. Some of the main phenomena in the
oxidation of metals are illustrated in Figure 2-2. When a clean metal is exposed to an
oxidant (e.g. O2), it is first adsorbed on the surface of the metal. Oxide nuclei form at
favourable sites and commonly grow laterally [4]. A continuous film is produced when
the nuclei grow and coalesce, completely covering the metal surface. The adsorption of
oxidant and initial oxide formation are dependent on the surface orientation, crystal
defects at the surface, surface preparation and impurities in both the metal and gas [3].
Once a continuous film/scale separates the metal and oxidant, further oxide growth
requires the transport of electrons through the layer to the oxygen atoms adsorbed on
the surface, and oxygen ions, metal ions or both diffusing through the layer. Hence,
further growth is determined by the availability of oxidant in the atmosphere and the
transport rate of the reactant or electrons through the layer [4].

7
Literature Review

Figure 2-2: Schematic illustration of main phenomena occurring during metal


oxidation [4].

If the oxide layer remains continuous and compact as its thickness increases and if
interfacial processes are rapid, solid state diffusion becomes the rate-limiting step.
Under such conditions the reaction is often described by a parabolic rate equation [4].
Cavities and microcracks can develop within the scale during its growth. The presence
of these defects will affect transport of reactants through the scale. Growth stresses can
result in the formation of cracks and scale spallation, exposing the metal surface to the
atmosphere. When scale growth is controlled by transport through the gas or by a phase
boundary reaction, it is common to observe oxidation rates that are constant with time.
Other processes that can occur include the formation of liquid or volatile oxides.

Whilst the initial formation of thin films is an important stage in oxide development, the
films quickly grow into thick scales at high temperatures, due to the high rate of
diffusional transport. Hence, further discussion will be limited to the growth of scales
thicker than 1 µm.

8
Literature Review

2.1.2.2 Kinetics of Scale Formation

When compact scales are formed on metals and where there is sufficient oxygen (or
oxidising gas) available at the scale surface, the rate of oxidation will be limited by
solid-state diffusion [3]. Diffusion may occur via lattice, grain boundary, or short-circuit
mechanisms. As the scale layer grows, the diffusion distance increases and as a result
the reaction rate decreases with time (illustrated in Figure 2-3). In such cases a parabolic
rate law is usually observed. This law can be expressed as [5]:

∂X k '
= (2.7)
∂t X
or
X 2 = 2k ' t + C (2.8)

where X is the scale thickness, t is the time, k' is the parabolic rate constant and C the
integration constant.

Linear

X Parabolic

Figure 2-3: Parabolic and linear oxidation (X is the thickness of oxide film, t is the
oxidation time) [3].

If the rate of scale formation of a compact, adherent scale is entirely limited by a surface
reaction or by gas-phase diffusion, the oxidation process obeys a linear rate law [6]. A
linear rate law might also be observed where the oxide is volatile or molten, if the scale
spalls off or cracks, or if a porous, unprotective oxide forms on the metal [2]. A linear

9
Literature Review

rate law indicates that solid-state diffusion through the scale is not rate limiting. The
kinetic equations are:

∂X
= kl (2.9)
∂t
or
X = kl t + C (2.10)

where kl is the linear rate constant.

Other rate equations have been suggested when presenting kinetic data for scale growth.
It is believed that these describe situations representing a combination of the processes
listed above [2].

2.1.2.3 Wagner Theory of Oxidation

The parabolic rate law was first derived from first principles by Carl Wagner [7] in
1933 and was the most important single contribution to the understanding of high
temperature oxidation of metals [3]. Wagner investigated the flux of particles through
an oxide layer formed on a pure metal and considered their concentration, mobility and
the forces acting upon them. Wagner made the following simplifications/assumptions
[1,5]:

• The oxide layer was assumed to be compact and adhered to the surface.
• Oxide growth occurred by a single diffusion mechanism, e.g. the diffusion of
cations to the scale/gas interface occurring in a p-type oxide.
• The transport of charged species was the rate controlling step. Interfacial
reactions were fast compared to the transport of charged species and hence were
not rate controlling.
• Local thermodynamic equilibrium was assumed to exist at each interface (i.e. at
the metal/oxide and oxide/gas interface) and throughout the scale.
• The oxide layer illustrated only small deviations from stoichiometry.

10
Literature Review

• Space-charge effects were neglected since the distances over which they act
were much smaller than the thickness of the scale.
• Oxygen solubility in the metal was neglected.

If thermodynamic equilibrium exists at each interface, then activity gradients would be


established for the components across the oxide layer. Figure 2-4 illustrates
concentration gradients within an oxide scale that contain mostly metal vacancies, i.e. a
cation deficient semi-conductor or p-type scale.

Metal Oxide Gas


(M) (MO) (O2)
Cations
Anions

Electrons
o o
aM pO2

 ∆G o MO  1  ∆G o MO 
pO2
i
= exp  aM =
o
 
(p )  2 RT
1
 RT  o 2 
O2

M = M2+ + 2e- M2+ + 2e- + ½ O2 = MO


or or
M + O2- = MO + 2e- ½ O2 + 2e- = O2-

Figure 2-4: Scale formation based on Wagner’s model for a p-type scale [1].

i
The partial pressure of oxygen at the metal/oxide interface, p O 2 , is equal to the

equilibrium dissociation pressure of the oxide in contact with its metal, whereas at the
o
oxide/oxygen interface it is equal to the oxygen pressure in the gas phase, p O 2 . For a

p-type scale, metal ions diffuse away from the M/MO to the MO/O2 interface, while
o i
vacancies migrate in the opposite direction. Since p O 2 > p O 2 , metal ion vacancies are

continuously produced at the MO/O2 interface and consumed at the M/MO interface.

Wagner stated that equivalent amounts of oppositely charged species are transported
across the scale layer such that no net current flows [5]. Such species were influenced
11
Literature Review

by chemical and electrical-potential gradients. The rates of migration of these species


(cations, anions and electronic defects) are such that a state of electroneutrality is
maintained within the scale. Wagner noted that electronic defects migrate much more
rapidly than ionic species and usually one ionic species predominates with respect to
another. Usually, one of the ionic mobilities can be ignored, and the scale can be
described as either a cation or anion diffuser. By making suitable substitutions, Wagner
eliminated the electric-field gradient, replaced mobilities with diffusivities and used
oxygen pressures instead of chemical potentials [5] and arrived at the result:

1 pO o  z 
kp = C o ∫ 2i  Do + c Dm d ln p O 2 (2.11)
2b p O2
 za 

where Co is the concentration of oxygen per unit volume, b the number of oxygen atoms
per molecule of oxide, zc and za the valences of the cation and anions, Do and Dm the
i
diffusivity of oxygen and metal in the oxide, and p O 2 is the oxygen pressure at the

oxide/metal interface (i.e. p Odiss


2
). Comprehensive reviews of Wagner’s treatment have

been given by Kofstad [3], Birks and Meier [1] and Yurek [6].

Eqn. (2.11) can be simplified for certain limiting cases. Consider the growth of an oxide
by metal transport diffusing via vacancies having an effective charge α. The value of
i o i
p O 2 is usually very low and thus can be ignored ( p O 2 >> p O 2 ). Oxygen diffusion is

insignificant compared to metal diffusion and can also be ignored (Dm >> Do). Hence,
the parabolic rate constant can be expressed as [4]:

k p = (α + 1)Dm
o
(2.12)

where Dmo is the metal ion diffusivity in the oxide at the oxide/gas interface. If oxide
growth occurs by interstitial metal ions (e.g. as for n-type oxides) with an effective
charge given as α, the parabolic rate constant can be expressed as:

k p = (α + 1)Dm
i
(2.13)

12
Literature Review

where Dmi is the metal ion diffusivity in the oxide at the metal/oxide interface. Since
Dmo ∝ p O 2 , the rate constant in Eqn. (2.12) is dependent on the ambient oxygen
o

pressure. Similarily, Dmi ∝ p O 2 , but since p O 2 is fixed by the thermodynamics of the


i i

metal/oxide equilibrium, the rate constant in Eqn. (2.13) should be independent of the
ambient oxygen pressure. Similar expressions can be found for oxide growth controlled
by oxygen diffusion on vacancy or interstitial sites.

Wagner’s theory predicts that parabolic scale growth in a p-type oxide is dependent
upon the ambient oxygen pressure, but is independent for an n-type oxide. An example
illustrating the dependence of the parabolic rate constant on oxygen pressure is
illustrated in Figure 2-5. The rate becomes independent of the oxygen pressure when
Co3O4 (an n-type oxide) forms. The transition point at each temperature and pressure
corresponds to the CoO/Co3O4 equilibrium.

Figure 2-5: Dependence of the parabolic rate constant on the oxygen pressure for
cobalt oxidation [3].

Simplifications made to Wagner’s relationship, e.g. Eqns. (2.12) and (2.13), illustrate
that the parabolic rate constant is a function of some diffusivity in the scale. In general,
the rate constant differs by less than an order of magnitude from the diffusivity of the

13
Literature Review

relevant species at the relevant pO2 [5]. Good agreement between the calculated
parabolic rate constants and those measured for iron [8], cobalt [9] and copper [10] have
been attained. However, Wagner’s theory presents an ideal model and in its original
form was limited to the growth of dense scales controlled by lattice diffusion of species
through the scale [11]. Such effects as grain boundary diffusion, the formation of multi-
layered scales and oxidation in mixed reactant gases were not considered in the original
model and so can lead to deviations from Wagner theory predictions. However, the
model serves as a useful basis for interpreting high temperature oxidation.

2.1.2.4 The Oxidation of Iron in Air or Oxygen

A description of the high temperature oxidation of iron is more complex because


different oxides may form. Iron can oxidise to form more than one valence (+2 and +3)
and hence a multi-layered scale can form. Assuming that the ambient oxygen pressure is
sufficiently high, the oxidation of iron in air forms a scale consisting of an inner wustite
(Fe1-xO) layer, a middle layer of magnetite (Fe3O4) and an outer layer of hematite
(Fe2O3) above temperatures of 570°C (see Figure 2-6). Note that p-type oxides are
denoted by the formula M1-xO, while n-type oxides are denoted by M1+xO. As illustrated
in the iron-oxygen phase diagram (see Figure 2-7), FeO is not stable for temperatures
below 560°C. It should be pointed out that many authors [1,8,12,13] suggest that the
lower temperature limit for FeO stability is 570°C. The layers are arranged so that that
the most metal-rich layer is adjacent to the metal and the most oxygen-rich layer on the
outside.

FeO and Fe3O4 are p-type oxides and are characterised by the migration of iron ions.
FeO exists over a wide range of stoichiometry that increases with increasing
temperature. The mobility of cations and electrons is thus extremely high in this phase
owing to the high cation vacancy concentration [1]. In contrast, Fe3O4 and Fe2O3 exhibit
only slight variations in stoichiometry, although some variation at high temperatures is
observed for Fe3O4 [1,14]. Some confusion exists over the exact defect structure of
Fe2O3. Disorder on the anion sub-lattice only has been reported [15] for which oxygen
ion migration is expected to dominate. However, it has also been suggested that Fe2O3
growth occurs by cation migration [16].

14
Literature Review

FeO Fe3O4 Fe2O3

200µm

Figure 2-6: Multi-layered scale formation on iron (oxidised for 3 hours at 1050°C
in flowing air).

Figure 2-7: The iron-oxygen phase diagram [17].

15
Literature Review

The mechanism for oxide growth on iron is represented schematically in Figure 2-8.
Iron oxidises at the Fe/FeO interface according to:

Fe = Fe2+ + 2e- (2.14)

Figure 2-8: Oxidation mechanism of iron to form a three-layered scale of FeO,


Fe3O4 and Fe2O3 above 570°C showing diffusion steps and interfacial reactions [1].

Migration of the iron ions and electrons occurs through the FeO layer via cation
vacancies and electron holes, respectively. When these species reach the FeO/Fe3O4
interface, Fe3O4 is reduced to FeO according to:

Fe2+ + 2e- +Fe3O4 = 4FeO (2.15)

Any surplus iron ions and electrons from this reaction are available to migrate through
the Fe3O4 layer to form Fe3O4 at the Fe3O4/Fe2O3 interface:

Fen+ + ne- +4Fe2O3 = 3Fe3O4 (2.16)

where n = 2 or 3 for Fe2+ or Fe3+, respectively. The reaction(s) to be considered for the
formation of Fe2O3 depend on whether iron or oxygen ions are mobile within this phase:

If iron ions mobile: 2Fe3+ + 6e- +1.5O2 = Fe2O3 (2.17)

16
Literature Review

If oxygen ions mobile: 2Fe3+ + 3O2- = Fe2O3 (2.18)

If iron ions are considered mobile, Fe2O3 formation occurs at the scale/gas interface. If
oxygen is mobile, Fe2O3 forms at the Fe3O4/Fe2O3 interface. The ionisation of oxygen at
the scale/gas interface is given by:

0.5O2 + 2e- = O2- (2.19)

When more than one oxide layer forms, the thickness of each layer depends on the
individual stabilities, the rates of growth (i.e. the diffusivity of the mobile species) and
on the molar volume at a particular temperature [14]. For the formation of a three-
layered scale on iron, the thickness of FeO is generally much larger than Fe3O4 and
Fe2O3. This is mainly due to the lower rate of diffusion within Fe3O4 and Fe2O3 [12].
Consider the flux of iron cations across the FeO/Fe3O4 interface. If the interface is
stationary, the amount of iron cations arriving at the interface must be equal to the
amount leaving. Since the mobility of cations is greater in FeO than Fe3O4, the
thickness of the FeO layer will be greater so that the diffusion gradient is shallower.
Paidassi [18] reported that the growth of each oxide layer when iron is oxidised in air
between 700 to 1250°C follows the parabolic rate law and that the relative thicknesses
of FeO/Fe3O4/Fe2O3 are relatively constant in the ratio 95:4:1. The ratio changes at
lower temperatures, as the growth of Fe3O4 is favoured over FeO.

Figure 2-9 illustrates a stability diagram for the Fe-O system in which the reciprocal
temperature is plotted against the oxygen partial pressure. The phase field boundaries
represent the equilibrium existing between adjacent phases with respect to the oxygen
partial pressure and temperature. For example, at 1100°C FeO is stable for oxygen
partial pressures between approximately 5×10-13 and 1×10-10 atm. If the oxygen partial
pressure could be varied between these limits and a state of equilibrium achieved at the
FeO/gas interface, it should be possible to observe a variation in the parabolic rate
constant. However the partial pressures required to form FeO-only scales are very low,
and in practice are achieved using gas mixtures. Investigations of the oxidation of iron
in CO-CO2 [19,20], CO2 [21] and H2O-H2 [22] have shown that oxidation was initially
controlled by either diffusion in the gas phase or reaction at the scale/gas interface.
Under such conditions linear oxidation kinetics would be observed. If the oxygen partial
17
Literature Review

pressure of a reaction gas were to be increased, Fe3O4 and then Fe2O3 would eventually
become stable. When these phases become stable, the resulting parabolic rate constant
of the scale layer would be relatively unaffected by the ambient oxygen partial pressure
[1]. The vacancy concentrations at the metal/FeO, FeO/Fe3O4 and Fe3O4/Fe2O3
interfaces are fixed by thermodynamic equilibrium and the growth of these layers would
be independent of the ambient oxygen pressure. In contrast, the growth of the Fe2O3
layer would be dependent. However, since this layer generally accounts for such a small
amount of the scale, the variation in the overall parabolic rate is minimal.

Figure 2-9: Stability diagram for Fe-O system [23].

In the discussion so far, the description of iron oxidation has been limited to situations
of classical scale growth. However, a number of deviations can occur and should be
discussed. The first deviation to be considered is separation of the scale from the metal
during its growth. Once this occurs, the flux of iron into the oxide is halted, and
therefore growth ceases, resulting in a thinner oxide layer. Scale separation or
detachment during reaction is a result of a stress-relieving mechanism. During
oxidation, the scale thickens while the oxide/metal interface recedes. Stresses are
developed in the scale as it attempts to maintain adherence to the metal substrate [4].

18
Literature Review

The compressive stresses generated in the scale can be relieved by one or more of the
following: plastic deformation of the scale or of the metal, detachment (spalling) at the
metal/scale interface and elastic failure (cracking) of the scale [24]. The plasticity of a
scale layer decreases with temperature, and detachment is expected to be more likely at
lower temperatures. Figure 2-10 illustrates the relationship between the relative
thickness of the oxide layers formed on iron with respect to temperature. Sachs and
Tuck [12] stated that although the lowering of the iron diffusivity in FeO relative to
Fe3O4 would have some effect on the change in relative thickness, it was more likely
that scale detachment at lower temperatures played a more significant role. Complete
detachment from the metal would not only result in a thinner scale layer, but the relative
proportion of the various oxides would change as diffusion took place. The detached
scale equilibriates with the gas, with iron and oxygen diffusing through the layers in an
attempt to eliminate their chemical potential gradient. The overall effect is that the
relative proportion of Fe3O4 and Fe2O3 increases at the expense of FeO. This is
illustrated in Figure 2-11, where a non-adhered layer has relatively more Fe3O4 and
Fe2O3 than the adhered layer.

Figure 2-10: Approximate percentages of FeO, Fe2O3 and Fe3O4 on iron oxidised in
oxygen [12,25].

19
Literature Review

Fe2O3
200µm
Fe3O4

FeO

Iron

Figure 2-11: Adherent and non-adherent scale formed on iron after 3 hours
oxidation at 1050°C in flowing air.

The second deviation to be considered from classical scale growth is the formation of
porous scales. Such scales contain voids and cavities of various shapes and sizes,
generally located in the inner scale region (see Figure 2-12). Engell et al. [26] reported
that porous scales formed at prolonged oxidation at 850°C and for progressively shorter
times at higher temperatures. Inert markers were found embedded in scales exhibiting
porosity [27], indicating the inward migration of oxygen. It is more probable that
oxygen is transported to the metal surface through the formation and healing of cracks
than by its diffusion through the scale [12].

The applicability of the classical scale model would depend on the relative amount of
porosity and gap formation. For example, deviations from parabolic scale growth have
been attributed to the presence of gap formation in iron during oxidation in air and
oxygen [12]. Figure 2-13 shows parabolic scale growth plots for iron oxidised in air,
oxygen and steam. Deviations from parabolic growth occur as the oxidation time
increases for both air and oxygen. This was attributed to the appearance of gaps at the
scale/metal interface. In contrast, after an initial period (corresponding to linear scale
growth), oxidation in steam resulted in ideal parabolic oxidation. There have been two
suggestions for the effect of steam. Firstly, scale formed in steam is more plastic, creeps
more readily and is therefore less likely to form gaps [28]. Secondly, gas reactions in

20
Literature Review

pores allow the diffusion of oxygen across the scale and the continued diffusion of iron
[29]. It is thought that this counteracts the disruption of the flux of iron to form scale,
enabling the oxidation rate to be maintained.

Figure 2-12: Porous scale formed on iron at oxidised in 4O2-N2 at 1200°C for 1
hour [30].

Figure 2-13: Oxidation of pure iron at 1000°C in various atmospheres [12].

21
Literature Review

2.1.3 The Oxidation of Alloys

The oxidation of pure metals leads to well-defined layered scales. However, alloy
oxidation presents a more complex situation. The extra thermodynamic degree(s) of
freedom of alloys can allow the formation of multiphase scales, internal oxidation and
non-planar phase interfaces [31]. Different elements with different properties may
interact simultaneously, making it difficult to isolate the effect of an element in an alloy.
Some other factors relevant to alloy oxidation illustrating the complex nature of the
phenomenon include [1]:

• Metals present in the alloy will have different affinities for oxygen.
• Ternary or higher order oxides may form.
• Some solubility between different oxides may exist.
• Different metal ions will have different mobilities through the oxide phases.
• Metals in the alloy will have different diffusivities in the alloy.

Alloy oxidation behaviour depends on a number of thermodynamic and kinetic


parameters. It may be described by classifying the possible types of scale structures that
may form. A comprehensive classification was presented by Bastow et al. [32]. The
situation is simplified by considering a binary, single phase alloy A-B, where A is the
more noble and B the more reactive component. The possible scale structures where the
two oxides are insoluble in each other and do not form spinels was summarised by
Gesmundo and Niu [33]. As illustrated in Figure 2-14, these include (1) external AO
scales; (2) external AO scales and internal oxidation of B (BOv); (3) mixed external AO
+ BOv scales; (4) external BOv scales, and (5) external and internal oxidation of B. The
approaches used to predict the nature of scales formed on A-B alloys will now be
discussed. Unless otherwise stated, the oxidant partial pressure is assumed to be higher
than that for the A-AO equilibrium, denoted as “high-oxidant pressure” [33]. Oxygen
pressures below the AO stability are referred to as “low-oxidant pressure” [34]. The
possible oxidation modes for this simplified system will be discussed in the following
sections.

22
Literature Review

Alloy Scale Alloy Scale


(1) (2)

Alloy Scale Alloy Scale


(3) (4)

Key
AO
BOv

Alloy Scale
(5)

Figure 2-14: Schematic representation of the modes of oxidation of a single phase


alloy A-B forming insoluble oxides, where A is the more noble and B the more
reactive component.

2.1.3.1 External Scale Formation

The simplest approach to predicting the nature of alloy scales was first outlined by
Wagner [35] and summarised by Gesmundo and Niu [33]. It assumes that the alloy
composition does not change as a result of oxidation. This enables the definition of a
unique composition for which the alloy is in simultaneous equilibrium with each oxide
[33]:

[K (AO)(1 − N )]
e
B
v
= K (BO v ) N Be (2.20)

where K(AO) and K(BO) are the equilibrium constants for the formation of AO and
BOv, and N Be is the mole fraction of B in the alloy, which is here taken to be ideal. If

23
Literature Review

N B is less than N Be , AO only will form, while if it is higher, BOv only will form [35].

As the difference in stability between the two oxides increase, N Be approaches zero.

This limiting situation is approximated by systems where the oxides have very low
scaling rates, so that there is a negligible difference in alloy composition between the
scale/metal interface and the bulk [33]. Wagner [35] noted that in most practical
situations, preferential oxidation of one component leads to its depletion in the
subsurface region and thus changes its concentration at the scale/metal interface. Thus
the quantity N Be appearing in Eqn. (2.20) is actually the alloy surface concentration, not
the bulk value. Figure 2-15 represents schematic concentration profiles existing within a
platinum-nickel alloy after oxidation. Wagner [35] used this system to analyse the
oxidation rate of an alloy containing a noble metal, platinum, and an oxidisable metal,
nickel. He found that the oxidation rate was limited by either the supply of nickel from
the alloy, or by the diffusion rate of nickel through the oxide. As illustrated in Figure
2-15 (b), the oxidation rate was found to approach that for pure nickel as the alloy
nickel content was increased.

24
Literature Review

(a)

(b)

Figure 2-15: (a) Schematic concentration profiles and (b) effect of nickel content
on the parabolic rate constant for the oxidation of Pt-Ni alloys [1].

Wagner [35] was first to develop an analysis taking into account depletion of one
component in a binary alloy. He identified two critical alloy compositions: N Bo∗ (1)

defined the transition from external AO scales to mixtures of AO and BOv; N Bo∗ (2)

defined the transition from mixed scales to external BOv scales. N Bo∗ (1) was calculated

from the corresponding value of N Ao∗ (1) , defined as [35]:

25
Literature Review

[ ]
N Ao∗ (1) = N Ai∗ (1) + 1 − N Ai∗ (1) F (u1 ) (2.21)

where N Ai∗ (1) is the value of NA at the scale/metal interface for the simultaneous

equilibrium with both AO and BOv (i.e. equal to 1 - N Be ) and F(r) is the auxiliary
function defined as:

( )
1
F (r ) = π r exp r 2 erfc(r )
2 (2.22)

The parameter u1 is defined as:

1 k c (AO)
u1 = (2.23)
2 DA

where DA is the alloy diffusion coefficient of A and kc(AO) is the parabolic rate
constant for AO scale growth, defined in terms of the thickness l of metal consumed:

l 2 = k c (AO)t (2.24)

If N Ai∗ (1) is replaced with N Ae and N Ao∗ (1) with N Bo∗ (1) , Eqn. (2.21) simplifies to [33]:

N Bo∗ (1) = N Be [1 − F (u1 )] (2.25)

From Eqn. (2.25), N Bo∗ (1) ≅ N Be when F(u1) << 1, while N Bo∗ (1) is much smaller than

N Be when F(u1) >> 0. Hence, N Bo∗ (1) may be either smaller than or equal to N Be .

The second critical concentration, N Bo∗ (2) , is calculated from [35]:

[ ]
N Bo∗ (2) = N Bi∗ (2) + 1 − N Bi∗ (2) F (u 2 ) (2.26)

where u2 is given by:

26
Literature Review

1 k c (BO v )
u2 = (2.27)
2 DB

Finally, DB is the alloy diffusion coefficient of B and kc(BOv) is the parabolic rate
constant for the growth of BOv, defined in terms of the thickness of metal consumed.
Substituting N Bo∗ (2) = N Be in Eqn. (2.26) yields [33]:

[ ]
N Bo∗ (2) = N Be + 1 − N Be F (u 2 ) (2.28)

Eqn. (2.28) illustrates that N Bo∗ (2) may be higher than or equal to N Be . Hence, N Be is
∗ ∗
always intermediate between N Bo (1) and N Bo (2) . In addition, if the difference in

stability between the two oxides is large, N Bo∗ (1) and N Be will become very small and
practically identical to each other [33].

2.1.3.2 Internal Oxidation beneath an External Scale

The formation of internal oxides of B beneath an external AO scale is not possible for

N Bo less than N Bo (1) [33]. For this condition, the oxygen pressure at the scale/metal
interface is insufficient to form BOv. Since it is expected that both the B and oxygen
concentration decrease with distance from the interface, the alloy will be undersaturated

for formation of BOv. For N Bo between N Bo (1) and N Be , a mixed external AO + BOv
scale will form. However, no internal oxides can form since the scale/metal interface
composition will be fixed at N Be and the alloy is therefore undersaturated with respect
to the formation of BOv. Internal oxides of BOv may only form when the alloy becomes
supersaturated, i.e. for N Bo values sufficiently larger than N Be . Gesmundo and Niu [33]

indicated that a calculation for the minimum value of N Bo required for internal oxidation

of BOv had never been attempted and would be quite difficult. They suggested that N Be
may be a reasonable approximation for this critical value.

27
Literature Review

2.1.3.3 Conditions for the Transition from Internal to External Oxidation

The critical B content required in the bulk alloy for the transition from internal to
∗ ∗
external oxidation of BOv is defined as N Bo (3) . Relationships expressing N Bo (3) for
low-oxidant pressures were analysed by Wagner [36] and Rapp [37], while Gesmundo
and Viani [38] analysed the situation for high-oxidant pressures. These calculations
were based on the requirement that the BOv volume fraction in the internal oxidation
zone (IOZ), fv, be large enough to decrease the oxygen flux sufficiently to favour
external scale formation [36,37]. In a study of the oxidation of Ag-In alloys, Rapp [39]
reported the critical value of fv, f v∗ , be equal to 0.3.

The value of N Bo∗ (3) , for both low and high-oxidant pressures, is calculated to be
[36,37]:

V (alloy)  ∗
N Bo∗ (3) =  m  f v F ( h) (2.29)
 Vm (BO v ) 

where Vm(alloy) and Vm(BOv) are the molar volumes of the alloy and BOv, respectively.
1
The parameter h is equal γφ , where γ is the kinetics parameter defined in the equation:
2

ξ 2 = 4γ 2 DO t (2.30)

where ξ is the thickness of the IOZ at time t and DO is the diffusivity of oxygen in A.
The remaining parameter φ is equal to DO DB .

As summarised by Gesmundo and Niu [33], equations determining values for γ differ
for low and high-oxidant pressure cases. Under low-oxidant pressures, γ can be
calculated from [36]:

N Os G (γ )
= (2.31)
vN B F (h)
o

28
Literature Review

where N Os is the solubility of oxygen in A. For the case of high-oxidant pressures, Eqn.
(2.31) must be modified [38]:

N Os G (γ ) erf (γ ) − erf (u O )
= (2.32)
vN B F (h)
o
erf (γ )

where uO is defined as:

1
1  k (AO)  2 (2.33)
uO =  c 
2  DO 

2.1.3.4 Conditions for the Simultaneous External and Internal Oxidation


of B

A final oxidation mode worth considering involves the simultaneous oxidation of B


both internally and externally. Gesmundo et al. [40] commented that a mathematical
treatment of the kinetics of this process had not been developed. However, the
conditions required for its occurrence were examined by Wagner [41] and summarised
by Gesmundo et al. [40]. This oxidation mode may only occur if the alloy is
supersaturated with respect to the precipitation of BOv within the alloy. Hence, the
oxygen activity within the alloy must be equal to or higher than that required to form
BOv under the local concentration of B [40].

Solution of the diffusion equations for this situation [41] leads to:

−1
 x   1

 k 
 erfc c o 
2
N O = N O erfc
i
 (2.34)
 2 DO t   4D 
 O 
o


where N O is the local oxygen concentration, N Oi its value at the alloy/scale interface, x

the distance from the original alloy surface and DOo the diffusivity of oxygen in the

alloy. The concentration of B in the alloy, NB, is similarly found to be [40]:

29
Literature Review

−1
   1

 
( ) x  k
 
2
N B = N B - N B − N B erfc  erfc
o o i c (2.35)
 2 DB t   4 DB  
 

where N Bi is the concentration of B at the alloy/scale interface.

Wagner [41] used a parameter, defined as the gradient of the logarithm of the
concentration product N B ( N O ) at the alloy/scale interface, to predict the formation of
v

BOv both externally and internally. The parameter, denoted by Gesmundo et al. [40] as
S, can be expressed as [41]:

S=
[
¶ln N B ( N O )
v
] (2.36)
∂x x= X

where X is the location of the alloy/scale interface at time t. From Eqns. (2.34) and
(2.35), S is defined as:

N Bo − N Bi u 2 1 w
S= -v (2.37)
N Bi F (u 2 ) DB t F ( w) DOo t

1
 k (BO )  2
where w =  c o v  . The alloy will be supersaturated with respect to the formation
 4 DO 
of BOv within the alloy for S > 0. Gesmundo et al. [40] commented that a finite degree
of supersaturation would be required for precipitation of BOv within the alloy,
corresponding to a critical value S* larger than 0. However, the authors noted that an
expression did not exist for this critical parameter. Regardless, the determination of S is
still useful as the likelihood of this oxidation mode would be greater for larger values of
S [40].

30
Literature Review

2.1.3.5 The Oxidation of Steel

In general, steels oxidise more slowly than pure iron [12]. This is illustrated in Figure
2-16, which compares the oxidation rate of pure iron and mild steel. However, the
presence of some alloying elements may increase the rate of oxidation. In addition, the
atmosphere in which oxidation takes place can have a significant affect on the rate and
mechanism of oxidation. Each of these issues will be discussed in the following
sections, with particular reference to low carbon steels oxidised in reheating
atmospheres.

250
Weight gain (mg/cm 2)

200

150

100
Pure iron
50
Mild steel
0
0 100 200 300 400
TIme (min)

Figure 2-16: Oxidation of pure iron and mild steel in oxygen at 1100°C [12].

2.1.3.5.1 Effect of Alloying Elements

The oxidation of iron may be significantly modified by the presence of alloying


additions. In general, the scale formed on low-carbon steels is composed mainly of the
oxides of iron. However, depending on the alloying and residual content of the steel, the
following phenomenon can be observed: modification of the morphology of the oxide
layer; enrichment of some elements at the scale/metal interface; modification of the
metal under the scale by selective oxidation of some elements and variations in
oxidation kinetics [42]. The effect on the oxidation behaviour of elements commonly
found in steels will now be discussed. The review will be limited to elements found in
low-alloy steels:

31
Literature Review

Carbon: Malik and Whittle [43] reported that additions of carbon decreased the
oxidation rate of iron. This was attributed to the rejection of carbon at the scale/metal
interface, which modified the behaviour at the interface. Although not observed in their
investigation, it has been found that carbon rejection leads to the formation of a thin
layer of graphite at the scale/metal interface [44]. This may offer a barrier to the flux of
iron and thus decrease the rate of oxidation. Malik and Whittle [43] also discussed the
possibility that carbon may be oxidised at the metal surface, evolving CO and CO2,
promoting a separation between the scale and metal. The authors analysed evolved CO2
during oxidation and concluded that very little carbon was lost to the atmosphere and
that gaseous oxides do not escape through the scale. Earlier studies [45,46] suggested
that appreciable pressures of CO and CO2 led to the formation of cracks in the scale. In
such instances, the oxidation rate may increase as the oxidant gains access to the
underlying metal. Malik and Whittle [43] commented that previous studies almost
exclusively used steels in their investigations, and that the presence of other elements
made interpretation difficult.

Manganese: Manganese can be incorporated into FeO and Fe3O4, however the
effect on the oxidation behaviour is slight [47]. Geneve et al. [142] commented that the
concentration of manganese was the same in the scale and steel for dilute iron alloys.

Chromium: If a sufficient quantity is present, an Fe-Cr spinel may form at the


scale/metal interface [12]. This offers a barrier to the diffusion of iron and thus slows
the rate of oxidation. However, the level of chromium found in low-alloy steel is
usually insufficient to form a continuous layer at the interface and its effect may be
minimal.

Aluminium: Aluminium is usually found in small quantities in low-alloy steels and


thus has negligible effects on oxidation behaviour.

Silicon: The amount of silicon found in steels is generally below that required
to form a protective scale of SiO2 [48]. However, diffusion of oxygen into the metal
may result in the internal oxidation of silicon to form SiO2. The formation of a subscale
at the scale/metal interface may result when SiO2 reacts with FeO to form fayalite,
Fe2SiO4. Fayalite has a melting temperature of 1205°C and may form the eutectics:
Fe2SiO4-FeO (at 1177°C) or Fe2SiO4-SiO2 (at 1178°C) [48]. The formation of a liquid

32
Literature Review

oxide phase increases the oxidation rate [48,49] due to increased diffusivity of iron from
the metal through the scale. In contrast, the formation of a solid silicon-rich oxide phase
at the scale/metal interface has been suggested to inhibit the diffusion of iron and thus
decreases the oxidation rate [50].

Nickel: Nickel is more noble than iron and its concentration at the scale/metal
interface will increase as iron is consumed. The diffusivity of nickel in iron is slow
compared to the motion of the scale metal interface. This will result in an enrichment of
nickel at the surface of the steel. The nickel-enriched surface region is more resistant to
oxidation and scale growth occurs preferentially along grain boundaries [51]. Nickel
does not significantly decrease the oxidation rate, as seen from the oxidation kinetics of
Fe-Ni alloys with up to 30 at%Ni, which varied only slightly with composition
[27,47,52].

Copper: As with nickel, copper will enrich at the scale/metal interface during
oxidation. In contrast, copper has a limited solubility in iron within the temperature
range of reheating practices. If sufficient enrichment has occurred, a copper-rich phase
will precipitate at the scale/metal interface. Nanni and Gesmundo [53] investigated the
oxidation of an Fe-4.54wt%Cu alloy in oxygen between 700 and 1000°C. They found
that the oxidation rate was significantly decreased with respect to pure iron. The authors
suggested a number of reasons for this observation. Firstly, the oxygen activity at the
alloy/scale interface increases when a copper-rich phase is present. This change
decreases the oxidation rate of FeO, which subsequently lowers the overall rate for the
alloy. Secondly, an observed fine distribution of dissolved copper within the scale
would also lead to a rate decrease. The presence of some foreign ions dissolved in
magnetite has been shown to decrease iron diffusivity [54], while the doping of iron
[55] and iron-chromium [56] alloys with lithium has been reported to decrease the
defect concentration in the scale. Finally, the presence of a second metal phase may act
as a barrier to the flux of iron into the scale.

The copper-rich phase is molten above 1096°C [57]. The effect of a liquid copper-rich
phase on the oxidation behaviour of iron or steels has received little attention. Houpert
et al. [58] measured the weight gain of various copper-containing low carbon steels in a
simulated reheating furnace atmosphere under non-isothermal conditions. The weight
gains were similar up to a temperature of 1100°C, after which the steels with increasing
33
Literature Review

copper content oxidised more rapidly. It was thought that a liquid film at the scale/metal
interface increased the oxidation rate by increasing the flux of iron into the scale.

2.1.3.5.2 Effect of Oxidant Atmosphere

The high temperature oxidation behaviour of iron has been studied extensively. The
bulk of the investigations have been carried out in low oxygen potential atmospheres,
e.g. CO-CO2 [19,20,59], CO2 [21] and H2O-H2 [22]. Oxidation was found to be initially
controlled by either diffusion in the gas phase or reaction at the scale/gas interface.
Once a critical scale thickness was met, reported as between 4×10-3 and 0.1mm by Petit
and Wagner [59] (700 – 1100°C), oxidation became controlled by diffusion through the
scale. Table 2-1 summarises the linear rate constants for oxidation of iron or low
carbon steel in low oxygen potential atmospheres. The rates of oxidation in H2O-
containing atmospheres are generally higher than those containing CO2 or CO. This is
most likely a result of the different gases having different rates of chemisorption of
oxygen ions at the scale surface [12].

Table 2-1: Reported linear rate constants (g/cm2s) in low oxygen potential
atmospheres

Temperature Iron in 1:1 Iron in CO2 [21] Low carbon Low carbon
(°C) (vol%) CO2 – steel in 10% steel in 10%
CO [19] CO2 - N2 [60] H2O – N2 [60]
1150 8.90×10-6 9.70×10-6
1100 9.95×10-7 3.50×10-6 8.72×10-6
1050 7.40×10-7 1.12×10-6 5.53×10-6
1000 2.83×10-8 5.38×10-7 1.80×10-7 1.80×10-7
900 1.73×10-8 2.62×10-7 4.00×10-8

Abuluwefa et al. [61] commented that the number of studies involving oxidation in
oxygen environments is limited. These authors investigated the oxidation behaviour of
low carbon steel between 1000 and 1250°C in O2-N2 atmospheres with oxygen contents
ranging between 1 and 15%. Initial oxidation rates were found to follow a linear
oxidation rate law up to scale thicknesses of between 400 and 500µm. The linear rate
constant was found to be dependent upon the oxygen content. As indicated in Table 2-2,

34
Literature Review

the rate constant increased with oxygen content. Abuluwefa et al. [61] concluded that
the transport of oxygen through the gas-phase boundary layer was rate controlling.

Table 2-2: Linear rate constants (g/cm2s) for oxidation of low carbon steel in O2-N2
[61]

Percent oxygen
Temperature (°C)
1 3 6 9 12
1250 1.43×10-5 3.88×10-5 8.65×10-5 1.58×10-4
1200 1.53×10-5 3.50×10-5 7.83×10-5 1.11×10-4 1.37×10-4
1150 1.20×10-5 3.25×10-5 8.30×10-5 1.07×10-4 1.36×10-4
1100 1.45×10-5 2.95×10-5 7.28×10-5 1.15×10-4 1.54×10-4
1050 1.55×10-5 2.92×10-5 7.25×10-5 1.18×10-4 1.66×10-4
1000 1.17×10-5 3.12×10-5 6.42×10-5 1.18×10-4 1.46×10-4

When iron or steel is oxidised in air or oxygen at high temperature, linear kinetics may
not be perceived since diffusion control has taken over by the time reliable experimental
readings can be taken [12]. Once a thick enough scale forms, parabolic kinetics will be
observed as diffusion through the scale becomes rate controlling. Table 2-3 summarises
the reported parabolic rate constants for oxidation of iron or steel in high oxygen
potential atmospheres. These values illustrate that iron oxidises more quickly than steel
and that oxidation in oxygen is more rapid than in air.

In conventional steelmaking, slabs or billets are reheated prior to final rolling. It is


desirable to minimise the amount of scale formed in order to maximise steel yield.
However, some scaling is required to remove surface defects and oscillation marks
produced during continuous casting [62]. Reheating is generally achieved by the
combustion of natural gas, fuel oil, coke oven or blast furnace gas [63]. Steel is often
charged cold and is heated continuously as it moves through the reheat furnace, exiting
at between 1230 and 1250°C [62]. During reheating, oxidation takes place in the
gaseous products of combustion, i.e. free O2 (from the addition of excess air), CO2, H2O
and N2 [60]. The relative amounts of these gases depend on the type of fuel and the
percent combustion air used. Cook and Rasnussen [64] investigated the equilibrium
products of the combustion of natural gas. Figure 2-17 summarises their results with
respect to the percentage of theoretical combustion air. Stoichiometric combustion,
35
Literature Review

referred to as 100% air, refers to the amount of air required for complete combustion.
As the percent combustion air decreases, the amount of oxidants H2O and CO2 decrease,
while the reducers CO and H2 increase. Hence, the atmosphere changes from an
oxidising to a reducing environment.

Table 2-3: Reported parabolic rate constants (g2/cm4s) for oxidation of iron or steel
in high oxygen potential atmospheres

Temperature Iron in air Iron in Low carbon steel Mild steel in Mild steel in
(°C) [65] oxygen [65] in 6% O2-N2 [61] air [65] oxygen [65]

1250 3.63×10-6
1200 2.60×10-6
1150 2.12×10-6
1100 2.82×10-6 3.54×10-6 1.25×10-6 4.17×10-7 2.00×10-6
1050 1.97×10-6 2.16×10-6 9.10×10-7 2.63×10-7 1.11×10-6
1000 8.31×10-7 1.14×10-6 4.90×10-7 1.59×10-7 5.76×10-7
900 1.87×10-7 2.24×10-7 5.25×10-8 1.51×10-7

Minaev et al. [66] oxidised various carbon steels in the combustion products from
natural gas, coke oven gas and blast furnace gas, burned with air at 0.7 to 1.1 of
stoichiometric gas requirements between 800 and 1250°C. They reported that oxidation
was most rapid in coke oven gas combustion products and least rapid in blast furnace
gas products. Oxidation in natural gas was only slightly higher than that in blast furnace
gas. Lee [67] oxidised a low carbon steel in a simulated reheat furnace atmosphere and
varied the percentage theoretical combustion air (see Figure 2-18). The scaling rate was
much lower in the sub-stoichiometric atmospheres (i.e. 95 and 99% air). In the 112% air
atmosphere, the oxidation kinetics was parabolic. In contrast, linear kinetics was
observed in the weakly oxidising, sub-stoichiometric atmospheres.

36
Literature Review

Figure 2-17: Equilibrium products of combustion of natural gas related to


percentage theoretical combustion air [64].

60

50
Weight gain (mg/cm 2)

40
112%
30 99%
95%
20

10

0
0 10 20 30 40 50 60

Time (minutes)

Figure 2-18: Oxidation of low carbon steel (A1006) in a simulated reheat


atmosphere (natural gas) of various percentage theoretical combustion air at
1100°C [67].

37
Literature Review

Selenz and Oeters [68] investigated the oxidation of Armco iron in binary, ternary and
quaternary gases containing 0-3% O2, 10% CO2, 10 and 20% H2O and N2 in the
temperature range 1000 to 1300°C. They reported that the linear rate constant in
multicomponent gas mixtures was the sum of the constants for the single oxidising
components. The parabolic rate constant in multicomponent gas mixtures was found to
be partially dependent on the gas composition. This is expected if the scale contains
FeO only; however, the gas mixtures used had an oxygen potential far higher than the
value for the FeO-Fe3O4 equilibrium. Selenz and Oeters [68] observed that only traces
of Fe3O4 were formed in oxygen-containing atmospheres. It was concluded that gas-
solid equilibrium was not met at the scale surface and the oxygen potential in the gas
boundary layer dropped below that of the FeO-Fe3O4 equilibrium.

Abuluwefa et al. [60] studied the oxidation of low carbon steel in atmospheres
containing O2, CO2, H2O and N2 in the temperature range 800 to 1150°C. The authors
found that temperature and the concentration of free oxygen were the two most
important parameters affecting oxidation rates. The presence of CO2 and H2O at
concentration levels typical of reheat furnace atmospheres contributed little to the linear
oxidation rate in atmospheres containing oxygen levels of 6%. Abuluwefa et al. [60]
experimental results supported the finding of Selenz and Oeters [68] that the linear
oxidation rate in a multicomponent atmosphere may be estimated by the addition of
rates of oxidation for the individual gaseous components. Reported linear rate constants
for multicomponent atmospheres are summarised in Table 2-4. The rate constants are
slightly higher than the respective values in O2-N2 atmospheres with equivalent O2
concentrations (see Table 2-2). Parabolic rate constants determined graphically from
data reported by Abuluwefa et al. [62] are summarised in Table 2-5. These values are
similar to those reported for oxidation in oxygen or air (see Table 2-3).

38
Literature Review

Table 2-4: Reported linear rate constants (g/cm2s) for oxidation of low carbon steel
in multicomponent atmospheres [60]

Temperature (°C) 1%O2-10%CO2-3%H2O-N2 6%O2-6%CO2-3%H2O-N2


1150 4.13×10-5 1.27×10-4
1100 5.22×10-5 8.55×10-5
1050 3.27×10-5 8.40×10-5
1000 3.10×10-5 8.33×10-5
900 2.30×10-5 5.18×10-5

Table 2-5: Reported parabolic rate constants (g2/cm4s) for oxidation of low carbon
steel in multicomponent atmospheres [62]

Temperature (°C) 1%O2-10%CO2-3%H2O-N2 6%O2-6%CO2-3%H2O-N2


1150 2.88×10-6 3.74×10-6
1100 1.74×10-6 3.08×10-6
1050 1.16×10-6 1.84×10-6
1000 1.08×10-6 1.22×10-6
900 5.03×10-7 5.76×10-7

Abuluwefa et al. [62] examined the amount of scale formed on samples of low carbon
steel that were placed on top of steel slabs being processed in an industrial walking-
beam reheat furnace. Samples were retrieved at various stages and the amount of scale
growth determined. The furnace was operated using low and high air/fuel ratios, for
which the average measured composition was 1.5%O2-12%CO2-12%H2O-N2 and
3%O2-7%CO2-12%H2O-N2, respectively. The weight gain results are shown in Figure
2-19 (a). The temperature profile of the samples during reheating was measured and is
identified as ‘Measured rider top surface’ in Figure 2-19 (b). As a comparison,
Abuluwefa et al. [62] carried out nonisothermal laboratory experiments in which the
weight gain of a low carbon steel sample was measured continuously. Samples were
introduced into the furnace at 700°C and heated at 5.6 or 9.6°C/min to 1150°C and held
at this temperature for 30 minutes. The results are summarised in Figure 2-20. The
laboratory results indicate a decrease in the oxidation rate at approximately 900°C.
Similar observations were made by Sachs and Tuck [65] who attributed the change in
rate to scale separation. Steel transforms from ferrite to austenite at this temperature and

39
Literature Review

it was believed that the dimensional change in the metal caused scale/metal separation.
The phenomenon was thought to be more likely for scales with an outer layer of Fe3O4
or Fe2O3, as these oxides are comparatively more rigid than FeO and hence decrease the
overall scale plasticity [65].

(a) (b)

Figure 2-19: (a) Weight gain curves from oxidation of low carbon steel during
reheating in an industrial walking-beam reheat furnace, and (b) measured
temperatures of slab and sample (rider) during a reheating cycle in an industrial
walking-beam reheat furnace [62].

40
Literature Review

Figure 2-20: Weight gains from nonisothermal oxidation of low carbon steel in the
low oxygen atmosphere for two heating rates. The top graph shows the
temperature profile during oxidation [62].

2.2 The Phenomenon of Surface Hot Shortness

2.2.1 Residual Elements in Steels

Residual (or tramp) elements are metallic or metalloid elements that are not efficiently
removed from the liquid metal during steelmaking and as a result may build up to
relatively high levels with continued recycling [69]. Such elements are described as
residual because they are always present in steel to some degree. Residual elements are
also noted for the significant effects that they may have on particular critical properties
of steel at levels of only a few hundredths to a few tenths of a percent [69].

The EAF steelmaking process uses scrap as its primary feed source [70,71]. Due to the
nature of recycling, high levels of residuals are found in scrap. For example, copper can

41
Literature Review

enter through the recycling of brass and electrical wiring, whereas nickel and tin can be
introduced through plated material [70]. Because scrap has higher levels of residuals, it
follows that EAF steel will also contain higher levels. A comparison of the typical
residual levels between EAF steel and an equivalent grade produced by the basic
oxygen furnace (BOF) process is summarised in Table 2-6.

Table 2-6: Typical residual levels (in wt%) in EAF and BOF low carbon steels
currently produced by BHP Steel [72]

Element BOF Steel EAF Steel


Copper 0.01 0.20
Nickel 0.005 0.08
Tin 0.003 0.016

Elements such as Cu, Ni, Sn, Sb, Pb, As, S and P remain in steel because they cannot be
preferentially oxidised during normal steelmaking processes [70,71,73]. The same
thermodynamics that prevent removal of residuals by oxidation during steel production
also lead to their rejection from and oxide scale during secondary processing. This can
be illustrated by use of the Ellingham diagram shown Figure 2-21, in which the Gibbs
free energies (in units of kJ per mole O2) of the relevant metal oxides are plotted as a
function of temperature. In this figure, the free energy for only the most stable iron
oxide, FeO, is shown. Consider a steel billet (containing Cu, Ni and Sn) being reheated
prior to a hot working process, such as rolling to a rod or sheet. The typical temperature
range used in BHP reheat furnaces is 1100 - 1250°C, for times varying from 1 to 3.5
hours [74]. When the steel billet is reheated in an oxidising atmosphere, iron will
oxidise to form a scale consisting primarily of FeO. Copper and other elements whose
oxides are less stable than FeO (i.e. oxides which are above FeO in the Ellingham
diagram) will not oxidise. These elements will become enriched in the subsurface
region of the steel. For example, copper enrichments of 5 to 10 times the bulk
concentration have been reported within the enriched zone [75]. The residual elements
can build up to such levels that the properties of the steel can be adversely affected [69].

42
Literature Review

-100

4Cu + O2 = 2Cu2O
-200 2Ni + O2 = 2NiO
∆G° (kJ/mole O2)

-300 Sn + O2 = SnO2

2H2 + O2 = 2H2O 2Fe + O2 = 2FeO


-400
C + O2 = CO2

-500 2C + O2 = 2CO

-600 Si + O2 = SiO2

-700
500 600 700 800 900 1000 1100 1200 1300

Temperature (°C)

Figure 2-21: Ellingham diagram for various oxides.

According to Noro et al. [76], scrap can be classified into home scrap, market scrap and
imported scrap. Market scrap can be further classified into process scrap (scrap resulting
from customers steel fabrication processes) and obsolete scrap (scrap resulting from
discarded steel products after their service life ends). Table 2-7 lists the composition of
typical market scrap from Japan. As can be observed from this table, the values of
copper and tin in obsolete scrap are an order of magnitude higher than that found in pig
iron scrap. Noro et al. [76] stated that due to the higher contents of residual elements in
obsolete scrap, products manufactured from EAF steel contain levels of copper and tin
approximately one order of magnitude higher than those produced from raw products or
using higher quality scrap.

43
Literature Review

Table 2-7: Chemical compositions of residual elements of typical market scrap in


Japan [76]

Chemical compositions after melting (%)


Category of market scrap
Cu Sn Ni Cr Zn*
Obsolete Car shredded scrap 0.230 0.052 0.069 0.123 0.050
Market scrap

Heavy scrap 0.234 0.017 0.070 0.130 0.210


Can scrap 0.050 0.128 0.032 0.061 0.000
Factory bundle 0.027 0.002 0.020 0.031 0.700
Home scrap 0.021 0.010 0.050 0.030 0.010
Pig iron scrap 0.010 0.002 0.020 0.020 0.002
* Before melting

To be able to use poorer quality scrap in the EAF process, high levels of residuals are
diluted by blending higher quality scrap (e.g. high purity home scrap or process scrap)
with pig iron and obsolete scrap [76]. Table 2-8 illustrates the composition of melts
when scrap is mixed with sponge iron (direct reduced iron or DRI) or hot metal. The
extent of dilution depends on the scrap being used and the steel being produced (and
hence the allowable levels of residual elements in the final product). However, the
limiting factor to this approach is the increase in production costs associated with using
higher quality scrap or increased levels of iron produced from raw products. As a result,
there is a need to be able to make use of poorer quality scrap for EAF steel production
[77].

Table 2-8: Residual element composition of various charge types [78]

Charge Cu Sn Ni S
100% scrap 0.2 0.015 0.07 0.03
60% DRI-40% scrap 0.1 0.010 0.05 0.02
100% DRI 0.02 0.002 0.01 0.01
BOF 70% hot metal-30% scrap 0.04 0.005 0.02 0.03

Furthermore, as Noro et al. [76] point out, obsolete scrap accounts for the majority of
scrap available for recycling, and its occurrence is increasing every year (see Figure
2-22). As many authors have indicated, the concentration of residual elements in scrap
is expected to rise further due to increased recycling as a consequence of environmental

44
Literature Review

and economic pressures [71,76,79]. Estimated trends in copper and tin contents are
shown in Figure 2-23. From these estimations, it is apparent that by 2015 the residual
element contents of obsolete scrap will be 1.2 - 1.5 times the present values. Hence it is
apparent that poorer quality scrap will be used for more EAF steel products in the future
and thus methods must be developed to be able to make use of such scrap. Marique [80]
commented that there is no technique available to remove residual elements from steel
melts. As a result, scrap preparation and ferrous/non-ferrous separation techniques are
gaining attention [58,80]. However, Emi et al. [81] stated that an economically viable
process for completely separating residual elements in scrap has not been developed.
Therefore, methods must be developed so that steels containing high levels of residual
elements can be processed and utilised.

70
Estimate
Import
60 Obsolete
Scrap generated (million tons/year)

Process
Home scrap
50

40

30

20

10

0
'80 '81 '82 '83 '84 '85 '86 '87 '88 '89 '90 '91 '92 '95 2000 2005 2010

Fiscal year

Figure 2-22: Change in the generation of scrap in Japan [76].

45
Literature Review

0.5 0.03
Bar steel scrap
Shaped steel scrap
Heavy scrap (high grade)
0.4
Concentration of Cu (%)

Concentration of Sn (%)
0.02

0.3

0.2
0.01

0.1

0.0 0.00
1985 1990 1995 2000 2005 2010 2015 2020 1985 1990 1995 2000 2005 2010 2015 2020
Fiscal Year Fiscal Year
Figure 2-23: Simulation of copper and tin contents for obsolete scrap in the future
[76].

Metallurgical phenomena related to residual elements are well documented


[58,69,80,82]. These include surface hot shortness, intergranular brittleness and changes
to the mechanical properties. As will be discussed later, surface hot shortness is a form
of surface cracking caused by the formation of a liquid phase during oxidation and hot
working. Intergranular brittleness is related to the segregation of elements to grain
boundaries and the subsequent precipitation of second phases that inhibit
recrystallisation and the mobility of austenite boundaries during deformation [58,70].
Residual elements in solid solution may increase hardness and decrease ductility
[58,69]. Copper may precipitate during aging treatments to strengthen by precipitation
hardening [58]. Such changes in mechanical behaviour may be beneficial or detrimental
depending on the product application.

Residual elements have also been associated with adherent scale, a type of oxide that is
difficult to remove during hot working. Scale that becomes rolled into the surface of the
workpiece may cause surface defects and excessive tool and die wear [83]. For
example, nickel has been shown to decrease the descalability of low carbon steel [84].
Asai et al. [84] observed that adherence increased as the scale/metal interface became
more uneven with increasing nickel additions. Hence, the uneven interface caused a

46
Literature Review

mechanical locking between the scale and metal. A similar mechanism was observed for
steels containing silicon, on which the formation of fayalite increases the scale/metal
interface unevenness [85,86].

2.2.2 Mechanism of Surface Hot Shortness

2.2.2.1 Introduction

Melford [70] categorised two forms of hot shortness, (i) hot shortness in the bulk of the
steel (liquation cracking) and (ii) surface hot shortness. Liquation cracking is a problem
associated with relatively insoluble residual elements (e.g. sulphur) enriching at grain
boundaries, precipitating as a liquid and resulting in cracking. Surface hot shortness is
caused by the enrichment of residual elements during oxidation that can give rise to a
liquid phase, which subsequently weakens austenite grain boundaries. The effect can be
observed by the appearance of surface cracks or fissures (see Figure 2-24). In many
cases, this renders the product unacceptable and can lead to the scrapping of a large
amount of product. The focus of this investigation is surface hot shortness and no
further attention will be given to liquation cracking.

47
Literature Review

1cm

Figure 2-24: Mild steel billet exhibiting surface hot shortness [87].

The basic mechanism of surface hot shortness is well known [87-89] and is represented
schematically in Figure 2-25. When a steel containing copper is oxidised, FeO is formed
due to its relatively high stability [88,89]. The preferential oxidation of iron leads to
enrichment of copper and other less reactive residuals, such as nickel, tin, antimony and
arsenic, at the steel surface. When the enrichment of copper reaches a level exceeding
the solubility limit in iron, a copper-rich phase precipitates at the scale/metal interface.
The iron-copper phase diagram in Figure 2-26 indicates that the solubility of copper in
austenite (γ) at 1096°C is 8.2 wt%. Melford [70] reported that when only copper is
present in steel, an enrichment level of over 9 wt% copper is quite possible under
conditions of severe oxidation in the temperature range 1100-1200°C.

48
Literature Review

Scale

O2 Steel O2 Steel
Concentration

Concentration
(a) Distance (b) Distance

Scale Copper-rich phase

O2 Steel
Concentration

Concentration profile key


Iron
Copper

(c) Distance

Figure 2-25: Schematic of copper enrichment during oxidation; (a) before


oxidation, (b) initial stages of oxidation and (c) later stages of oxidation where the
solubility of copper in the metal has been exceeded.

49
Literature Review

1600
1583°C
1500 12.9 Liq.
δ
8.1
1400
1394°C
1300
γ
Temperature °C

1200
1096°C 96.7
1100
8.2 95.9
1000 1084.87°C
912°C
900
ε
3.1 850°C
800 2.2
α
700

600
0 10 20 30 40 50 60 70 80 90 100
Weight Percent Copper

Figure 2-26: Iron-copper phase diagram [57].

It should be noted that where surface hot shortness is concerned, copper is the crucial
element, because it is the only residual element of significance that can form a molten
phase at reheating temperatures [70,82,87]. Once a liquid copper-rich phase has formed
(i.e. temperatures above 1096°C, as indicated from Figure 2-26), it can penetrate along
austenite grain boundaries. Many authors have concluded that this penetration weakens
the boundaries and can cause surface cracks in the presence of stresses developed
during hot working [70,82,87,90]. An example illustrating the penetration of a copper-
rich phase and crack formation is given in Figure 2-27. If sufficient weakening of the
boundaries has occurred, surface cracks will form a checked or alligator surface [87].
Cracks may also manifest themselves in other forms. Nicholson and Murray [88]
reported that black arrow-shaped marks observed on cold rolled product tended to lift
and detach when rolled at high reductions. These defects were attributed to the
development of small cracks during the roughing stage of secondary hot rolling. Van
Wijngaarden et al. [91] reported that observed seams in rolled bar initiated when tearing
of billets occurred during the first two passes in the roughing mill. Habraken and
Lecomte-Beckers [92] stated that small surface cracks caused by surface hot shortness

50
Literature Review

result in the formation of ‘slivers’ of metal or tears on the surface during rolling of
slabs.

Figure 2-27: Photomicrograph of the surface of a reheated billet, showing


penetration of residual-rich phase down austenite grain boundaries [71].

Residual element enrichment not only occurs during reheating but also during casting.
Indeed, surface hot shortness cracking has been reported to develop during continuous
casting [91,93-95]. Wolf [93] reported that cracks on continuously cast billets were
fairly shallow, having an average and maximum depth of 0.5mm and 2mm,
respectively. Such cracking would not pose a problem in ordinary steelmaking (e.g.
rebar, structural steels) but would be a concern in special steelmaking (e.g. forging
billets, rails, etc.) [93]. Houpert et al. [58] reported that reheating prior to secondary
rolling could be used to remove cracking developed during continuous casting. They
found that longer hold times in the reheat furnace reduced cracking associated with
continuous casting. However, prolonged hold times gave rise to cracking developed
during secondary rolling.

The embrittlement of steel by liquid copper has been reported in other areas. For
example, copper is known to embrittle iron-base alloys during welding [96-98]. Surface
contamination of the base metal by copper may be introduced from accidental abrasions

51
Literature Review

of copper weld tooling [97] or by the use of copper-base alloys during brazing [96]. The
copper contamination melts during welding and under an applied stress may cause
surface cracks. Vander Voort [99] reviewed locomotive axle failures that had been
attributed to liquid copper embrittlement. In these cases, overheating caused bronze
sleeves or bearings to melt. A liquid copper phase was then available to penetrate the
steel axles, resulting in embrittlement and failure. Fisher et al. [100] investigated
cracking on the bore surface of cannon tubes. The authors found that copper was
transferred by abrasive contact between the steel surface and the copper rotating band
on the projectile. Cracks developed when copper embrittled the steel during the thermal-
mechanical stresses generated during firing. It is thus obvious that under particular
conditions, steel is susceptible to defects arising from liquid copper embrittlement.

2.2.2.2 Copper-rich phase formation

For a steel to be prone to surface hot shortness, a copper-rich phase must first precipitate
at the scale/metal interface. The ease with which this is achieved depends on the copper
content in the steel, oxidation rate and the rate of copper ‘back-diffusion’ from the
surface to the interior [101]. The rate of oxidation and back-diffusion are both diffusion-
controlled mechanisms and thus affected by temperature. Hence it is not surprising that
copper-rich phase formation has been observed to be temperature dependent [88,89]. In
general:

• At temperatures below approximately 1000°C, the copper-rich phase is solid and


appears as a dispersion of globules entrapped in the inner oxide layer or at the
scale/metal interface (see Figure 2-28). Surface hot shortness will not occur at these
temperatures [69,101,102], as a liquid phase is required to penetrate and weaken
grain boundaries. The melting point of the copper-rich phase predicted from the
iron-copper phase diagram is 1096°C and surface hot shortness may be expected
above approximately 1100°C. However, the melting point is affected by the
presence of other elements. For example, it has been reported that a copper-rich
phase may remain molten at temperatures as low as 900°C when tin is present
[70,103].

52
Literature Review

Figure 2-28: Solid copper-rich phase particles dispersed in the inner scale layer
(0.108%C-3.04%Cu-Fe alloy oxidised in air for 24h) [104].

• In an intermediate temperature range of approximately 1100 - 1150°C, the copper-


rich phase is molten and usually remains at the scale/substrate interface (see Figure
2-29). Under these conditions the copper-rich phase may penetrate metal grain
boundaries, leading to their embrittlement during hot working.

• For temperatures above approximately 1200°C, the copper-rich phase is molten but
mainly occluded into the scale layer (see Figure 2-30). It has been proposed [88]
that occlusion (or separation into the scale) is the result of internal oxidation along
grain boundaries, resulting in engulfment of grains, or by formation of a layer of
oxide particles at the copper-rich phase/matrix interface (leading to undermining of
the phase). Surface hot shortness is avoided in this situation since the molten
copper-rich phase is isolated from the substrate.

53
Literature Review

Figure 2-29: Experimental steel containing 3% copper oxidised for 3 hours at


1100°C (oxidant unknown) [105].

Scale

Metal
50µm

Figure 2-30: Metallic globules occluded into scale from the oxidation of Ni/Cu-
containing steel [106].

54
Literature Review

The important points to note from these observations is that surface hot shortness only
occurs when the copper-rich phase is molten and in contact with the austenite grain
boundaries. Hence for surface hot shortness to occur, the temperature of reheating/hot
working must be such that the copper-rich phase is molten, but not occluded harmlessly
into the scale. Indeed, cracking test investigations [102,107] have been carried out
which illustrate that hot shortness does not occur at or around 1000°C, and there is also
a great deal of evidence that occlusion inhibits surface cracking [73,89,107].

When copper is enriched at the metal surface, a concentration gradient is established


and diffusion takes place from the surface to the interior [82]. Melford [87] proposed
that there is a balance between the level of enrichment and diffusion of copper, and
above some temperature, diffusion is so rapid that the amount of enrichment at the
surface never exceeds the solubility limit in austenite. As a result surface hot shortness
would not occur. Therefore, it is conceivable that this mechanism operates in addition to
the three modes of copper-rich phase formation discussed above. Melford [87]
calculated the rate of copper enrichment and back-diffusion for a mild steel containing
0.27% copper (Figure 2-31). Back-diffusion was predicted to become significant for
temperature greater than 1080°C, which was in poor agreement with experience. The
author stated that the affect of other residual elements had not been considered in the
calculation and that more accurate diffusivity data was required to properly model this
process.

55
Literature Review

Figure 2-31: Comparison of enrichment and dispersion rates for copper in iron.
Experimental data for the oxidation of a mild steel containing 0.27%Cu were used
to provide data on the rate of copper enrichment [87].

2.2.2.3 Penetration and Cracking

2.2.2.3.1 Liquid Metal Embrittlement and Wettability

The weakening and subsequent crack formation along austenite grain boundaries by the
penetration of a liquid copper-rich phase can be considered a form of liquid metal
embrittlement (LME) [49,107-110]. LME describes the phenomenon whereby the
ductility or fracture stress of a solid metal is reduced by exposure of its surface to a
liquid metal [111]. Vander Voort [99] noted that some confusion exists in the literature
regarding LME and that this was partly a result of the different forms that exist. Stoloff
[111] classified four distinct forms:

1. Instantaneous fracture of a particular metal under an applied or residual stress when


in contact with particular liquid metals. This is the most common form of LME.

56
Literature Review

2. Delayed failure of a particular metal in contact with a particular liquid metal after a
certain time interval at a static load below the ultimate tensile stress of the metal.
This form results from grain boundary penetration by the liquid metal and is less
common than the first form.
3. Grain boundary penetration of a particular solid metal by a particular liquid metal
causing eventual disintegration of the metal. A stress is not a prerequisite for this
form of LME.
4. High temperature corrosion of a solid metal by a liquid metal resulting in
embrittlement. This is entirely different to the other forms of LME.

It is apparent that surface hot shortness falls into the fourth classification. Vander Voort
[99] commented that studies of LME of this form were less common than others.
Indeed, Savage et al. [97] stated that many studies of LME consider systems where a
large amount of liquid metal is in contact with a solid base metal. These authors
proposed that minute quantities of particular metals with melting points below the
matrix material can embrittle the surface through LME [97]. Certainly, this would need
to be the case if LME was the mechanism responsible for cracking due to surface hot
shortness.

The following conditions are required but not necessarily sufficient for LME to occur
between a solid-liquid metal couple [111]:

(i) adequate wetting of the solid by the liquid


(ii) the presence of an applied or residual stress in the solid
(iii) a barrier to plastic flow in the base metal at some point in contact with the liquid.

When a steel prone to surface hot shortness is reheated and worked, it appears that all
three conditions for LME are satisfied. In view of the varying forms of LME, care must
be taken in comparing results from specific LME studies. A number of experimentally-
derived conditions have been identified that are usually, but not always, obeyed [99].
For example, solid metals that are highly soluble in the liquid metal and solid/liquid
systems that form intermetallic compounds are usually immune to LME. When a
material fails due to LME, the crack path is generally intergranular [111]. In addition,

57
Literature Review

the fracture stress and ductility are extremely sensitive to temperature, strain rate, grain
size, thermo-mechanical history and composition of the solid and liquid.

LME is believed to result from the interaction between the liquid metal atoms and the
metal atoms that define the crack tip [98]. The presence of a liquid phase can cause
normally ductile metals to fail in a completely brittle fashion. When a continuous
supply of liquid is available, the crack tip is kept atomically sharp which allows the
stress to concentrate to such a level where the crack continually propagates.

Hough and Rolls published a number of investigations concerning the LME of iron by
copper [112-115]. These authors performed notched tensile-creep tests between 1100
and 1130°C under non-oxidising conditions. Copper was introduced into the notch by
coating a thin layer using a cyanide bath or by winding pure copper wire about the
notch surface. Copper was found to significantly alter the creep behaviour of iron and
cause premature rupture. Embrittlement was found to be of the delayed failure type,
resulting from diffusion-controlled, grain boundary penetration by copper. A copper-
rich phase was found ahead of the cracks and the depth of cracking was found to be
dependent on the depth of copper penetration.

Suzuki [109] measured the hot ductility of low carbon steel samples whose surface had
been coated with copper. The hot tension tests were conducted between 900 and 1070°C
for strain rates varying between 5×10-4 to 50s-1. The hot ductility decreased markedly at
1050°C and higher temperatures. Under these conditions copper was observed to have
penetrated grain boundaries and failure occurred by intergranular fracture. However, the
hot ductility was found to return with increasing strain rate (see Figure 2-32). For
example, excellent hot ductility was reported for all temperatures at a strain rate of 50 s-
1
. The fracture mode was observed to change from intergranular to transgranular at the
higher strain rates. Transgranular fracture was also observed for temperatures at which
the copper layer was solid. Suzuki [109] concluded that transgranular fracture was not
related to LME. It was proposed that for temperatures where LME by liquid copper may
occur, there exists a strain rate where deformation in the matrix is faster than the
penetration of liquid copper along grain boundaries. Seo et al. [116] reported a similar
strain dependence for which embrittlement by liquid copper was not observed above a
strain rate of 3×10-2 s-1.

58
Literature Review

Figure 2-32: Strain rate dependence of hot ductility for various test temperatures
for a low carbon steel plated with copper [109]. RA(%) refers to the percentage
reduction of area to failure.

It is pertinent now to introduce the term wetting. This can be described as a


phenomenon involving a solid and liquid that are in intimate contact, such that the
adhesive force between the two phases is greater than the cohesive force within the
liquid [117]. In the situation where a liquid copper-rich phase is in contact with
austenite grains, the liquid phase wets the austenite grain boundaries in what is known
as intergranular penetration. Smith [118] predicted that the penetration by a liquid phase
increases as the dihedral angle (θ) between the liquid and solid phase decreases. Figure
2-33 schematically represents the situation that arises when a liquid copper-rich phase is
present at the triple point of austenite grain boundaries. Smith [118] proposed that the
shape of the liquid phase adjusts itself such that the surface energy of the system is
minimised. As a result, θ could be predicted by a balance of the surface free energies of
the system, (or in this case the surface tensions) by balancing the surface tension forces
at the junction of the grain boundary, OC, with that of the two interphase boundaries,
OA and OB (see Figure 2-33). If the surface free energy of the austenite-austenite grain
boundary (i.e. OC) is denoted by γFe-Fe and the surface free energy of the austenite-
liquid phase boundary (i.e. OA and OB) is denoted by γFe-Cu, then the dihedral angle can
be determined from the equation [118]:

59
Literature Review

θ
γ Fe− Fe = 2γ Fe−Cu cos (2.38)
2

Austenite grain

Austenite grain
θ
B
C O A

Liquid copper-rich phase

Austenite grain

Figure 2-33: Schematic of a liquid copper-rich phase wetting austenite grain


boundaries (adapted from [96]).

If 2γFe-Cu < γFe-Fe the dihedral angle is 0° and the liquid phase will wet the grain
boundary as a continuous film. Savage et al. [97] note that penetration along grain
boundaries is still possible for dihedral angles greater than 0° but less than 60°. In such
cases, the liquid would penetrate along the grain edge until the predicted equilibrium θ
angle is met. Smith [118] proposed that liquid metal transport along grain boundaries is
a result of a capillarity mechanism. This form of penetration is accelerated by stress and
it is therefore possible that a tensile stress could cause continued penetration for
dihedral angles between 0 and 60° [97].

A number of investigations have sought to elucidate the role of surface tension on the
embrittlement of iron by liquid copper. Van Vlack [119] reported the interfacial surface
free energy at 1105°C of the austenite (γFe-Fe) and liquid copper/austenite (γFe-Cu) grain
boundaries as 850 ergs/cm2 and 430 ergs/cm2, respectively. Substitution of these values

60
Literature Review

in Eqn. (2.38) predicts a dihedral angle of 18°. This suggests that liquid copper would
penetrate rapidly along austenite grain boundaries at 1105°C. Such low dihedral angles
are supported by work carried out by Salter [120]. Model alloys were heat treated at
various temperatures and the dihedral angle existing between the copper-rich and iron-
rich phases were determined (see Figure 2-34). Figure 2-35 illustrates that the dihedral
angle approached zero at a temperature of approximately 1100°C [120]. Interestingly,
the minimum dihedral angle temperature corresponded to the temperature at which the
most liquid copper-rich phase was present (see Figure 2-36). This suggests that the
maximum hot shortness effect would occur at 1100°C due to a combination of the
maximum amount of liquid copper-rich phase available and a maximum in its tendency
to penetrate the austenite grain boundaries.

×750

Figure 2-34: Model alloy heat treated at 1100°C illustrating copper-rich phase
wetting metal grain boundaries (15%Cu-8.7%Ni-mild steel alloy) [120].

30
Dihedral angle (degrees)

20

10

900 1100 1300


Temperature (°C)

Figure 2-35: Dihedral angle-temperature relationship for copper-mild steel system


[120].

61
Literature Review

Figure 2-36: Variation of undissolved copper-rich phase with temperature for


various copper contents [120].

Hough et al. [114] determined the dihedral angle of liquid copper that had penetrated
iron during a notched tensile-creep test. The dihedral angle was found to remain
constant at 34° for samples tested between 1100 and 1130°C. Hough et al. [114] also
performed sessile drop experiments between liquid copper and iron. This type of
wettability study is based on the determination of the contact angle, φ, between a liquid
droplet and solid substrate (see Figure 2-37). A balance of the horizontal forces acting
upon the liquid droplet due to surface tensions yields the following relationship:

γ sv − γ sl
cos φ = (2.39)
γ lv

where γlv, γsv and γsl are the liquid-vapour, solid-vapour and solid-liquid surface
energies, respectively. Hough et al. [114] reported the contact angle between a liquid
copper droplet and iron substrate as 35° and 28° at 1100 and 1130°C, respectively. The
authors commented that contact angle measurements are more sensitive to changes in
temperature than dihedral angle measurements.

62
Literature Review

Liquid droplet

γlv
γs φ γsl

Substrate

Figure 2-37: Schematic of sessile drop configuration.

Savage et al. [97] stated that the use of surface tension equations, such as Eqn. (2.38), to
predict the incidence of surface hot shortness has met with varying levels of success.
However, surface tension is dependent upon composition, temperature, grain
boundaries, surface condition and trace element additions [97]. Obviously, slight
variations in the systems studied by various investigators could account for the
discrepancies.

2.2.2.3.2 Proposed Mechanisms of Cracking

There have been two conflicting mechanisms proposed to describe the process of
cracking resulting from surface hot shortness. The main difference between the
mechanisms is whether an applied stress is required for the material to be embrittled by
the formation of a liquid copper-rich phase. For example, Suzuki [109] commented that
a liquid copper-rich phase penetrates grain boundaries during oxidation, i.e. in the
absence of an applied stress. Subsequently, tensile stresses generated during rolling
cause cracking by LME. Habraken and Lecomte-Beckers [92] reported similar results
by stating that cracks are formed in the reheating furnace and surface break-up occurs
during hot rolling. Under these conditions, defects are introduced into the material
before hot rolling has occurred. In contrast, Zou and Langer [121] and Born [122] stated
that a liquid copper-rich phase penetrates grain boundaries when the material is subject
to tensile stresses. This condition differs to the previous in that an applied stress is
required to embrittle the material.

Determining which proposed mechanism is correct is of considerable practical


importance, as it would influence means of overcoming or minimising surface hot

63
Literature Review

shortness. If the first mechanism is correct, steel is permanently damaged during the
oxidation process. If a liquid copper-rich phase could not be prevented from forming,
then one could minimise the embrittling effect by minimising the degree of penetration,
or by solidifying the copper. If the second mechanism is correct, cracking would be
prevented by introducing a method that avoids copper-rich phase penetration during hot
working.

Results reported from laboratory investigations are not clear on which mechanism is
correct. Kajitani et. al. [107] oxidised mild steel tensile specimens containing 0.20 to
1.5% copper between 1000 and 1350°C in air. The specimens were strained to a
displacement of up to 5mm. Before deformation, some of the precipitated copper-rich
phase was observed to have penetrated grain boundaries. The penetration depth was
relatively small and did not exceed 50 µm. After deformation, copper was observed
either ahead of, or along the crack surface and had clearly penetrated much further than
before deformation. Although not reported, the penetration depth was estimated from a
micrograph at 460µm. Other investigators [102,106,123] have reported observing
penetration of a copper-rich phase from oxidation experiments (i.e. without the
application of stress) but the depth or relative amount of penetration was not quantified.
Chen et al. [124] examined the penetration of liquid copper into low carbon steel by
dipping steel coupons into a liquid copper bath between 1150-1250°C for 10-60
minutes. Penetration along grain boundaries was observed under all conditions and the
depth of copper penetration increased with temperature and time. The average
penetration depth was not reported, but unusually deep penetrations of approximately
200µm were observed.

A further complexity to the interpretation of copper penetration from laboratory


experiments was reported by Seo et al. [116]. These authors found that the rate of
cooling of oxidised samples had an indirect effect on copper-rich phase penetration.
Copper-containing low carbon steels were oxidised at 1100°C for 1 hour in air and
cooled either in air (fast cool) or in a furnace (slow cool). It was reported that the
penetration of grain boundaries rarely occurred for the air-cooled specimens. They
postulated that cooling in air results in scale separation during the early stages of
cooling, whereas in furnace cooling separation is delayed. It was proposed that thermal

64
Literature Review

stresses associated with the difference in thermal shrinkage between the scale and metal
would promote copper penetration and this would be more prevalent in the furnace-
cooled specimens. The observation of copper causing cracking in weld heat-affected
zones [97,98] supports the finding that thermal stresses alone are sufficient to cause
penetration. In these cases, the heat of welding was able to melt small amounts of
copper contamination, which under the action of thermal stresses, penetrated the grain
boundaries by LME. These findings suggest that little or no deformation is required for
cracking to occur. However it is quite possible that the thermal stresses developed in the
heat-affected zone during welding, or in an oxidised sample due to thermal shrinkage
differences, would be higher than those experienced by a billet during oxidation in a
reheat furnace.

These findings suggest that once a liquid copper-rich phase is formed during oxidation,
some penetration will occur. The penetration depth can be influenced by the presence of
stress (whether thermal or mechanical). However, it is unclear whether the degree of
penetration during oxidation contributes significantly to cracking during hot working.

Kajitani et al. [107] investigated crack development from surface hot shortness using a
hot tension test. The authors proposed that there were two stages that described the
growth of cracks: in stage I the cracks develop rapidly in depth; while in stage II the
cracks tend to open considerably in a lateral direction (i.e. the crack depth does not
change appreciably). Experimental results are illustrated in Figure 2-38. It was reported
that liquid copper was present up to the crack tips in stage I, but was not present at the
tips in stage II. The authors concluded that for stage I, there is enough copper present to
readily wet grain boundaries and coat crack surfaces. Cracks do not develop greater in
depth in stage II because the liquid copper is no longer present at the ends of the cracks,
having been consumed in wetting the surfaces of the cracks in stage I. A two-stage
development of crack growth was supported by Seo et al. [116]. Figure 2-39 illustrates
crack depth as a function of their width after 10 and 30% deformation. Deformation
greater than 10% resulted in increased crack width without noticeably increasing the
crack depth.

65
Literature Review

Figure 2-38: Change of crack depth and width as a function of sample


displacement (0.2%C-1.0%Cu steel oxidised at 1100°C for 10 minutes, strain rate
of 5s-1) [107].

Figure 2-39: Crack depths as a function of their width for an interstitial-free,


0.175%Cu-containing steel after oxidation in air at 1100°C for 30 minutes, strain
rate of 2.8×10-1s-1 [116].

If the two-stage crack growth mechanism is correct, the amount of copper precipitated
at the metal surface would determine the depth of cracking. In support of this, Kajitani
et. al. [107] reported that within a temperature range where cracking arose, the depth of
cracking was found to increase with temperature. One would assume that an increase in
temperature would produce an increased copper enrichment. However wettability is

66
Literature Review

temperature dependent and therefore an increase in temperature may have either aided
or hindered the penetration ability. Observation of the dihedral angle vs. temperature
graphs of various copper + x containing mild steel systems (where x is Sn, Ni, Sb, Mn
or As) presented by Salter [120] indicate that in general the relationship appears as in
Figure 2-35. Therefore a change in wettability may either aid or hinder penetration,
depending on whether the temperature range is approaching or advancing from 1100°C.
However, it seems logical to suggest that an increase in the amount of liquid copper-rich
phase would result in increased penetration and thus deeper cracks.

It is proposed that further work be carried out to support or deny the requirement of an
applied stress for copper-rich phase penetration.

2.2.3 Factors Affecting Surface Hot Shortness

The presence of other residual or alloying elements can affect the severity of surface hot
shortness. Salter [120] proposed that the effect these ternary additions have on a copper-
bearing steel can be classified according to whether they are austenite or ferrite
stabilisers. For example, nickel will have a beneficial effect because it is an austenite
stabiliser and therefore increases the solubility of copper in austenite. On the other hand,
tin, antimony and arsenic are ferrite stabilisers and will reduce the solubility of copper
in austenite and therefore be detrimental. The influence of these various elements on the
solubility of copper in austenite at 1200°C and 1250°C is shown in Figure 2-40.

This effect on the solubility of copper in austenite is obviously an important issue.


However, for surface hot shortness to occur, the precipitated copper-rich phase must be
molten. Hence changes in the melting point of the copper-rich phase in the presence of
other elements are also important. However, other elements can have other effects,
including changes to the oxidation process (e.g. promotion of internal oxidation and
occlusion) and changes to the wettability characteristics of the copper-rich phase. A
summary of the different effects of important elements follows.

67
Literature Review

(a) 12 (b) 12
Ni Ni
9 9
Cu (wt%) Cu (wt%)
As
Mn
6 6
As
3 Sb 3
Sn
Sb Sn
0 0
3 6 9 3 6 9 12
Sb, Sn, As, Mn or Ni (wt%) Sb, Sn, As, Mn or Ni (wt%)

Figure 2-40: Influence of various elements on the solubility of copper in austenite,


at (a) 1200°C [70] and (b) 1250°C [125].

2.2.3.1 Steel Chemistry

2.2.3.1.1 Copper

Copper is the only residual element of significant concentration in steel that can form a
molten phase at reheating/hot working temperatures [70,82,87]. The likelihood of a
copper-rich phase precipitating during oxidation is dependent on a number of factors,
including the copper content of the steel, the solubility limit of copper in austenite, the
rate of oxidation and the rate of copper diffusion in the steel. From a steelmaker’s point
of view, the most significant factor is the copper content of the steel. Results from both
laboratory investigations and practical experience have indicated that there is a limiting
copper content below which defects arising from surface hot shortness will not occur.
The limit is commonly cited as 0.2wt%Cu [77,101], although values of 0.15 and
0.3wt% have also been reported [80,91]. As an example, Figure 2-41 illustrates the
rejection rates from surface defects as a function of copper content and illustrates that a
copper content lower than 0.15wt% does not normally cause problems. Of course the
critical copper limit would be influenced by the presence of other elements and the steel
reheating/hot working conditions.

68
Literature Review

Figure 2-41: Influence of copper content on the rejection rates arising from surface
defects for a low carbon steel during wire rod rolling [80].

A qualitative measure, referred to as ‘severity of cracking’, of the amount of cracking


on hot bend test specimen surfaces indicate that the severity increases with increasing
copper levels [88,90,101]. Figure 2-42 illustrates the relationship between copper
content and the cracks developed during hot bend testing. In this case, both the severity
of cracking and maximum crack depth increases with increasing copper content. If
oxidation conditions are assumed constant, steels containing a higher copper content
will precipitate a copper-rich phase more readily (i.e. the solubility limit will be
exceeded earlier) and more embrittling copper-rich phase will be precipitated during
oxidation.

Figure 2-42 also illustrates that at deformation temperatures lower than the melting
point of copper (~1080°C), no cracking will result. This supports the belief that surface
hot shortness requires the formation of a liquid copper-rich phase [69,101,102]. As will
be shown subsequently, the presence of other elements alters the melting point of the
copper-rich phase, and the temperature at which cracking develops.

69
Literature Review

Severity of cracking
3

3.0 Bending temperature:


1150°C
2.5 1100°C
Maximum crack depth (mm)

1050°C
2.0

1.5

1.0

0.5

0 0.2 0.4 0.6 0.8 1.0 1.2


Copper (%)

Figure 2-42: Effect of copper content on the severity of cracking and maximum
crack depth from hot bend tests. The samples were oxidised in air at 1150°C (time
not reported) prior to testing [101].

2.2.3.1.2 Nickel

The addition of nickel to a copper-bearing steel reduces the severity of cracking or the
number of cracks resulting from hot bend or hot tension tests (after oxidation)
[101,102,106,123]. If nickel is present alone with copper, it will have a beneficial effect
by stabilising the austenite [70], and increasing the solubility of copper in austenite.
Nickel has also been reported to raise the melting point of the copper-rich phase [88],
and would thereby contribute to reducing cracking. Akamatsu et al. [103] used Thermo-
Calc [126] to calculate the effect of nickel on the melting point and solubility limit of

70
Literature Review

copper in austenite (see Figure 2-43). This illustrates that both the melting point and
solubility limit increase with nickel addition.

Figure 2-43: Calculated solubility limits of copper in austenite (γ) in the Fe-Cu-Ni
system at various nickel:copper ratios. ∆ refers to peritectic points, while ε and L
refer to the copper-rich and liquid phase, respectively [103].

Because of these effects, effort has been devoted to quantifying the level of nickel
required to inhibit surface hot shortness. Nicholson and Murray [88] performed hot
cracking tests on copper-bearing steels at various nickel contents. They reported that the
incidence of surface cracking was negligible when the nickel:copper ratio was greater
than unity. To further quantify this work, Fisher [106] oxidised steels with
nickel:copper ratios ranging from 0.03 to 1.0. Fisher found that nickel:copper ratios of
1.5-2.0 were required to increase the solubility of copper in austenite sufficiently to
prevent the formation of a molten copper-rich phase. However, the beneficial effect of
nickel additions can be offset by the presence of elements such as tin or antimony,
which decrease the solubility and/or melting point of the copper-rich phase [70]. Salter
[127] investigated the effect of simultaneous additions of nickel and tin on the solubility
of copper in austenite. The effect of tin was found to be so overriding that it was only at
the lowest tin contents that nickel increased the solubility notably.

71
Literature Review

While nickel:copper ratios of 1.5-2.0 may be required to prevent the formation of a


liquid copper-rich phase, Fisher [106] also pointed out that ratios of 1:1 or less can be
effective by promoting internal oxidation and subsurface occlusion for temperatures as
low as 1150°C. The author reported that for nickel:copper ratios as low as 0.1, large
quantities of the copper-rich phase had occluded into the scale. This mechanism has
been shown to suppress cracking in hot tensile tests after oxidation [102]. Fisher [106]
proposed that the enriched layer (of nickel and copper) at the scale/metal interface has a
greater oxygen solubility than the matrix. Hence this allows oxygen to diffuse inward
through the enriched layer until it reaches higher iron concentrations, where it forms
FeO. Perhaps a more correct explanation is that the iron activity in the subsurface region
is significantly reduced by the enrichment of nickel and copper so that the internal
oxidation of iron becomes thermodynamically favourable. The internal oxides coalesce
and impingement with the external scale leads to engulfment of enriched metal surface
regions, resulting in occlusion.

Kohsaka and Ouchi [101] reported that the critical nickel:copper ratio for preventing
surface cracking depends on the reheat temperature. The amount of nickel required
increased when the specimen was reheated at lower temperatures (i.e. 1150°C cf. 1250).
It is possible that this is related to the fact that the occlusion mechanism is more
pronounced at higher temperatures (i.e. where oxidation is more rapid), and therefore
requires less nickel to inhibit cracking at higher temperatures. In addition, Salter [120]
demonstrated that a maximum in the amount of undissolved copper-rich phase occurred
between 1050-1150°C for copper/nickel-containing mild steel alloys (see Figure 2-36),
indicating that higher nickel contents would be required at these temperatures to
suppress cracking.

Recently published work by Fukagawa and Fujikawa [128] is similar to that carried out
in the current project, and their results are briefly reviewed. Low carbon steel with a
copper content of 0.50wt% and nickel additions between 0.001 and 0.25wt% were
oxidised in air at various temperatures (1100-1300°C) for 2 hours. The morphology of
the scale/metal interface and the hot cracking behaviour (using hot tension) was
characterised. Hot tension tests were performed at 1100°C in argon using a strain rate of
0.8×10-2 s-1. For the steel containing negligible amounts of nickel (0.50%Cu-
0.001%Ni), surface cracking was observed for all oxidation temperatures. For the
72
Literature Review

0.50%Cu-0.023%Ni steel, no surface cracking occurred for oxidation temperatures of


1250°C or higher. No surface cracking was observed at any temperature for the
0.50%Cu-0.25%Ni steel.

Cross-sections of oxidised samples revealed a copper-rich phase at the scale/metal


interface at all oxidation temperatures for the 0.50%Cu-0.001%Ni steel. In the case of
the 0.50%Cu-0.023%Ni steel, a copper-rich phase was observed for oxidation
temperatures between 1100-1200°C. Above 1200°C, the scale/metal interface was
found to become irregular and less copper-rich phase was observed. This phenomenon
was more obvious in the 0.50%Cu-0.25%Ni steel. Figure 2-44 illustrates the increased
irregularity of the scale/metal interface with increasing addition of nickel. Fukagawa
and Fujikawa [128] proposed that as the scale/metal interface becomes irregular, the
copper-rich phase becomes occluded into the scale. Surface cracking is inhibited if the
quantity of copper-rich phase present at the metal surface is decreased markedly.

(a) (b)

20µm

Figure 2-44: Subsurface region of (a) 0.50%Cu-0.023%Ni steel and (b) 0.50%Cu-
0.25%Ni steel oxidised at 1200°C in air for 2 hours [128].

This illustrates that the presence of nickel favours the formation of an uneven
scale/metal interface. The interface unevenness increased with oxidation temperature
and nickel content, decreasing the amount of copper-rich phase present at the metal
surface by promoting occlusion. As a result, the hot ductility improved with nickel
addition.

73
Literature Review

Other recent work on the role of nickel in preventing surface hot shortness has been
published by Akamatsu et al. [103]. Steels containing 1wt% copper and up to 1wt%
nickel were oxidised in air at 1200°C for 30 minutes. They observed that the amount of
liquid copper-rich phase was reduced considerably by adding 0.5wt% nickel. No liquid
phase was observed when the nickel content was raised to 1wt%. The scale/metal
interface became progressively more uneven as the nickel content increased as oxide
particles formed at the metal surface.

For the 1%Cu-0.5%Ni and 1%Cu-0.2%Ni steels, Akamatsu et al. [103] reported
copper-rich and copper/nickel-enriched phases existing at the scale/metal interface.
Chemical analysis of these phases for the 1%Cu-0.5%Ni steel is summarised in Table
2-9. These results, along with chemical analysis from the 1%Cu, 1%Cu-0.2%Ni and
1%Cu-0.1%Ni steels, were plotted on a calculated isothermal Fe-Cu-Ni phase diagram
(see Figure 2-45). The compositions of the copper/nickel and copper-rich phases agreed
well with the γ / γ+Liquid and γ+Liquid / Liquid phase boundaries, respectively.
Akamatsu et al. [103] observed that the amount of copper and nickel in the
copper/nickel-enriched phase was nearly equal (i.e. a nickel/copper ratio of 1). Although
there was some enrichment of nickel in the copper-rich phase, most of the enriched
nickel was found in the copper/nickel-enriched phase. For the 1%Cu-1%Ni steel, the
amount of copper and nickel in the copper/nickel-enriched phase gradually decreased as
the distance between the particles and the scale/metal interface increased. However the
nickel/copper ratio remained nearly constant at 1. As a result, they proposed that copper
and nickel enrich in the same ratio at which they existed in the bulk material.

Table 2-9: Reported chemical analysis (wt%) of phases at scale/metal interface for
a 1%Cu-0.5%Ni steel oxidised in air at 1200°C for 30 minutes [103].

Phase Cu Ni Sn Fe
Copper-rich phase 87.0 1.9 2.3 6.1
Copper/nickel- 14.0 11.0 0.1 72.0
enriched phase 13.0 9.2 <0.1 75.0

74
Literature Review

Figure 2-45: Comparison between experimental results and calculated Fe-Cu-Ni


isotherm at 1200°C (C - 1%Cu, N2 - 1%Cu-0.2%Ni, N5 - 1%Cu-0.5%Ni, N10 -
1%Cu-1%Ni) [103].

The gradient of the dashed lines in Figure 2-45, which pass through the initial
composition of the steels, are equal to the nickel/copper ratio of the initial composition
of the steels. As iron is selectively oxidised and if it is assumed that copper and nickel
enrich in the same ratio at which they existed in the bulk material, these lines may be
used to interpret the reported composition of phases in the subsurface region of the
steels [103]. For example, the 1%Cu-1%Ni steel does not form a copper-rich phase
since the solubility limit isn’t exceeded (i.e. the γ / γ+Liquid phase boundary is never
crossed).

Akamatsu et al. [103] calculated a peculiar diagram illustrating a relative fraction


measurement of the amount of liquid copper-rich phase formed at different temperatures
for varying nickel:copper ratios (see Figure 2-46). It is not clear how the relative
fraction was calculated, however the diagram does illustrate that only a small amount of
copper-rich phase forms at high temperatures for an alloy with a nickel:copper ratio of
0.9. The diagram supports the observation that the amount of copper-rich phase formed
in a steel with a nickel:copper ratio of 0.5 is significantly reduced. For example, the
amount of copper-rich phase calculated for temperatures below 1250°C is reduced by
more than 80% (relative to a steel containing no nickel).

75
Literature Review

Figure 2-46: Calculated relative fraction of liquid phase (ratio of the fraction of the
liquid phase of a copper/nickel-bearing steel to that of a steel containing only
copper) formed at various nickel:copper ratios [103].

The addition of nickel to mild steel has been found to slightly decrease the dihedral
angle of the copper-rich phase at austenite grain boundaries, i.e. aiding penetration of
the copper-rich phase along grain boundaries [120]. Nonetheless, Salter [120] proposed
that the addition of nickel would be beneficial in preventing hot shortness due to the
overriding effect of the increased solubility of copper in austenite and the increased
melting temperature of the copper-rich phase.

2.2.3.1.3 Tin

Trace levels of tin (and other elements such as antimony and arsenic) have been shown
to promote surface hot shortness in copper-bearing steels [87]. However, in the absence
of copper, these elements have little effect. Melford [70] showed that Sn, Sb and As
decrease the solubility of copper in austenite. For example, 1% tin, more or less halves
the copper solubility [87]. Tin and antimony are particularly effective at decreasing the
solubility of copper, whereas arsenic reduces the solubility only slightly and is therefore
less detrimental. In addition, Melford [70] stated that the copper-rich phase remains

76
Literature Review

molten to a significantly lower temperature (i.e. about 820°C). No reference was made
under what conditions (e.g. tin content) this observation was made. Melford [70]
proposed that the presence of tin in sufficient quantities would lead to the precipitation
of a molten phase and grain boundary penetration under conditions of much less severe
copper enrichment.

Akamatsu et al. [103] oxidised 1% copper-bearing steels containing up to 0.5% tin in


air at 1200°C for 30 minutes. More liquid phase was observed at the scale/metal
interface for the tin-bearing steels than the copper-only steel. The scale/metal interface
was smooth and frequent penetrations of the liquid phase along grain boundaries were
observed. Such penetrations were observed in copper-only steel, however the presence
of tin appeared to promote penetration. The authors stated that an increase in the
wettability and fluidity of the liquid phase would be expected since tin decreases the
melting point of the copper-rich phase. This statement was not supported by any
reference and is in disagreement with the work of Salter [120], who found that
wettability is partly decreased with tin addition. A calculated Fe-Cu-Sn isotherm (see
Figure 2-47) illustrated that a liquid copper-rich phase may exist for temperatures as
low as 900°C. The diagram also illustrates the drop in solubility of the copper-rich
phase in austenite when only small amounts of tin are present.

Figure 2-47: Calculated boundaries of the γ phase at various temperatures [103].

77
Literature Review

Tin was found to be partly beneficial with respect to the wettability of the copper-rich
phase on austenite. Salter [120] found that the dihedral angle increases with increasing
tin additions to mild steel. However, the author proposed that this would have a limited
effect due to the severe reduction in solubility which tin has on the copper-rich phase in
austenite. Interestingly, Salter [120] found that other elements which cause a reduction
in the solubility of the copper-rich phase (i.e. antimony and arsenic) also cause an
overall increase in the dihedral angle. However, it was proposed that the effect of these
elements on the solubility and melting point of the copper-rich phase dictate to a larger
extent how detrimental their presence will be.

2.2.3.1.4 Silicon

Silicon has been found to be an effective alloying element in suppressing surface hot
shortness cracking [49,73,88,89,107,129]. A contributing factor to this is the
modification of the oxidation mechanism of copper-bearing steels by the formation of
2FeO.SiO2 (fayalite). Fayalite results from the internal oxidation of the silicon in the
steel, and has been observed to form in the scale or subscale regions [50,88,107,129].
This can result in the scale becoming partly liquefied, as fayalite has a melting
temperature of 1205°C and will form a eutectic with FeO (FeO-2FeO.SiO2) at 1177°C
[48,107] and with SiO2 (Fe2SiO4-SiO2) at 1178°C [48]. As described below, the
presence of a solid or liquid fayalite/FeO-fayalite phase will modify the oxidation
mechanism of copper-bearing steels.

The mechanism by which silicon additions decrease surface hot shortness when the
silicon-rich oxide is solid has been attributed to increased levels of internal oxidation,
and hence occlusion of copper-rich phase into the scale (see Figure 2-48a) [89,110,129].
Alternatively, the presence of a solid silicon-rich oxide phase (Fe2SiO4 and/or SiO2) at
the scale/metal interface inhibits the diffusion of iron from the metal to the scale
[48,89]. This decreases the rate of oxidation which would therefore liberate less copper-
rich phase. Lanteri et al. [48] oxidised binary Fe-Si alloys (silicon content up to
1.420wt%) in 1.8%O2-N2 between 1000 – 1300°C. The oxidation rates were found to
decrease with increasing silicon content below 1177°C. Above 1177°C, the presence of
a liquid phase at the scale/metal interface led to higher oxidation rates, with the rate
increasing with silicon content. In addition, Nagumo and Hida [49] oxidised
78
Literature Review

copper/silicon-containing steels and reported similar results to Lanteri et al. [48] with
respect to the change in oxidation rate when the subscale was solid or liquid. However,
a recent study by Seo et al. [129] showed that at both 1100 and 1200°C, the oxidation
rate of 0.1%C-0.5%Mn-0.5%Cu steels increased with a 0.4%Si addition. The authors
reported lower amounts of copper-rich phase at the scale/metal interface owing to
increased levels of occlusion. No explanation was offered by Seo et al. [129] for the
increased oxidation rate at 1100°C.

(a) (b)

20µm 20µm

Figure 2-48: Examples of occlusion associated with fayalite subscale: (a) solid
subscale formed at 1000°C [89] and (b) liquid subscale formed at 1220°C [73].

The mechanism by which surface hot shortness is minimised when a liquid subscale
forms was analysed by Kajitani et al. [107] and attributed to increased levels of
occlusion (see Figure 2-48b). The authors considered the wettability of the scale and the
copper-rich phase, and the molten fayalite/scale and the copper-rich phase. The silicon-
rich liquid phase in the scale was believed to take-up the liquid copper due to a
reduction in the interfacial energy between the scale and the liquid copper as the scale
was liquefied. Unfortunately, there is no reported data of the wettability of copper-rich
phases with SiO2 and iron oxides and so this cannot be confirmed [129]. Generally,
liquid metals have low wettability with oxides, and the wettability varies among
different oxides [130,131]. As Seo et al. [129] point out, this indicates that the
difference between the wettability of the copper-rich phase with FeO and SiO2 would
certainly affect the occlusion behaviour.

79
Literature Review

In general, it appears that the addition of silicon to copper-bearing steels decreases the
amount of copper-rich phase at the scale/metal interface for all reheating/working
temperatures. Kajitani et al. [107] showed that cracking was only suppressed when
fayalite was liquid, which appears to contradict previously mentioned findings
[89,110,129]. Habraken and Lecomte-Beckers [92] reported that the presence of a
viscous or liquid inner oxide layer would negate any beneficial effect of silicon
addition, due to increased oxidation rate and hence copper enrichment. It appears that
there is more experimental evidence to suggest that this is not the case, and that silicon
additions are beneficial at higher temperatures [50,89,107,129]. These apparent
anomalies may be associated with the various testing methods, compositions, heating
conditions etc. employed. However, it must be emphasised that from a scaling kinetics
point of view, the formation of liquid fayalite is detrimental due to increased levels of
metal loss.

2.3 Modelling of Surface Hot Shortness

Hewitt and Meadows [132] reported the development of a computer model based on an
analytical solution of copper and tin diffusion during oxidation. It has been referred to
in some papers as the Bilby model after Dr. B.A. Bilby's contribution to the theoretical
development of the model [82,108]. Unfortunately, the mathematics of the model were
not published. The authors indicated that the model made use of the scaling rates of
steel in a mill furnace, diffusion rates of copper and tin in iron, and the solubility of
copper in the Cu-Sn-Fe system. With respect to the solubility, it was indicated that
when copper alone was present in the steel, a solubility limit of up to 9% was used.
When tin was present, the solubility was reduced to about 4%. Details of the scaling and
diffusion rates were not provided.

The model predicted the thickness of the copper-rich layer, by taking into account the
net effect of the degree of residual enrichment by oxidation and the amount of back-
diffusion into the metal (i.e. diffusion away from the enriched region into the metal) by
these elements. The model was declared a success, based on an excellent correlation
between the number of defects observed microscopically and the predicted layer
thickness (see Figure 2-49). Various simulations were then run to predict the effect of

80
Literature Review

residual content and temperature (see Figure 2-50). From these simulations the critical
temperature, defined as the temperature at which the rate of removal by back-diffusion
exceeds the rate of enrichment, was determined. Using this analysis, various mill
practices were set within a strip rolling mill. Slabs containing no more than 0.15 wt%Cu
were rolled successfully with very few defects when held above 1250°C for 10 to 20
minutes prior to discharge from the reheat furnace. The apparent effect here was to
cause sufficient back-diffusion of the enriched residuals. No observation was made of
increased levels of occlusion at this higher reheat temperature.

Figure 2-49: Relation between defects and predicted layer thickness [132].
Rate of formation (cm/s×108)

80
60 0.25 %Cu - 0.01 %Sn
40 Critical temperature
20 0.05 %Cu - 0.01 %Sn
0
800 1000 1200 1400
-20
Layer redissolving
-40

Temperature (°C)

Figure 2-50: Variation in rate of formation of layer with temperature and copper
content [132]. The units in the original figure were cm × 10-8, however it is believed
they should be cm/s × 108 (as indicated).

81
Literature Review

Hartley et al. [71] reported that the model discussed in [132] had been revised so that it
took into account the combined effects of copper, tin and nickel; using a finite
difference method to solve the diffusion equations. The phenomenon of occlusion had
also been incorporated into the revised model, although the methodology was not
discussed. It appears from the account of Hartley et al. [71] that the revised model was
accurate, allowing prediction of the degree of enrichment under all reheating conditions
and was being used to modify mill practices. A great deal of effort has been made by the
current author to contact the authors of these articles, in an attempt to obtain
information regarding the mathematics of the model. Unfortunately, it appears that the
Bilby model contains proprietary information and is thus unavailable for investigation.

Bergman and West [108] published details of a diffusion model predicting the thickness
of a liquid copper-rich phase during oxidation of an iron-copper alloy. It is believed that
a number of typographical errors and/or incorrect assumptions were made in the paper.
Indeed, substituting their own data into the analytical solution describing the rate of
layer thickening produced a singularity. However, it does appear that the authors were
able to extract sensible results from their model. For example, Figure 2-51 compares the
predicted thickness between the Bergman/West and Bilby models after 1 hour
oxidation. The parameters used for this simulation are summarised in Table 2-10. Under
these conditions, the Bilby model predicts back-diffusion of the copper to exceed
enrichment levels at temperatures greater than approximately 1050°C. In contrast, the
Bergman/West model predicts continuous copper enrichment with temperature. Since it
is generally accepted that either occlusion and/or back-diffusion reduces the amount of
copper-rich phase at the scale/metal interface at higher temperatures [69,71,87,105], it
appears that predictions from the Bilby model are more realistic.

82
Literature Review

Bergman/West model

Thickness of copper layer (m×106)


1

Bilby model

-1
500 700 900 1100 1300 1500
Temperature (K)

Figure 2-51: Calculated thickness of the copper layer after one hour isothermal
heat treatment [108].

Table 2-10: Parameters used in simulations summarised in Figure 2-51 [108]

Parameter Value
 280100 
Diffusivity of copper in iron (cm2/s) 0.434 exp − 
 RT 
Rate constant in terms of metal  169764 
3.04 exp − 
consumption (cm2/s) †  RT 
Copper concentration in copper-rich layer 95%
Copper solubility limit 6%
Original copper concentration in alloy 0.215%
Copper concentration in scale 0.06%

This value was modified so that k cp is defined as l = k cp t , i.e. consistent with
previously reported rate constants in this thesis. Bergman and West [108] expressed
k cp in terms of l = 2k cp t

It seems that the only other model simulating enrichment of residual elements during
oxidation is that of Lewis [94]. In this case, the author attempted to simulate enrichment
from oxidation occurring in continuously cast products. It was assumed that the copper-
rich phase formed as a continuous, uniform layer. Lewis [94] determined that back-
diffusion would not contribute significantly due to the rapid oxidation that takes place
during continuous casting, i.e. metal consumption is rapid compared to residual element

83
Literature Review

diffusion in the metal. As a result it was ignored and the copper-rich layer thickness was
determined solely by the amount rejected due to metal consumption and the amount
passing into the scale (occlusion). Lewis [94] measured the level of copper occlusion in
laboratory castings by dissolving the scale in ammonium chloride solution. The copper
content was then measured by atomic absorption spectroscopy. The results indicated
that approximately 50% of the copper-rich layer precipitated was occluded.
Subsequently, an occlusion factor of 50% was used for all simulations. Agreement
between the predicted and measured thickness was reasonable for small thicknesses.
However the model greatly underestimated the thickness at high measured values (see
Figure 2-52a). A modification was made assuming that complete detachment of the
scale occurred whenever the copper-rich layer thickness reached any multiple of
0.75µm. Lewis [94] felt that the scale remained adhered when the amount of enriched
copper-rich phase was small, but detached when the copper-rich layer thickness reached
a critical value. It can only be assumed that the value of 0.75µm was chosen since very
few thicknesses in excess of this were observed. It was assumed that complete scale
detachment resulted in access of the fresh metal surface to the atmosphere, causing
accelerated oxidation. The result is shown in Figure 2-52b, with good agreement for all
measured results.

84
Literature Review

1.4 Pressure w ater

Predicted thickness ( m )
(a)
1.2 Air-mist
1.0
0.8
0.6
0.4
0.2
0.0
0.0 0.4 0.8 1.2 1.6 2.0 2.4
Measured thickness (µ m )

2.8
(b) Pressure w ater
Predicted thickness ( m )

2.4
Air-mist
2.0

1.6

1.2

0.8
0.4

0.0
0.0 0.4 0.8 1.2 1.6 2.0 2.4
Measured thickness (µ m )

Figure 2-52: Comparison of predicted and measured copper-rich layer thickness


assuming (a) a coherent scale forms and (b) complete scale detachment occurs
when copper-rich layer reaches a critical thickness [94]. Pressure water and Air-
mist refer to different methods of cooling the cast strand.

From the accounts of Hewitt and Meadows [132] and Hartley et al. [71], the Bilby
model appears to have been quite successful in simulating copper enrichment during
reheating for secondary hot rolling. Unfortunately, details on the mathematics of the
model and how it took into account the process of occlusion were never published.
Although the Lewis model [94] accounted for fluctuating oxidation rates caused by
either temperature changes or scale detachment, it appears simplified in the sense that
diffusion was ignored. However, the Lewis model was applied to copper enrichment
during continuous casting. Lewis [94] commented that the oxidation conditions in this
situation are quite different from those existing in the reheat furnace. Nonetheless, it is
the current author’s opinion that a copper enrichment model should take into account
diffusion. Back-diffusion cannot be ignored when the diffusivity in the metal

85
Literature Review

approaches or exceeds the oxidation rate. Lastly, either the mathematics of the
Bergman/West model [108] have been reported incorrectly or are flawed. Since the
details of previously reported models are either flawed, not applicable or not accessible,
the development of a new model is warranted.

2.4 Aims and Focus

This study aims to investigate surface hot shortness in a commercially-produced low


carbon steel grade. Laboratory conditions will be chosen that simulate a processing
environment that leads to surface hot shortness: reheating and hot rolling of copper-
bearing steels for secondary processing. The mechanism of copper enrichment during
oxidation is well agreed upon, however, little has been published on the kinetics of this
process. The current project will define how a copper-rich phase forms and its rate of
formation in a set of model low carbon steels. The effect of oxidation conditions and
alloying additions (nickel and silicon) will also be investigated.

The extent and behaviour of the copper enrichment process during oxidation will be
both modelled and determined experimentally. Details of previous models predicting
copper enrichment are either flawed or not accessible, and so the development of a new
model is warranted. The cracking behaviour of the model steels will be investigated
using a hot compression test. The results will be interpreted in terms of the measured
amount of copper-rich phase formed during oxidation. Oxidation conditions and
alloying additions will be identified that reduce or eliminate cracking arising from
surface hot shortness.

86
CHAPTER
THREE
Experimental Procedure

3.1 Introduction

To fulfil the aims of this investigation, the following areas were defined: measurement
of the oxidation behaviour of a set of copper-bearing model steels; characterisation of
the enrichment behaviour of residual elements during oxidation; modelling of residual
elements accumulation and diffusion during oxidation; and investigation of the cracking
susceptibility of the model steels using a simulated hot working test. The following
sections will outline the procedures used to investigate each research area.

3.2 Oxidation Behaviour of Model Steels

The aim of this area of investigation was to characterise the oxidation behaviour of a set
of model steels. These had carefully controlled compositions, permitting comparison of
steels of different copper content with and without additions of nickel and silicon. The
chemical composition of the steels, determined by atomic emission spectroscopy, is
summarised in Table 3-1. The model steels were designated as copper, copper/nickel
and copper/silicon steels. They were based on a commercially-produced low carbon
grade, A1006, with various additions of copper, nickel and silicon. The amount of tin
was held constant at a value typical of that encountered in low carbon steel produced via
electric arc furnaces. Samples of A1006 from hot-rolled 3mm thick plate and
electrolytic iron were included in the investigation to serve as references.

87
Table 3-1: Composition (wt%) of steels used in investigation

Identification C Si Mn P S Ni Al Cu Sn
0.19Cu 0.042 0.010 0.25 0.010 0.011 0.006 0.019 0.19 0.016
Copper steels
0.47Cu 0.048 0.010 0.25 0.012 0.014 0.006 0.008 0.47 0.017
0.17Cu-0.20Ni 0.060 0.015 0.27 0.009 0.012 0.20 0.040 0.17 0.016
Copper/nickel
0.48Cu-0.13Ni 0.047 0.010 0.27 0.011 0.012 0.13 0.036 0.48 0.017
steels
0.46Cu-0.22Ni 0.048 0.010 0.26 0.009 0.010 0.22 0.046 0.46 0.015
Copper/silicon
0.48Cu-0.52Si 0.065 0.52 0.27 0.012 0.013 0.009 0.035 0.48 0.017
steel
Reference A1006 0.060 < 0.005 0.25 0.060 0.010 0.022 0.034 0.006 <0.002
alloys Iron 0.0012 <0.005 <0.01 <0.001 0.0010 0.002 <0.003 <0.002 <0.002

88
Experimental Procedure

The model steels were produced by BHP Research Melbourne Laboratories. They were
melted in a Heraeus vacuum induction furnace and cast under an argon atmosphere to
ingots with a cross-section of 127×127mm. The ingots were rolled to three thicknesses
(30, 17 and 12 mm) so that material was available for various hot working simulation
tests. Reheating was carried out in a Birlec electric resistance furnace in an 8%CO-
6%H-N2 atmosphere. The ingots were rolled at 1020°C from 130 mm to 30 mm thick
plate. After sectioning, the remaining material was heated for two hours with an exit
temperature of 850°C for rolling from 30 mm to 17 mm. Once sufficient material had
been sectioned, the remaining material was heated for 1 hour with an exit temperature
of 900°C for rolling from 17 to 12 mm.

Samples of the model steels for oxidation experiments were obtained from the 12 mm
rolled plate. They were sectioned from the plate as 12×15×3 mm samples (see Figure
3-1). Identically-sized samples were obtained from the hot-rolled 3mm thick A1006
plate. Smaller, non-uniform samples were obtained from buttons of electrolytic iron
which had been melted several times using a Centorr 5 SA single-arc machine. The
buttons were annealed at 1100°C for 24 hours in high-purity grade argon (99.997%
pure, O2 < 10 ppm, H2O < 12 ppm) prior to sample machining.

3 mm
15 mm

Rolling
direction

Transverse
direction 12 mm

Figure 3-1: Schematic of model steel sample orientation in rolled plate.

89
Experimental Procedure

All samples were polished to a 600 grit finish and ultrasonically degreased in acetone
prior to oxidation. The experiments were performed in a horizontal resistance-tube
furnace using a high-purity recrystallised alumina work tube (inside diameter of 50
mm), illustrated in Figure 3-2. A total of five samples could be oxidised during each
experiment. Electronic mass flow meters were used to control gas flow rates. A total
flow rate of 300 ml/min (i.e. approximately 1.9 cm/s at reaction temperature) was used
for all conditions. The samples were introduced into the cold zone of the furnace
(approximately 200°C) and the work tube sealed. The furnace was purged with flowing
high-purity grade argon and the samples were moved to the furnace hot zone during
purging. Once the experiments were concluded, the work tube was purged with argon
and the samples removed. To avoid damaging the work tube, the samples were
gradually introduced and removed from the reaction zone of the furnace. This process
took approximately 15 minutes. Once samples were removed from the furnace they
were allowed to cool in air. No further oxidation would have occurred, as the samples
would have been no hotter than approximately 200°C upon removal.

90
Experimental Procedure

(a) To fume
cupboard
From purge
gas cylinder
Purge gas tap

Silica pusher rod Bypass tap

From water
Furnace baths or mass
flow meter array

Mass flow
meter array

On-off tap

To furnace or Water baths


Reaction gas cylinders water baths

(b) Resistance-heated furnace Samples


Alumina sample boat

Gas exit Gas entry

Silica pusher
rod

Figure 3-2: (a) Layout of apparatus and (b) detail of furnace used for
discontinuous oxidation experiments.

Initial oxidation experiments were carried out using a simulated reheat furnace
atmosphere. Table 3-2 shows the partial pressure of the equilibrium gaseous products of
natural gas combustion calculated by a free energy minimisation program called
Chemix [133] at 1100°C for a total pressure of 1 atm at 112% stoichiometric
combustion (i.e. an excess of 12% air over stoichiometric combustion of natural gas
with air). Details of this calculation have been summarised in Appendix A.1.

91
Experimental Procedure

Table 3-2: Calculated partial pressures of the equilibrium products of natural gas
combustion for 112% stoichiometric combustion at 1100°C and 1 atm

Species Partial Pressure (atm)


CO2 0.086
CO 3.63×10-7
N2 0.725
O2 0.020
H2O 0.169
H2 3.48×10-7

Unfortunately, oxidation in the simulated reheat furnace atmosphere was non-uniform


(discussed in Section 4.2.1). In view of these problems, the reaction gas was changed to
clean, dry air. Discontinuous oxidation experiments were carried out at 1050, 1150 and
1250°C for times ranging from 10 minutes to 3 hours. In order to slow the oxidation rate
and minimise scale separation during reaction, an 8.6%CO2-N2 gas mixture was
substituted for air at 1250°C. The composition was chosen so that the gas would have a
CO2 quantity equivalent to that of the simulated reheat atmosphere (see Table 3-2).

Oxidised samples were subjected to microstructural analysis carried out on cross-


sections which were mounted and polished to a 1µm finish. An Olympus PMG-3 optical
microscope with a digital camera attachment and the Leica Q500MC image analysis
package (ver. 01.00) was used for general observations and measurement of the scale
thickness. The average scale thickness was determined from 10 random measurements
along a sample side to which the scale was adhered. These values, measured after
different oxidation times, were used to determine the scaling rate constant for the steels.

A Cahn D-200 thermogravimetric analysis (TGA) balance (see Figure 3-3) was used to
determine scaling rate constants at 1250°C. A JEOL 840 scanning electron microscope
(SEM) was used to acquire secondary electron (SE) or back-scattered electron (BSE)
images of the scale and subsurface regions of the oxidised samples.

92
Experimental Procedure

TGA balance

Gas exit Gas entry


Computer

Furnace Silica reaction


controller tube

Furnace

Sample

Figure 3-3: Schematic of thermo-gravimetric analysis balance used for obtaining


oxidation kinetics.

3.3 Enrichment Behaviour of Residual Elements during


Oxidation

The aim of this area of investigation was to characterise the enrichment of residual
elements during oxidation. The amount of copper-rich phase formed during oxidation
was measured and compared against predicted values.

The amount of copper-rich phase formed at the scale/metal interface was measured by
image analysis using BSE images. Although the copper-rich phase can be observed
under an optical microscope, the contrast between it and the steel was poor. The ability
to distinguish between these phases was far more successful using BSE images. Ten
randomly-selected images at a magnification of 1000× were obtained for each sample.
The amount of copper-rich phase at the metal surface, ACu, was expressed as an average
surface area per unit scale/metal interface length (see Figure 3-4):

Surface area of copper - rich phase


ACu = (3.1)
Scale - metal interface length

93
Experimental Procedure

Scale/metal interface length

Scale

Cu-rich phase Cu-rich phase

Metal

Figure 3-4: Schematic illustrating method of measuring amount of copper-rich


phase formed at the scale/metal interface.

The enrichment of residual elements in the subsurface region of the model steels was
determined by measuring the composition profiles perpendicular to the scale/metal
interface. A Cameca SX50 electron probe micro-analyser was used for quantitative
composition analysis using wave-length dispersive spectroscopy (WDS). All samples
and electron probe micro-analysis (EPMA) standards were carbon-coated prior to
analysis as some of the phases analysed were non-conductive (e.g. the oxide scale). The
EPMA standards used were pure Fe, Cu, Ni, Sn, Si, Mn, Al, Fe2O3 (for oxygen),
Ca5(PO4)3(OH) (Apatite- for phosphorus) and FeS2 (Marcasite - for sulfur). An
accelerating voltage and probe current of 15 keV and 20 nA, respectively, were used for
all analysis. The intensity of the characteristic x-ray peak of each element was corrected
by subtracting the background x-ray level and absorption and fluorescence was
accounted for on an atomic number and sample composition basis via the usual matrix
ZAF correction procedure [134]. Beam scans with a step size of approximately 0.5µm
were used to measure composition profiles across the scale/metal interface.

94
Experimental Procedure

3.4 Cracking Behaviour of Model Steels

The aim of this area of investigation was to examine the cracking behaviour of the
model steels using a simulated hot working test. Previous hot workability investigations
of surface hot shortness have used hot tension [102,107,109,123,128] and hot bend
[88,101,122,132] tests. The stress state developed in tension and bend tests differs from
that formed in rolling operations, since the primary stress component of rolling is
compression. As a result, hot compression testing may provide a more useful
representation of industrial processes [135]. Cracks developed due to surface hot
shortness during hot rolling are transverse [136], and must form due to the action of
secondary tensile forces. A compression test developing secondary tensile forces about
the mid-span bulge zone more closely simulates conditions of crack development during
hot rolling (see Figure 3-5).

Mid-span bulge zone

Cracks
Transverse Primary
direction compressive
force

Longitudinal Secondary
direction tensile force

Figure 3-5: Schematic illustrating cylindrical test specimen having undergone a


compression test. Secondary tensile forces developed at the mid-span bulge zone
may induce cracking.

Cylindrical specimens of the model steels (10mm diameter × 110mm) were hot
compressed after oxidation, at a high strain rate using a Gleeble 3500. The Gleeble 3500
is a thermo-mechanical testing unit capable of high speed heating (via direct resistance
heating) and compression or tension (via a hydraulic servo system). Figure 3-6
illustrates the testing chamber of the unit during oxidation of the specimen. The

95
Experimental Procedure

temperature was measured by a thermocouple (type K) spot-welded to the centre of the


specimen. Cooling was aided by directing compressed air at the specimen using the
quench nozzles.

Quench
Thermocouple nozzles

Copper
grips

Figure 3-6: Sample chamber of Gleeble during oxidation stage of hot compression
test.

Only a small amount of material was available for specimen machining and the possible
number of tests was limited. As a result, a selection of the model steels and A1006 were
chosen for testing. Specimens of A1006 were sourced from 42mm thick plate. All
specimens were machined to a tolerance of ± 0.025mm and cleaned in ethanol prior to
testing. The testing conditions are summarised in Figure 3-7. Oxidation temperatures
ranged between 1000 and 1250°C, while deformation temperatures ranged between 900
and 1100°C. The majority of tests were performed using the oxidation conditions
outlined in Table 3-3 and a compression temperature of 1050°C. These times were
chosen so that a similar thickness of scale formed at each temperature. The standard
compression temperature of 1050°C was chosen as it was consistent with two different
processing conditions of commercially produced A1006. The secondary hot rolling of
A1006 is carried out between 870 and 1100°C [137]; while in thin slab casting, hot

96
Experimental Procedure

rolling generally occurs between 950 and 1050°C [138]. Flowing air in the specimen
chamber was achieved by directing compressed air from the quench nozzles at the
specimen. The effect of strain and strain rate could not be investigated due to limited
specimen numbers. Hence, the compression was set at 10mm achieved in 0.05s for all
tests.

Oxidation Compression
(in flowing air) (in still air)

Cross-head speed of
50°C/s 10mm in 0.05s
50°C/s

25°C/s

Figure 3-7: Oxidation/hot deformation conditions.

Table 3-3: Standard oxidation conditions for hot compression testing

Oxidation temperature (°C) Oxidation time (minutes)


1050 30
1150 10
1250 5

Deformed samples were cross-sectioned along the transverse plane of the mid-span
bulge zone (see Figure 3-5) and polished to a 1µm finish. Crack characteristics (amount
of cracking, average and maximum crack depth) were measured on the cross-sectioned
samples using an optical microscope and Leica Q500MC image analysis software (ver.
01.00). To account for variations, all measurements were conducted over half of each
samples circumference (see Figure 3-8a). The amount of cracking was determined by
measuring the surface area of the cracks at a magnification of 200×. The amount of
cracking, ACrack, was expressed in terms of the total interface length sampled (see Figure
3-8b):

97
Experimental Procedure

Total area of cracking


ACrack = (3.2)
Total interface length sampled

(a)

Interface
length Cracks
analysed
Cross-sectioned
specimen

(b) Interface length

Cracks Cracks

Metal

Figure 3-8: Schematics illustrating the method used to (a) sample and measure
crack characteristics and (b) measure the surface area of cracking per unit sample
interface length.

98
CHAPTER
FOUR
Oxidation Behaviour of Model Steels

4.1 Introduction

This chapter is concerned with characterising the oxidation behaviour of a set of model
steels. Residual element enrichment at the steel surface is dependent on, among other
things, the metal consumption rate. As a result, determining the oxidation rate of the
model steels is an important factor in describing residual element enrichment.

Three groups of model steels were produced: copper, copper/nickel and copper/silicon
steels. Two reference materials, A1006 and electrolytic iron, were included in the
investigation. Discontinuous oxidation experiments were conducted in clean, dry air at
1050 and 1150°C, and in 8.6%CO2-N2 at 1250°C for times ranging from 10 minutes to 3
hours. Oxidised samples were characterised using optical and electron microscopy. The
rate of oxidation was determined by measuring the scale thickness at 1050 and 1150°C,
while at a 1250°C, TGA was used. Details of alloy production, sample preparation,
oxidation experiment conditions and analysis techniques are discussed in Section 3.2.

4.2 Results

4.2.1 Initial Oxidation Experiments

Initial oxidation experiments were conducted in a simulated reheat furnace atmosphere.


The oxidation behaviour of all samples was found to be non-uniform. Figure 4-1
illustrates that the thickness of scale varied across the sample face. This indicates that
there was insufficient oxygen available for uniform scale growth. As a result, the total
gas flow rate was increased from 300 to 500 mL/min. Non-uniform oxidation behaviour
was still present using the higher flow rate. An analysis of this behaviour is discussed in
detail in Section 4.3.1. Although it was believed that further increases in the total gas

99
Oxidation Behaviour of Model Steels

flow rate would solve the problem, limitations in the experimental setup prevented
implementation. In view of these problems, the reaction gas was changed to air.

(a)

(b)

100µm

Figure 4-1: Appearance of scale layer across sample face after oxidation in 112%
Air atmosphere at 1100°C for 2 hours: (a) Leading edge and (b) trailing edge of
sample.

Although uniform oxidation was observed in air, scale separation and cracking was
found to occur readily. Figure 4-2 illustrates the difference in appearance between a
scale layer that remained adhered and one that had separated during oxidation. Most
noticeable is that the adhered layer is much thicker than the separated layer. This
indicates that scale separation occurred during oxidation. Evident in Figure 4-2 is that

100
Oxidation Behaviour of Model Steels

separated scales often contained a higher proportion of oxygen-rich iron oxides (e.g.
Fe2O3).

FeO (+ decomposed FeO) Fe3O4 Fe2O3

(a)

Fe3O4 Fe2O3

(b)

100µm

100µm

Figure 4-2: (a) Adhered and (b) separated scale on 0.47Cu after oxidation in air at
1050°C for 1 hour.

As will be described in Section 4.3.2, the level of residual element enrichment could not
reliably be related to the amount of consumed metal when scale separation occurred. In
addition, the thickness of scale formed could not be related to the oxidation time. As a

101
Oxidation Behaviour of Model Steels

result, analysis of scale thickness and residual element enrichment was confined to
adhered scale layers. In many instances scale separation occurred on both sample sides
which led to experiments having to be repeated. In view of the large number of repeats
occurring at 1050 and 1150°C, a reaction gas of lower oxidation potential was used at
1250°C. No scale separations were observed using 8.6%CO2-N2 at 1250°C. Table 4-1
compares the thickness of scale formed on iron at 1050, 1150 and 1250°C. This
illustrates that very thick scales were formed when oxidation took place in air. Rapid
scale growth obviously contributed greatly in promoting scale separation.

Table 4-1: Thickness of scale formed on iron after 30 minutes oxidation

Condition Scale thickness (µm)


1050°C in air 511 ± 5
1150°C in air 674 ± 17
1250°C in 8.6%CO2-N2 119 ± 7

4.2.2 Oxidation Behaviour

The type of scale formed depended on the steel composition and the conditions of
oxidation. When oxidation was carried out in air, longer oxidation times favoured the
formation of an inner FeO, middle Fe3O4 and thin outer Fe2O3-layered scale (see Figure
4-3a). Shorter oxidation times favoured the formation of FeO-only scale (see Figure
4-3b). In some instances the FeO layer decomposed to FeO + Fe3O4 during sample
cooling. In many instances, a whisker/blade scale morphology was observed at the
scale/gas interface (see Figure 4-3c). A row of pores within the scale adjacent the
scale/metal interface was observed in the copper, copper/nickel, A1006 and iron scales
formed in air (see Figure 4-3a and b).

102
Oxidation Behaviour of Model Steels

(a)

(b)

(c)

Figure 4-3: Oxidation in air: (a) FeO (+ decomposed FeO)/Fe3O4/Fe2O3 scale


formed on 0.17Cu-0.20Ni oxidised at 1050°C for 3 hours, (b) FeO-only scale
formed on 0.19Cu oxidised at 1150°C for 30 minutes, (c) whisker scale growth on
0.48Cu-0.13Ni oxidised at 1150°C for 10 minutes.

103
Oxidation Behaviour of Model Steels

When oxidation was carried out in the CO2-N2 mixture, FeO-only scales were formed at
all oxidation times analysed. The outer scale surfaces of these alloys were wavy,
differing from the relatively flat outer surface of the air-oxidised samples (see Figure
4-4).

Figure 4-4: Oxidation in CO2-N2: wavy appearance of scale/gas interface formed


on 0.48Cu-0.13Ni oxidised for 90 minutes.

The 0.48Cu-0.52Si steel formed a two-phase subscale between the metal and iron oxide
scale under all oxidation conditions. This phase was identified by EPMA as containing
fayalite (2FeO.SiO2) and FeO. At 1050°C the subscale formed as a solid (see Figure
4-5) whereas at 1150 and 1250°C it was liquid (see Figure 4-6). The change in state is
attributed to the reported melting point of the fayalite-FeO eutectic as 1177°C [48]. A
relatively planar scale/metal interface formed when the subscale was solid (see Figure
4-5a) compared to a wavy scale/metal interface when it formed as a liquid (Figure
4-6b). Liquid fayalite was often observed to have penetrated the scale (see Figure 4-6c).

104
Oxidation Behaviour of Model Steels

(a)

(b)

Figure 4-5: Appearance of subscale when solid: (a) fayalite-FeO subscale/FeO (+


decomposed FeO)/Fe3O4 scale formed on 0.48Cu-0.52Si oxidised at 1050°C for 3
hours, (b) higher magnification image of subscale/metal interface.

105
Oxidation Behaviour of Model Steels

(a)

(b)

(c)

Figure 4-6: Appearance of subscale when liquid: (a) fayalite-FeO subscale/FeO (+


decomposed FeO)/Fe3O4 scale formed on 0.48Cu-0.52Si oxidised at 1150°C for 45
minutes, (b) higher magnification image of subscale/metal interface, (c) fayalite
penetration into scale.

106
Oxidation Behaviour of Model Steels

4.2.3 Rate of Oxidation

The kinetics of oxide scale growth can be described using the following equation:

X n = k't (4.1)

where X is the scale thickness, k' the rate constant, t the time and n = 1 for linear, 2 for
parabolic etc. All steels oxidised at 1050°C according to parabolic kinetics, as
illustrated in Figure 4-7. Parabolic oxidation signifies that diffusion through the scale is
rate-determining [5].

0.014
0.012
0.01 Iron
0.47Cu
X 2 (cm2)

0.008
0.17Cu-0.20Ni
0.006
A1006
0.004 0.48Cu-0.52Si
0.002
0
0 0.5 1 1.5 2 2.5 3
Oxidation time (hours)

Figure 4-7: Oxidation kinetics of a selection of steels at 1050°C.

At 1150°C, the kinetics of scale formation became less clear, as scale growth for all
steels was non-uniform. Results for a selection of steels in Figure 4-8 illustrate irregular
variation of scale thickness with time. Multiple experiments revealed that separation
during reaction was occurring readily. For scales that were taken to be adhered, it is
possible that separation and healing occurred during reaction. Such a mechanism would
explain the observed non-uniform growth.

107
Oxidation Behaviour of Model Steels

1400
1200
1000
Iron
X (µ m)

800
0.46Cu-0.22Ni
600
0.48Cu-0.52Si
400
200
0
0 10 20 30 40 50 60
Oxidation time (minutes)

Figure 4-8: Oxidation kinetics of a selection of steels at 1150°C.

Scales formed on steels at 1250°C grew according to kinetics which were initially linear
and then parabolic, as illustrated in Figure 4-9. The time of transition between linear
and parabolic kinetics was estimated for most steels to be approximately 60 minutes.
Scale growth controlled by diffusion in the gas phase or by a surface reaction gives rise
to linear growth. As scales thicken, diffusion through the scale becomes rate controlling
and further growth obeys a parabolic rate law.

Rate constants at 1050 and 1250°C were determined by linear regression on Eqn (4.1)
and have been summarised in Table 4-2. At 1050°C, the 0.19Cu and 0.47Cu steels
oxidised at a similar rate to iron. Additions of nickel slightly lowered the oxidation rate,
while silicon had a marked effect in decreasing the rate of oxidation. The rate constant
for the oxidation of iron in air at 1050°C has been reported earlier as 1.2×10-6 cm2/s
[65]. This rate constant has been expressed in terms of scale thickness by a conversion
assuming a multilayered iron oxide scale formed (see Appendix A.2). This compares
favourably to that obtained in the present work and serves to validate the experiment.

108
Oxidation Behaviour of Model Steels

0.06

Normalised weight (g/cm 2)


0.05

0.04 Iron
0.47Cu
0.03
0.46Cu-0.22Ni
0.02 0.48Cu-0.52Si

0.01

0
0 20 40 60 80 100 120
Time (minutes)
(a)

0.005
(Normalised weight)2 (g2/cm 4)

0.004
Iron
0.003
0.47Cu
0.46Cu-0.22Ni
0.002
0.48Cu-0.52Si
0.001

0.000
40 60 80 100 120
Time (minutes)
(b)

Figure 4-9: Oxidation kinetics of a selection of steels at 1250°C: (a) weight gain vs.
time and (b) (weight gain)2 vs. time.

109
Oxidation Behaviour of Model Steels

Table 4-2: Rate constants for steels, where the subscripts l and p refer to the linear
and parabolic rate constants, respectively (the rate constants at 1250°C have been
expressed in terms of scale thickness by a conversion that assumes only an FeO
scale forms, see Appendix A.2)

1050°C 1250°C
Steel
kp' (cm2/s) kl' (cm/s) kp' (cm2/s)
0.19Cu 1.4×10-6 7.1×10-6 4.0×10-7
0.47Cu 1.3×10-6 6.5×10-6 3.3×10-7
0.17Cu-0.20Ni 9.9×10-7 6.7×10-6 1.9×10-7
0.48Cu-0.13Ni 1.0×10-6 6.1×10-6 2.3×10-7
0.46Cu-0.22Ni 6.6×10-7 6.0×10-6 2.4×10-7
0.48Cu-0.52Si 2.9×10-7 4.2×10-6 2.0×10-7
A1006 6.2×10-7 6.8×10-6 2.2×10-7
Iron 1.3×10-6 8.0×10-6 5.5×10-7

At 1250°C, the linear rate constants for all alloys were similar apart from the 0.48Cu-
0.52Si steel which was significantly lower. The parabolic rate constants of the
copper/nickel and copper/silicon steels were slightly lower than the remaining steels at
1250°C. No reported values for the oxidation of iron or steel in similar atmospheres at
1250°C could be found. Rate constants using similar atmospheres at lower temperatures
have been reported (see Table 2-1 and Table 2-3). Rate constants at 1250°C have been
estimated from published data (see Table 4-3) assuming an Arrhenius relationship for
scale growth. There is a large variation in the estimated linear rate constant. One would
expect a higher kl' for pure CO2 compared to CO2-CO or CO2-N2 atmospheres. This is
not reflected in the estimated result from [60]. Regardless, the linear rate constants of
the current work fall within the range of previously estimated values.

The estimated parabolic rate constants from published data are all similar, which is
expected when the rate-determining step is diffusion through the oxide. However, the
parabolic rate constants obtained from the current work are an order of magnitude lower
than those estimated from published data. This result was not investigated further, due
to time limitations.

110
Oxidation Behaviour of Model Steels

Table 4-3: Rate constants estimated at 1250°C from published data

Conditions Rate constant


Iron in 50%CO2-CO (922-1037°C) [19] 5.8×10-8
Iron in CO2 (920-1100°C) [21] kl' (cm/s) 1.7×10-6
Low carbon steel in 10%CO2-N2 (1000-1150°C) [60] 7.1×10-5
Low carbon steel in 6%O2-N2 (1000-1250°C) [61] 2.3×10-6
Low carbon steel in 1%O2-10%CO2-3%H20-N2 (900-
kp' (cm2/s) 2.6×10-6
1150°C) [62]
A1006 in 112% Air (700-1200°C) [67] 1.3×10-6

4.2.4 Internal Oxidation

In addition to external scale formation, all model steels and A1006 internally oxidised
under all oxidation conditions. Micrographs of the subsurface region of various steels
oxidised in air and in the CO2-N2 atmosphere are presented in Figure 4-10 and Figure
4-11, respectively. The internal oxides of the copper steels formed as a zone of fine
internal oxides (generally less than 1-2µm in diameter) with larger internal oxides (up to
10 µm in diameter) nearer the scale/metal interface. Large internal oxides appeared
more frequently in the copper/nickel steels and often formed an interlinking network
adjacent to the scale/metal interface (see Figure 4-10d). The internal oxides of the
copper/silicon steel were mostly in the form of small spherical particles, although some
larger oxides were observed.

Figure 4-10a, b and c; and Figure 4-11a, b, c and d reveal that some of the larger
internal oxides are not single phase. The poor resolution of the EMPA electron image
made differentiating these components very difficult and thus chemical analysis of the
individual components was not obtained. Nonetheless, chemical analysis of whole
internal oxide particles were performed. Due to the small size of the fine internal oxides,
analysis was confined to the large internal oxides. The reported chemical analysis was
based on the average of a minimum of 5 analysis points.

111
Oxidation Behaviour of Model Steels

(a)

(b)

(c)

Figure continued overleaf

112
Oxidation Behaviour of Model Steels

(d)

(e)

Figure 4-10: Appearance of internal oxides formed in steels oxidised in air: (a)
0.47Cu oxidised for 10 minutes at 1150°C, (b) 0.17Cu-0.20Ni oxidised for 30
minutes at 1050°C, (c) 0.46Cu-0.22Ni oxidised for 10 minutes at 1150°C, (d)
0.46Cu-0.22Ni oxidised for 15 minutes at 1150°C, (e) 0.48Cu-0.52Si oxidised for 45
minutes at 1150°C.

113
Oxidation Behaviour of Model Steels

(a)

(b)

(c)

Figure continued overleaf

114
Oxidation Behaviour of Model Steels

(d)

Figure 4-11: Appearance of internal oxides formed in steels oxidised in CO2-N2: (a)
0.47Cu oxidised for 90 minutes at 1250°C, (b) 0.17Cu-0.20Ni oxidised for 90
minutes at 1250°C, (c) 0.46Cu-0.22Ni oxidised for 90 minutes at 1250°C, (d)
0.48Cu-0.52Si oxidised for 120 minutes at 1250°C.

Table 4-4 summarises analysis of large internal oxide particles formed in copper/nickel
steels exposed to air at 1150°C. These results indicate that internal oxides found deeper
within the metal were complex and enriched in P, Mn, Al and Si. Internal oxides
adjacent to the scale/metal surface were FeO. Similar results were found for the copper
steels. Interestingly, no noticeable quantities of phosphorus were detected in the large
internal oxides formed in the copper and copper/nickel steels exposed to CO2-N2 at
1250°C.

Table 4-4: Chemical analysis of large internal oxides in copper/nickel steels formed
at 1150°C

Composition (at%)
Steel Region
P Mn Si Al Fe O
0.17Cu- Adjacent interface ~ 0.15 0.17 0.26 47.00 52.33
0.20Ni 25 µm from interface 5.25 3.15 8.95 1.06 22.44 58.46
0.46Cu- Adjacent interface ~ 0.15 0.12 0.22 46.44 52.78
0.22Ni 40 µm from interface 6.64 2.34 6.15 0.65 25.52 58.46

Table 4-5 summarises chemical analysis of large internal oxides formed in the
copper/silicon steel. The composition of these are noticeably different from those
formed in the other steels, being mainly composed of Fe, Si and O. Lanteri et al. [48]

115
Oxidation Behaviour of Model Steels

oxidised iron-silicon alloys in 1.8%O2-N2 at 1200°C and reported that SiO2 particles
formed at the internal oxidation front. Internal oxides nearer the metal surface were
reported to have converted to fayalite. It is likely that the internal oxides of the
copper/silicon steel were similar to those reported by Lanteri et al. [48]. The chemical
analysis summarised in Table 4-5 suggests the internal oxides were mixed fayalite/SiO2
oxides.

Table 4-5: Chemical analysis of large internal oxides formed in the 0.48Cu-0.52Si
steel

Temperature (°C) P Al Mn Si Fe O
1050 ~ 0.06 0.27 8.45 45.44 45.49
1150 0.25 0.99 3.33 13.20 24.52 56.78
1250 0.21 0.12 0.24 11.79 30.16 57.41

Not all particles observed within the metal were identified as internal oxides. Figure
4-12 illustrates an irregular-shaped, light-coloured particle adjacent to the scale/metal
interface. EPMA analysis of this particle is summarised in Table 4-6, indicating that it
was copper/iron sulfide. It should be noted that the occurrence of these particles was
rare.

Figure 4-12: Irregular-shaped particle (indicated by arrow) adjacent scale/metal


interface observed in 0.48Cu-0.13Ni oxidised for 60 minutes in air at 1050°C.

116
Oxidation Behaviour of Model Steels

Table 4-6: Composition of particle in Figure 4-12

Element Concentration (at%)


Cu 43.91
Fe 15.37
S 40.72

4.3 Discussion

4.3.1 Oxidation in Simulated Reheat Atmosphere

As mentioned in Section 4.2.1, the non-uniform oxidation behaviour was most likely a
result of there being insufficient oxygen available to sustain scale growth. The flux of
oxygen to the sample surface can be estimated from kinetic equations. If it is assumed
that the gas flow within the hot zone of the furnace is in the viscous regime, the
maximum flux, Ji, of oxidant i to the sample surface can be approximated by [139]:

k m (i ) p i
Ji = (4.2)
RT

where km(i) is the average boundary layer mass transfer coefficient for i, pi is the partial
pressure of i, R is the universal gas constant and T the temperature. Details of this
calculation are summarised in Appendix A.3. The maximum flux of oxygen in the
112% Air combustion atmosphere at 1100°C at atmospheric pressure (1 atm) was
calculated as 7.19×10-2 g/m2s.

ox
The expected flux of oxidant, J i , required to sustain oxide scale growth can be
expressed as:

ox ∂∆W
Ji = (4.3)
∂t

where ∆W is the normalised weight change and t the time of oxidation. If the parabolic
rate constant, k pp , is defined as:

117
Oxidation Behaviour of Model Steels

∆W 2 = k pp t (4.4)

ox
then J i becomes:

ox k pp
Ji = (4.5)
2∆W

In the present work, rate constants could not be determined reliably. In a previous study
[67] in which A1006 was oxidised in 112% Air at 1100°C, k pp was determined as 47

mg2cm-4min-1. Normalised weight change data from this study and the corresponding
ox ox
values for J i are summarised in Table 4-7. For the bulk of the oxidation period, J O 2

> J O 2 . Hence, the flux of oxygen required to sustain the expected growth was

insufficient. This calculation proves that the observed non-uniform oxidation behaviour
was a result of there being insufficient oxygen available for oxide growth.

Table 4-7: Calculated expected oxygen flux using data for the oxidation of A1006
in 112% Air at 1100°C [67]

ox
Time (min) ∆W (mg/cm2) J O2 (g/m2s)
10 24 1.63×10-1
20 34 1.15×10-1
30 40 9.79×10-2
40 47 8.33×10-2
50 53 7.39×10-2
60 57 6.87×10-2

4.3.2 Scale Separation

Scale separation during oxidation is a result of a stress-relieving mechanism. During


oxidation, the oxide layer thickens while the oxide/metal interface recedes. Stresses are
developed in the scale layer as it attempts to maintain adherence to the metal substrate
[4]. The compressive stresses generated in the scale can be relieved by one or more of

118
Oxidation Behaviour of Model Steels

the following: plastic deformation of the scale or of the metal, detachment (spalling) at
the metal/scale interface and elastic failure (cracking) of the scale [24]. When oxidation
was carried out in air, both cracking and/or detachment of an oxide layer were observed
for most samples. Stress build-up is dependent on the sample size and the rate at which
oxidation occurs. The high proportion of scale separation when samples were oxidised
in air reflects the relatively high rate of oxidation in this atmosphere at the temperatures
of interest.

When a scale layer separates from the metal, the flux of iron through the oxide is halted,
and therefore growth ceases, resulting in a thinner scale layer. This is clearly illustrated
in Figure 4-2. The higher proportion of oxygen-rich iron oxides (e.g. Fe2O3 and Fe3O4)
observed in separated scales is due to conversion of FeO to these oxides. A separated
scale equilibriates with the gas and iron and oxygen diffuse through the oxide layers in
an attempt to eliminate their chemical potential gradient. The overall effect is that the
relative proportion of Fe3O4 and Fe2O3 increases at the expense of FeO. This effect is
also clearly observed in Figure 4-2.

There are two important implications of scale separation during oxidation. Firstly, the
degree of residual element enrichment is dependent on the amount of metal consumed
during oxidation. A smaller oxide thickness would produce a lower degree of residual
element enrichment. In addition, if oxide growth (and thus metal consumption) had
ceased during oxidation, so too would residual element enrichment. This would allow
time for the enriched elements to diffuse back into the metal, decreasing the level of
enrichment at the metal surface. Secondly, scaling rate constants calculated from scale
thickness measurements from a separated side would be lower than those determined
from adhered sides. Because scale separations occurred on one side of almost all
samples oxidised in air, weight gain measurements were meaningless. As a result,
observations and analysis were confined to regions of the sample where scale separation
during reaction had not occurred.

4.3.3 Scale Morphology

In section 4.2.2, it was stated that longer oxidation times in air favoured the formation
of a multilayered scale while shorter oxidation times favoured the formation of an FeO-
119
Oxidation Behaviour of Model Steels

only scale. Table 4-8 summarises the equilibrium dissociation pressure for formation of
iron oxides between 1050 and 1250°C. It is obvious that all three iron oxides should
have formed when oxidation took place in air. The fact that distinct layers of Fe3O4 and
Fe2O3 were not observed for short oxidation periods indicates that FeO consumed these
oxides when it’s growth rate was high, i.e. during the earlier stages of oxidation when it
grew rapidly.

Table 4-8: Equilibrium dissociation pressures for formation of iron oxides

Equilibrium dissociation pressure (atm)


Oxide
1050°C 1150°C 1250°C
FeO 1.1×10-14 3.0×10-13 5.3×10-12
Fe3O4 2.6×10-12 1.4×10-10 4.5×10-9
Fe2O3 3.7×10-6 8.9×10-5 1.4×10-3

When oxidation was carried out in the CO2-N2 atmosphere, an FeO-only scale was
formed on all alloys (neglecting the fayalite subscale formed in the copper/silicon steel).
The equilibrium p O 2 in 8.6%CO2-N2 at 1250°C was calculated using Chemix [133] to

be 4.4×10-5 atm. FeO and Fe3O4 are stable under these conditions if the equilibrium
oxygen partial pressure in contact with the sample is maintained (see Table 4-8). The
oxygen partial pressure at the sample surface is likely to be far lower than that present
away from the surface. Oxygen atoms passing over the sample surface are expected to
be consumed quickly, resulting in a surface partial pressure lower than that required for
formation of Fe3O4.

The whisker/blade scale morphology observed in some samples oxidised in air is most
likely a result of oxygen starvation. Oxide whiskers may develop if gas phase diffusion
of oxidant to the surface controls the growth rate [1]. It is likely that the whiskers
developed as the furnace was purged with argon and the samples were in the process of
being removed and cooled. Sample removal from the furnace was a gradual process,
taking approximately 15 minutes. During this time, the oxygen content of the reaction
atmosphere would have dropped significantly. This would have provided conditions
favourable for whisker growth.

120
Oxidation Behaviour of Model Steels

Pore formation within the scale layer adjacent to the scale/metal interface (see Figure
4-3a and b) was observed by Sheasby et al. [30] for the oxidation of iron in 4%O2-N2 at
1200°C. They analysed polished cross-sections of samples oxidised between 30 seconds
and 1 hour (see Figure 4-13). Pores developed during the early stages of oxidation when
idiomorphic crystals formed. Small pores were trapped at crystal junctions when the
crystals began to grow. Larger pores were trapped between crystals when the upper
surface of the crystals were bridged by oxide. Since all pores observed in this
investigation were located near the scale/metal interface, it is proposed that their growth
occurred in a manner similar to that reported by Sheasby et al. [30]. The fact that a
network of pores was not observed in the copper/silicon steel must be related to the
modification of oxide growth when a subscale is present. This point was not
investigated further as it was unrelated to the project aims.

121
Oxidation Behaviour of Model Steels

(a) (b)

(c)

(d)

(e)

Figure 4-13: Iron oxidised in 4O2-N2 at 1200°C for (a) 30 s, (b) 3.75 min (c) 6 min,
(d) 10 min and (e) 1 hour [30].

The waviness of the scale surface when the steels were oxidised in CO2-N2 (see Figure
4-4) is related to the fact that for most of its growth, the scale layer grew according to a
linear rate law. This implies that the rate-controlling mechanism is either the transport
of oxidant from the atmosphere, a surface reaction or the incorporation of oxygen
anions into the scale lattice. As outlined previously (see Section 4.3.1), a mass transfer
calculation can be used to determine whether transport of oxidant was rate controlling.

122
Oxidation Behaviour of Model Steels

The equilibrium partial pressure of components in an 8.6%CO2-N2 gas mixture was


calculated using Chemix [133] and is summarised in Table 4-9. Using these values, J O 2

and J CO 2 were calculated to be 1.64×10-4 and 2.84×10-1 g/m2s, respectively.

Table 4-9: Equilibrium partial pressure of components in 8.6%CO2-N2 at 1250°C


(1 atm total pressure)

Species Partial Pressure (atm)


CO2 8.59×10-2
CO 8.81×10-5
N2 9.14×10-1
O2 4.40×10-5

For linear oxidation kinetics:

∆W = k pl t (4.6)

where k pl is the linear rate constant. Hence, Eqn. (4.5) must be modified:

ox
Ji = k pl (4.7)

The linear rate constant for iron exposed to the CO2-N2 atmosphere at 1250°C was
measured at k pl = 1.0×10-5 g/cm2s. Note that the rate constants summarised in Table 4-2

at 1250°C were converted from the measured rates expressed in terms of weight gain.
ox ox
As a result, J O 2 = 1.0×10-1 g/m2s. Since J O 2 > J O 2 , it is clear that molecular

oxygen was not the oxidant. However the flux of CO2 was high enough to support the
linear rate. This was not observed for most of the oxidation period. As a result, this
analysis indicates that rate control by gas diffusion did not occur. It can be concluded
that processes at the scale-gas interface controlled the rate of oxidation. The observed
surface irregularities may have developed since it is likely that surface reaction control
is sensitive to oxide grain orientation. For example, it is expected that adsorption of

123
Oxidation Behaviour of Model Steels

oxygen atoms or anions would occur most readily at preferred planes of oxide crystal
[12]. This would promote the formation of an irregular scale/gas interface.

4.3.4 Rate of Oxidation

4.3.4.1 Copper Steels

At 1050°C, the copper steels oxidised at a similar rate to iron. As will be discussed in
Section 5.2.1, a semi-continuous layer or discrete particles of a copper-rich phase
formed at the scale/metal interface. The point of interest is whether this enrichment of
residual elements would be expected to alter the oxidation rate. When compared to the
oxidation of pure iron, localised enrichment of noble elements at the scale/metal
interface would be expected to lower the iron activity there. For example, the copper-
rich phase in the 0.47Cu steel at 1050°C was measured to be 92.18at%Cu-6.90%Fe-
0.77%Sn-0.15%Ni. The effect of a lower iron activity at the scale/metal interface on the
oxidation rate is investigated using Wagner’s [35] expression for the diffusion
controlled process:


( ) − (p ) 
1 1
o i n
k p ∝  pO2 n
O2 (4.8)
 

o i
where p O 2 is the partial pressure of oxygen at the oxide/gas interface, p O 2 the partial

pressure of oxygen at the metal/oxide interface and n is equal to 6 for doubly charged
o i
cation vacancies. Generally p O 2 >> p O 2 and thus the rate of FeO formation becomes

( )
1
i o n
independent of p O 2 , i.e. k p ∝ p O 2 . Under these conditions, variation in the activity

of iron (and consequently the oxygen activity) at the scale/metal interface will not affect
i
the oxidation rate. However, if p O 2 cannot be assumed negligible, then one would

expect the oxidation rate reduced compared to pure iron. This has been reported by
Nanni and Gesmundo [53], who oxidised an Fe-4.54wt%Cu alloy in oxygen between
700 and 1000°C. They found that the oxidation rate was significantly decreased with
respect to pure iron. This was attributed mainly to the increase of the oxygen activity at

124
Oxidation Behaviour of Model Steels

the alloy/scale interface due to the presence of a copper-rich phase layer. In the current
study, however, the copper-rich phase was not present as a continuous layer and
therefore the boundary conditions at the scale/metal interface are complex. The metal in
contact with the scale differed, being either a copper-rich phase or iron enriched in
residual elements. In either case, one would expect there to be an increase in the oxygen
activity at the scale/metal interface compared to the pure iron situation. The fact that the
copper steels and iron oxidised at a similar rate indicate that none of the factors
discussed contributed significantly in altering the oxidation rate.

The oxidation rates measured in the CO2-N2 atmosphere at 1250°C did not differ
significantly from that obtained for pure iron. Under these conditions, oxidation initially
obeyed a linear rate law. As there was no variation in the reaction atmosphere between
samples and the scale formed on all samples was identical, one would expect similar
linear rate constants.

4.3.4.2 Copper/Nickel Steels

The parabolic rate constants of the copper/nickel steels determined at 1050 and 1250°C
steels were slightly lower than that for pure iron. As with the copper steels, localised
enrichment of noble elements at the scale/metal interface would be expected to lower
the iron activity there, which may result in a decrease in the oxidation rate. Figure 4-14
illustrates concentration profiles across the scale/metal interface of the 0.48Cu-0.13Ni
steel oxidised at 1250°C. No copper-rich phase was present in this particular analysis,
however localised enrichment of both copper and nickel are clearly observed.

125
Oxidation Behaviour of Model Steels

Scale Metal

5
Concentration (at%)

Cu
3
Ni

0
0 5 10 15 20 25
Distance (µ m)

Figure 4-14: Concentration profiles across scale/metal interface of 0.48Cu-0.13Ni


steel oxidised for 90 minutes at 1250°C in 8.6%CO2-N2.

4.3.4.3 Copper/Silicon Steel

The parabolic and linear rate constants of the copper/silicon steel at 1050 and 1250°C,
respectively, were lower than that obtained for pure iron. Microstructural observations
at 1050°C revealed that the fayalite-FeO subscale formed as a solid (see Figure 4-5).
The solid subscale most likely offers a barrier to iron diffusion and thus slows the rate
of scale formation [48]. At 1150 and 1250°C, the subscale was observed to have formed
as a liquid (Figure 4-6). The presence of a liquid fayalite is likely to increase the
oxidation rate due to increased levels of diffusion in the liquid phase [129]. However,
microstructural observations at 1150 and 1250°C revealed large gaps between the metal
surface, subscale and scale layer. If the presence of a liquid subscale promotes scale
separation, the overall oxidation rate may decrease.

Interestingly, the linear rate constant at 1250°C of the copper/silicon steel was lower
than that obtained for the other steels. It is expected that if scale growth is controlled by

126
Oxidation Behaviour of Model Steels

diffusion in the gas phase or by a surface reaction, that a similar rate constant be
determined for steels forming the same type of scale. Since FeO was formed on all
steels and the change in the rate constant could not be accounted for by experimental
error, some other mechanism must be responsible. This point was not investigated
further as it was unrelated to the project aims.

4.3.5 Internal Oxidation

It is apparent from Table 4-4 that some internal oxides contained surprisingly high
concentrations of P, Mn, Al and/or Si when considering their concentration in the bulk
material (see Table 3-1). The likelihood that diffusion from the bulk supplied sufficient
quantities of these elements to form the internal oxides may be investigated by a
simplified diffusion calculation. Figure 4-15 schematically illustrates the concentration
profile of a solute across a spherical internal oxide and adjacent metal (i.e. along section
A - A').

Internal oxide

A A’

Cp
Concentration

Co

Xr Xd Distance

Figure 4-15: Schematic of solute concentration profile existing across a spherical


internal oxide and adjacent metal. The solute concentration in the internal oxide
particle is Cp while in the bulk metal is Co.

127
Oxidation Behaviour of Model Steels

For a particle of radius Xr and concentration Cp of solute, the distance Xd required for
solute atoms to diffuse to form the particle can be approximated by:

(C p − Co )
Xd = Xr (4.9)
Co

where Co is the solute concentration in the bulk metal, and the diffusion profile in the
matrix has been ignored. Xd can be compared with the characteristic diffusion distance
Xc for a solute atom:

X c = 2 Dt (4.10)

where D is the diffusivity of the solute in the bulk metal and t is the time available for
diffusion. For this analysis, t was assumed to be the time the sample was oxidised. This
is most certainly an overestimation, as the time available for growth of the internal
oxides would be much shorter. The results are summarised in Table 4-10, illustrating
that the solute elements could not have diffused from the bulk in sufficient quantity to
form the oxides. It therefore is concluded that these elements were enriched/segregated
in these regions prior to oxidation.

Table 4-10: Comparison between Xd and Xc for solute atoms found in an internal
oxide (Xr = 2µm) in 0.17Cu-0.20Ni oxidised at 1150°C for 15 minutes (diffusivity
data taken from [140,141])

Solute Xd (cm) Xc (cm)


-2
P 6.5×10 1.7×10-3
Mn 2.1×10-3 3.8×10-4
Al 2.4×10-3 1.1×10-4
Si 5.6×10-2 4.8×10-3

In the case of the copper and copper/nickel steels, all large internal oxides adjacent to
the scale/metal interface were identified as FeO. This may be interpreted by assuming
that near the scale/metal interface, FeO was more stable than a mixed impurity oxide.
As the scale/metal interface receded and the oxides were approached by the interface,

128
Oxidation Behaviour of Model Steels

the concentrations of the impurity elements within the internal oxide were diluted as
they were converted to FeO. The situation for the copper/silicon steel differs somewhat,
since oxygen diffusion within the metal results in the internal oxidation of silicon
forming SiO2. This may react with FeO to form fayalite. It is for this reason that the
large internal oxides analysed in the copper/silicon steel were identified as mixed
fayalite/SiO2 oxides.

A zone of fine internal oxides was observed to have formed in all steels. Chemical
analysis was impossible due to their small size (less than 1-2 µm). In the case of the
copper and copper/nickel steels it seems logical that these were FeO. For example, a
network of interlinking particles of FeO adjacent to the scale/metal interface in the
copper/nickel steels (see Figure 4-10d) were commonly observed. It is highly unlikely
that the majority of these large internal oxides were the result of internal oxides
nucleated on the impurity particles discussed above. It is more likely that they are the
result of iron being internally oxidised. This is possible due to the enrichment of noble
elements at the metal surface, so that the iron activity is reduced and the internal
oxidation of iron becomes thermodynamically favourable. The following analysis will
focus on the simplified situation of the oxidation of a binary iron-nickel alloy in view of
the marked effect that nickel has on promoting internal oxidation.

Consider a single phase, iron-nickel alloy. In situations in which both iron and nickel
are reactive with oxygen at the temperature and gas pressure involved (i.e. high-oxidant
pressure), the alloy composition determines which oxide forms. Since an alloy with a
low nickel concentration is being considered (i.e. residual level), an external layer of
FeO is expected to form while nickel diffuses back into the alloy. Providing nickel does
not enrich at the metal surface sufficiently for NiO to become stable, FeO will continue
to form. This situation is represented schematically in Figure 4-16.

129
Oxidation Behaviour of Model Steels

FeO Fe-Ni
NFeo

NFei

Concentration NNii
NOi
NNio

Distance

Figure 4-16: Schematic concentration profiles (Fe, Ni and O) for exclusive growth
of an external layer of FeO in a single phase Fe-Ni alloy.

In internal oxidation analysis, it is customary to use the solubility product, Ksp, to


determine under what conditions internal oxidation becomes favourable. The solubility
product is defined as the inverse of the equilibrium constant [37], K. For iron to oxidise,
Ksp for the following reaction must be exceeded:

Fe Fe− Ni + O Fe− Ni = FeO (s) (4.11)

where i Fe− Ni represents species i dissolved in the Fe-Ni alloy and the solubility product

is defined as:

K sp = N O Fe − Ni ⋅ N FeFe − Ni (4.12)

where N i Fe- Ni is the mole fraction of i in the Fe-Ni alloy. The situation of iron oxidising

both internally and externally is represented schematically in Figure 4-17. It is obvious


that Ksp must be exceeded at the scale/metal for external FeO formation. However,
simultaneous internal oxidation of iron requires the concentration product NO.NFe to be
equal to or greater than Ksp within the alloy (see Figure 4-17b). This is possible since,
although the oxygen concentration decreases with distance from the scale/metal
interface, the concentration of iron increases.

130
Oxidation Behaviour of Model Steels

FeO Fe-Ni
(a) NFeo

NFei

Concentration
Internal
oxidation
zone

NNii
NOi
NNio

(b)
Ksp
NO.NFe

Distance

Figure 4-17: Schematic of (a) concentration profiles for simultaneous internal and
external oxidation of iron and (b) concentration product as a function of distance.

The Gibbs energy change for Eqn. (4.11) can be determined from its constituent
equations:

1
Fe (s) + O 2(g) = FeO (s) ∆GI (4.13)
2

1
O Ni = O 2(g) ∆GII (4.14)
2

Fe Fe − Ni = Fe (s) ∆GIII (4.15)

The corresponding Gibbs energy equations (in J/mole) are as follows:

∆GI = −264889 + 65.4T [142] (4.16)

131
Oxidation Behaviour of Model Steels

∆GII = −181800 + 70.2T (adapted from [143]) (4.17)

∆GIII = 4310 − RT ln N Ni + 1.2T [142] (4.18)

The Gibbs energy for Eqn. (4.11), ∆GIV, is thus the sum ∆GI + ∆GII + ∆GIII :

∆GIV = −442379 − RT ln N Ni + 136.8T (4.19)

The expression relating ∆GIV to the solubility product, defined in Eqn. (4.12), is:

 ∆GIV 
K sp = exp  (4.20)
 RT 

Hence, conditions are favourable for the internal oxidation of iron when
N O Fe − Ni ⋅ N Fe Fe − Ni ≥ K sp . The variation of Ksp with temperature for an original nickel

concentration of 0.002 is summarised in Table 4-11. A typical example of the


concentration profile formed in a copper/nickel steel at 1150°C is shown in Figure 4-18.
However, the concentration values (especially for oxygen) are spurious within the metal
between the scale and the internal oxide regions. This is a result of adjacent phases
being sampled simultaneously. A value for NFe at the metal/internal oxide interface was
estimated by assuming that the NFe:NNi:NCu ratio at the interface was correct and by
scaling the concentrations to 1. Taking NFe = 0.97, the minimum oxygen concentration
required for internal oxidation, NOmin, was determined to be 4.9×10-8. The oxygen
solubility limit in the alloy, NO(s), was estimated by assuming that practically all iron is
consumed in the subsurface region and that the alloy in this region can be considered to
be pure nickel. The oxygen solubility limit in pure nickel at 1150°C was taken as
9.8×10-5 [143]. Since NO(s) > NOmin, internal oxidation is expected to occur.

132
Oxidation Behaviour of Model Steels

Table 4-11: Variation in Ksp with temperature (NNi° = 0.002)

Temperature (°C) Ksp


1050 2.4×10-8
1150 4.0×10-7
1250 4.7×10-6

Scale Metal Internal oxide Metal

100
90
80
Concentration (at%)

70
Fe
60
O
50
Cu
40
Ni
30
20
10
0
0 2 4 6 8 10 12
Distance (µ m)

Figure 4-18: 0.17Cu-0.20Ni steel oxidised for 15 minutes at 1150°C.

4.4 Summary and Conclusions

Initial oxidation experiments were conducted in a simulated reheat furnace atmosphere.


Non-uniform scale growth was observed for all samples. The flux of oxygen to the
sample surface was estimated using kinetic equations. This calculation confirmed that
there was insufficient oxygen available for oxide growth. When the reaction gas was
changed to air, uniform scale growth was observed. However, scale separation and
cracking was found to occur readily. The rapid scale growth in air caused stresses to
develop that could be relieved by cracking or scale detachment. The level of residual
element enrichment could not reliably be related to the amount of consumed metal when
scale separation occurred. Consequently, all measurements of scale thickness and

133
Oxidation Behaviour of Model Steels

residual element enrichment were confined to regions of the sample where scale
separation during reaction had not occurred. In many instances, scale separation
occurred on both sample sides which led to experiments having to be repeated. In view
of the large number of repeats occurring at 1050 and 1150°C, oxidation was carried out
in 8.6%CO2-N2 at 1250°C. No scale separations were observed at 1250°C due to the
much lower rate of scale growth.

Oxidation at 1050 and 1150°C in air formed multi-layered scales of FeO, Fe3O4 and
Fe2O3. Shorter oxidation times favoured the formation of FeO-only scales, indicating
that FeO consumed the higher order oxides when its growth rate was high. FeO-only
scales were formed in 8.6%CO2-N2 at 1250°C due to the much lower effective oxygen
potential of the CO2-N2 atmosphere. In addition to external scale formation, all model
steels and A1006 internally oxidised under all oxidation conditions. Analysis of internal
oxides of the copper and copper/nickel steels revealed that internal oxides found deeper
within the metal were complex and enriched in P, Mn, Al and Si. Internal oxides
adjacent to the scale/metal surface were FeO. It is likely that the model steels contained
some impurity particles rich in P, Mn, Al and/or Si that were converted to internal iron-
rich oxides during oxidation. The mixed impurity internal oxides were converted to FeO
as they approached the scale/metal surface, since FeO is more stable. The copper/silicon
steel formed internal oxides of SiO2. These reacted with FeO to form fayalite at the
scale/metal interface resulting in an FeO-fayalite subscale.

All steels oxidised at 1050°C according to parabolic kinetics. The kinetics at 1150°C
could not be determined due to scale separation and healing during oxidation. Scales
formed at 1250°C obeyed an initially linear law followed by a parabolic rate growth
law. The copper steels were found to oxidise at a similar rate to iron. The copper/nickel
and copper/silicon steels oxidised more slowly than iron. The decrease was more
prominent in the copper/silicon steel. At 1050°C a solid fayalite/FeO subscale offered a
barrier to iron diffusion and thus lowered the oxidation rate. At 1250°C, the presence of
a liquid subscale promoted the formation of gaps between the metal and scale, resulting
in a decrease in the oxidation rate.

134
CHAPTER
FIVE
Enrichment Behaviour of Residual Elements during Oxidation

5.1 Introduction

This chapter is concerned with characterising the enrichment of residual elements


during oxidation of a set of model steels. An understanding of where copper and other
residual elements enrich, how quickly they enrich and the importance of factors that
affect their enrichment is obviously paramount in designing strategies to overcome or
minimise the damage caused by surface hot shortness.

Analysis was performed on polished cross-sectioned samples from discontinuous


oxidation experiments. Residual element enrichment was investigated using optical and
electron microscopy. The effect of steel composition and oxidation conditions (e.g.
atmosphere, temperature and time) on copper-rich phase formation were investigated.
Quantitative measurement of residual element enrichment was performed using EPMA.
In addition, the amount of copper-rich phase formed was measured by performing
image analysis on BSE images. The measured values were compared with predicted
levels. Details of alloy production, sample preparation, oxidation experiment conditions
and analysis techniques are discussed in Section 3.2.

5.2 Results

5.2.1 Copper-rich Phase Morphology

This section describes the way copper-rich phase was distributed and its morphology in
the oxidised model steels. The composition of the phase and the effect of alloy
composition and oxidation conditions (atmosphere, temperature and time) on its
formation will be reported in Sections 5.2.2 and 5.2.3.

135
Enrichment Behaviour of Residual Elements during Oxidation

When a copper-rich phase was observed to have formed, most was found distributed on
the scale/metal interface. It was not distributed uniformly across the interface and
existed as discrete particles and/or as a semi-continuous layer. Figure 5-1 illustrates the
variation in distribution on different areas of the same sample. The copper-rich phase
appears as the brightest since these are BSE images, in which a higher intensity
indicates a phase of higher atomic weight.

(a)

(b)

Figure 5-1: Variation in the copper-rich phase morphology at the scale/metal


interface; (a) semi-continuous layer and (b) discrete particles formed in 0.48Cu-
0.13Ni oxidised for 10 minutes at 1150°C.

Smaller amounts of copper-rich phase were often observed to have penetrated along
metal grain boundaries. Figure 5-2 illustrates examples of a copper-rich phase being
present at both the scale/metal interface and within the metal.

136
Enrichment Behaviour of Residual Elements during Oxidation

(a)

(b)

Figure 5-2: Penetration of copper-rich phase into the metal; (a) 0.48Cu-0.52Si
oxidised for 3 hours at 1050°C and (b) 0.47Cu oxidised for 10 minutes at 1150°C.

Copper-enriched particles were also observed to have been occluded in the scale.
Occasional particles were observed for most model steels at all oxidation temperatures,
with some steels exhibiting a larger degree of occlusion. These characteristics will be
discussed in more detail in Section 5.2.2. Figure 5-3 illustrates that occluded particles
were often distributed together and were of a very small size (usually less than 1µm in
diameter). In many cases when occluded particles were not observed on flatly-polished
regions of scale, they were identified along regions where portions of the scale had
become detached or damaged during sectioning/polishing. This is illustrated in Figure
5-4, where the bright particles are copper-enriched and the dark regions are holes
remaining from occluded particles that were removed as the scale was damaged. It is
believed that the surfaces along which occluded particles were observed were oxide

137
Enrichment Behaviour of Residual Elements during Oxidation

grain boundaries surfaces. It therefore appears that the occluded particles were
preferentially located at oxide grain boundaries.

Scale/gas interface

Figure 5-3: Occluded particles (indicated by arrows) in scale layer on 0.17Cu-


0.20Ni oxidised for 30 minutes at 1050°C.

138
Enrichment Behaviour of Residual Elements during Oxidation

Damaged region
Scale/metal interface
of scale

(a)

(b)

Figure 5-4: Occluded particles along surface of damaged/detached portion of scale;


(a) SE image of scale width and (b) BSE of same region illustrating copper-
enriched particles (0.47Cu oxidised for 60 minutes at 1250°C).

When oxidation was carried out in air (at 1050 and 1150°C), copper-enriched particles
were often observed at or near the outer scale surface. These particles were not observed
at this location for the model steels oxidised at 1250°C. Figure 5-5 illustrates particles at
and near the base of the scale whiskers, which were often observed for steels oxidised in
air at 1050 and 1150°C.

139
Enrichment Behaviour of Residual Elements during Oxidation

Figure 5-5: Copper-enriched particles (indicated by arrows) at scale edge of


0.47Cu oxidised for 10 minutes at 1150°C.

5.2.2 Copper-rich Phase Formation

5.2.2.1 General Remarks

Copper-rich phase formation was observed for all steels except iron and A1006. Its
presence at the scale/metal interface depended on both alloy composition and oxidation
conditions (atmosphere, temperature and time). Table 5-1 summarises under what
conditions a copper-rich phase was observed at the scale/metal interface. Oxidation
conditions at the temperature investigated are summarised in Table 5-2. A copper-rich
phase was observed at the interface for all model steels at 1050°C. At 1150°C, this
phase was present for all but one of the model steels (i.e. 0.17Cu-0.20Ni). In general,
the oxidation experiments were carried out at much shorter times at this temperature to
avoid scale detachment. The shorter experiments highlighted the fact that a copper-rich
phase formed rapidly (i.e. less than 10 minutes) at this temperature for most model
steels. At 1250°C the formation of the copper-rich phase was either delayed for an
extended period of time or it did not form at all.

140
Enrichment Behaviour of Residual Elements during Oxidation

Table 5-1: Conditions under which a copper-rich phase was observed at the
scale/metal interface (see Key for details)

Oxidation Temperature (°C)


Steel
1050 1150 1250
0.19Cu X X \
0.47Cu X X Y
0.17Cu-0.20Ni X \ \
0.48Cu-0.13Ni X X Z
0.46Cu-0.22Ni X X Z
0.48Cu-0.52Si X X [

Key X Copper-rich phase observed for all oxidation times


Y Copper-rich phase commenced formation between 30 and 60 minutes
of oxidation
Z Copper-rich phase commenced formation between 60 and 90 minutes
of oxidation
[ Copper-rich phase commenced formation between 90 and 120 minutes
of oxidation
\ No copper-rich phase observed for any oxidation time

Table 5-2: Oxidation conditions for discontinuous experiments

Temperature (°C) Oxidant Oxidation time


1050 Air 30, 60, 90 and 120 minutes
Range of times between 10 – 120 minutes,
1150 Air mostly less than 60 minutes due to scale
separation problems
1250 8.6%CO2-N2 30, 60, 90 and 120 minutes

The following sections describe copper-rich phase formation and residual element
enrichment with respect to the various classes of model steels.

141
Enrichment Behaviour of Residual Elements during Oxidation

5.2.2.2 Copper Steels

Figure 5-6 illustrates the appearance of the copper-rich phase at the scale/metal interface
in the copper steels. In general, it appeared in the form of discrete particles or as a semi-
continuous layer. Some copper-rich phase was observed to have penetrated metal grain
boundaries at all temperatures when a copper-rich phase was present at the interface.
Occluded particles were observed within the scale at all temperatures, with the
exception of the 0.19Cu steel at 1250°C. The occluded particles were rare at 1050°C,
but more numerous at 1150 and 1250°C.

Analysis of the copper-rich phase for all model steels was complicated by the often
small quantities formed and their small thickness. The spatial resolution of the EPMA is
in the order of 1-2µm, and in many cases the thickness of the copper-rich phase was less
than this. However, larger deposits were occasionally observed and compositional
analysis was confined to these regions. A minimum of five beam points were performed
and averaged and the results are summarised in Table 5-3. Significant partitioning of tin
to the copper-rich phase was observed at all temperatures. In addition, the concentration
of copper decreased, while that of iron increased with increasing temperature.

142
Enrichment Behaviour of Residual Elements during Oxidation

(a)

(b)

(c)

Figure 5-6: Appearance of copper-rich phase at scale/metal interface in copper


steels; 0.47Cu oxidised for (a) 30 minutes at 1050°C, (b) 10 minutes at 1150°C and
(c) 90 minutes at 1250°C.

143
Enrichment Behaviour of Residual Elements during Oxidation

Table 5-3: Copper-rich phase composition (at%) of 0.47Cu steel

Temperature (°C) Cu Fe Sn Ni
1050 92.2 6.9 0.8 0.1
1150 91.3 6.3 2.4 ~
1250 89.2 7.6 3.2 ~

Typical concentration profiles of various elements about the scale/metal interface of the
0.47Cu steel are represented in Figure 5-7. In considering residual element enrichment
after oxidation it must be remembered that in measuring steep concentration gradients
by EPMA the x-ray source generated is about 1-2µm diameter. This means that the
EPMA will generate signals from adjacent phases in the region of a phase boundary,
leading to the diffuse interfaces observed in Figure 5-7. This artefact is more obvious in
regions of sharp discontinuities in the concentration gradient, e.g. at the scale/metal or
copper-rich phase/metal interface).

(a) Scale Cu-rich phase Metal

100
90
80
Concentration (at%)

70
Fe
60 O
50 Cu
40 Sn
Ni
30
20
10
0
0 2 4 6 8 10 12
Distance (µ m)

Figure continued overleaf

144
Enrichment Behaviour of Residual Elements during Oxidation

(b) Scale Cu-rich phase Metal Internal oxide Metal

100
90
80
Concentration (at%)

70 Fe
O
60
Cu
50
Sn
40
Mn
30 Si
20
10
0
0 1 2 3 4 5 6 7 8 9 10 11 12
Distance (µ m)

(c) Scale Cu-rich phase Metal

100
90
80
Concentration (at%)

70
60 Fe
O
50
Cu
40
Sn
30
20
10
0
0 2 4 6 8 10 12 14 16
Distance (µ m)

Figure 5-7: Concentration profiles about scale/metal interface of 0.47Cu oxidised


for (a) 30 minutes at 1050°C, (b) 15 minutes at 1150°C and (c) 90 minutes at
1250°C.

145
Enrichment Behaviour of Residual Elements during Oxidation

A copper-rich phase was present at the scale/metal interface in each case in Figure 5-6.
An internal oxide enriched in manganese and silicon was also analysed in Figure 5-7b.
Concentration profiles at 1050 and 1150°C revealed that the enrichment depth of copper
within the metal was in the order of 1-2µm. However, at 1250°C the enrichment depth
was significantly higher (5-20µm).

5.2.2.3 Copper/Nickel Steels

Figure 5-8 illustrates the appearance of the copper-rich phase at the scale/metal interface
in the copper/nickel steels. As with the copper steels, the copper-rich phase was in the
form of discrete particles or as a semi-continuous layer. In addition, some copper-rich
phase was observed to have penetrated metal grain boundaries at all temperatures when
a copper-rich phase was present at the interface. The lighter-coloured particles in the
inner scale layer of Figure 5-8b and c are occluded particles. In contrast to the relatively
smooth scale/metal interface of the copper steels, the interface was noticeably rougher
in the copper/nickel steels.

146
Enrichment Behaviour of Residual Elements during Oxidation

(a)

(b)

(c)

Figure 5-8: Appearance of copper-rich phase at scale/metal interface in the


copper/nickel steels; 0.46Cu-0.22Ni oxidised for (a) 60 minutes at 1050°C, (b) 15
minutes at 1150°C and (c) 120 minutes at 1250°C.

147
Enrichment Behaviour of Residual Elements during Oxidation

Typical examples of the copper-rich phase composition of the copper/nickel steels are
summarised in Table 5-4. No analysis was performed at 1250°C due to the error
involved with analysing the small quantities formed at this temperature. In contrast to
the copper steels (see Table 5-3), a significant amount of nickel partitioned to the
copper-rich phase. When viewed under the optical microscope, the colour of the copper-
rich phase in the copper/nickel steels was duller than the bright yellow observed in the
copper and copper/silicon steel.

Table 5-4: Copper-rich phase composition (at%) of 0.46Cu-0.22Ni steel

Temperature (°C) Cu Fe Sn Ni
1050 82.5 10.8 1.7 5.0
1150 76.6 10.8 2.9 9.7

Although a copper-rich phase was observed at the scale/metal interface for all
copper/nickel steels at 1050°C, it was not observed for the 0.17Cu-0.20Ni steel at
1150°C (see Table 5-1). Figure 5-9a illustrates the appearance of the scale/metal
interface for this alloy at 1150°C. Occluded, copper-enriched particles are clearly
observed within the scale, even though no copper-rich phase was observed at the
interface. Figure 5-9b shows the complete scale width of this sample with occluded
particles appearing throughout the scale. Similar particles were observed within the
scale for all copper/nickel steels at all temperatures. In addition, they were more readily
observed in the copper/nickel steels than in the copper steels.

148
Enrichment Behaviour of Residual Elements during Oxidation

(a)

(b)

Figure 5-9: Occluded particles in 0.17Cu-0.20Ni steel oxidised for 15 minutes at


1150°C: (a) adjacent scale/metal interface and (b) across scale width.

Typical concentration profiles in the vicinity of copper-rich phase decorated scale/metal


interfaces are represented in Figure 5-10. Unfortunately, no profiles were obtained at
1250°C due to difficulties in locating the small amounts of copper-rich phase formed in
these steels at this temperature. The main difference between the profiles in Figure 5-10
and the ones obtained from the copper steels is the enrichment of nickel. Apart from its
enrichment within the copper-rich phase, it appeared to be most highly enriched at the
copper-rich phase/metal interface.

149
Enrichment Behaviour of Residual Elements during Oxidation

(a) Scale Cu-rich phase Metal

100
90
80
Concentration (at%)

70
Fe
60 O
50 Cu
40 Sn
Ni
30
20
10
0
0 2 4 6 8 10 12 14
Distance (µ m)

(b) Scale Cu-rich phase Metal Internal oxide Metal Internal oxide

100
90
80
Concentration (at%)

70
Fe
60 O
50 Cu
40 Sn
Ni
30
20
10
0
0 2 4 6 8 10 12 14 16
Distance (µ m)

Figure 5-10: Concentration profiles near the scale/metal interface at 1050 and
1150°C: 0.46Cu-0.22Ni oxidised for (a) 30 minutes at 1050°C, (b) 15 minutes at
1150°C.

150
Enrichment Behaviour of Residual Elements during Oxidation

Figure 5-11 illustrates typical concentration profiles developed at 1250°C. On rare


occasions, copper-rich phase particles were observed at the interface in these particular
examples. However these profiles illustrate enrichment of elements in regions without a
copper-rich phase at the scale/metal interface. Both copper and nickel were found to
enrich in the subsurface region of the metal. As with the copper steels, the enrichment
depth in the metal at 1250°C was significantly higher than that observed at
1050/1150°C.

As previously mentioned, the 0.17Cu-0.20Ni steel did not form a copper-rich phase at
1150 or 1250°C. Concentration profiles at 1150 and 1250°C differed from the other
copper/nickel steels in that the concentration of nickel was higher than copper. A typical
example is shown in Figure 5-12.

151
Enrichment Behaviour of Residual Elements during Oxidation

(a) Scale Metal Internal oxide Metal Internal oxides Metal

60

50
Concentration (at%)

40
Fe
O
30
Cu
Ni
20

10

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
Distance (µ m)

(b) Scale Metal

5
Concentration (at%)

Cu
3
Ni

0
0 5 10 15 20 25
Distance (µ m)

Figure 5-11: Concentration profiles near the scale/metal interface at 1250°C; (a)
0.46Cu-0.22Ni oxidised for 90 minutes and (b) 0.48Cu-0.13Ni oxidised for 90
minutes at 1250°C.

152
Enrichment Behaviour of Residual Elements during Oxidation

Scale Metal Internal oxide Metal

60

50
Concentration (at%)

40 Fe
O
30 Cu
Sn
20 Ni

10

0
0 2 4 6 8 10 12
Distance (µ m)

Figure 5-12: Concentration profiles near the scale/metal interface of 0.17Cu-0.20Ni


oxidised for 15 minutes at 1150°C.

5.2.2.4 Copper/Silicon Steel

Figure 5-13 illustrates the appearance of the copper-rich phase at the scale/metal
interface in the copper/silicon steel. The copper-rich phase was observed in the form of
discrete particles at all temperatures. Penetrations of copper-rich phase within the metal
were observed at 1050 and 1150°C. The morphology of the scale/metal interface varied
with the nature of the subscale. At 1050°C when the FeO-fayalite subscale was solid,
the interface was microscopically rough (see Figure 5-13a). At 1150°C the subscale was
liquid and the interface had a wavy appearance (see Figure 5-13b). The subscale was
also liquid at 1250°C, however the interface differed, as the subscale also protruded into
the metal (Figure 5-13c).

153
Enrichment Behaviour of Residual Elements during Oxidation

(a)

(b)

(c)

Figure 5-13: Appearance of copper-rich phase at scale/metal interface in the


copper/silicon steel; 0.48Cu-0.52Si oxidised for (a) 3 hours at 1050°C, (b) 45
minutes at 1150°C and (c) 2 hours at 1250°C.

154
Enrichment Behaviour of Residual Elements during Oxidation

The copper-rich phase composition developed by the 0.48Cu-0.52Si steel at 1150°C is


summarised in Table 5-5. Reliable results at 1050 and 1250°C could not be obtained as
large enough particles for analysis could not be identified. Comparison of this result
with those in Table 5-3 and Table 5-4 illustrates that the copper-rich phase of the
copper/silicon steel was similar in composition to those found in the copper steels.

Table 5-5:Copper-rich phase composition of 0.48Cu-0.52Si steel at 1150°C

Element Cu Fe Sn
Composition (at%) 91.6 5.4 3.0

As with the copper/nickel steels, occluded particles were observed in the scale. They
were most evident at 1150°C and were often associated with the subscale that had
penetrated deep within the iron oxide scale. Figure 5-14 illustrates a typical example,
with the occluded particles being the bright particles observed both in the subscale and
iron oxide scale itself.

Typical concentration profiles in the vicinity of the scale/metal interface are represented
in Figure 5-15. The development of a copper-rich phase at 1150°C is shown in Figure
5-15a. Due to the scarcity and small size of the copper-rich phase particles developed at
1250°C, no reliable concentration profiles could be obtained. As a result, the profile
shown in Figure 5-15b is of a region where no copper-rich phase was present. As with
the other model steels at 1050 and 1150°C, the copper profile within the metal is
shallow (see Figure 5-15a). The situation differs at 1250°C, where copper enrichment
occurs over a wider distance (see Figure 5-15b).

155
Enrichment Behaviour of Residual Elements during Oxidation

Penetrating
subscale and
occluded particles

Gap

Metal
50µm

Figure 5-14: Occluded particles associated with penetrating subscale in 0.48Cu-


0.52Si oxidised for 1 hour at 1150°C.

156
Enrichment Behaviour of Residual Elements during Oxidation

(a) Scale Fayalite Cu-rich phase Metal

100
90
80
Concentration (at%)

70
Fe
60 O
50 Cu
40 Sn
Si
30
20
10
0
0 2 4 6 8 10 12 14 16 18
Distance (µ m)

(b) Scale Fayalite Metal Internal oxide Metal

4
Concentration (at%)

3
Cu
Si
2

0
0 2 4 6 8 10 12 14 16 18 20 22
Distance (µ m)

Figure 5-15: Concentration profiles near the scale/metal interface of 0.48Cu-0.52Si


oxidised for (a) 15 minutes at 1150°C and (b) 90 minutes at 1250°C.

157
Enrichment Behaviour of Residual Elements during Oxidation

5.2.3 Amount of Copper-rich Phase

The method used to measure the amount of copper-rich phase at the scale/metal
interface was described in Section 3.3. The following parameter, ACu, was introduced to
describe the analysis:

Surface area of copper - rich phase


ACu = (3.1)
Scale - metal interface length

The surface area of the copper-rich phase was determined by image analysis of BSE
images. Since the area is expressed as a fraction of the scale/metal interface length, it
can be thought of as an average thickness. Nonetheless, the copper-rich phase was not
distributed uniformly at the scale/metal interface. It was commonly observed as discrete
particles or a semi-continuous layer (see Section 5.2.1). However, the perception of ACu
as an average thickness will be useful later in this chapter when it is compared against
predicted values that assume the formation of a uniform, continuous layer.

In the following figures, ACu has been plotted as a function of scale thickness. Along
with the measured data, predicted copper-rich layer thicknesses have also been
included. These are based on a simplified mass balance calculation for an iron-copper
alloy. It was assumed that all copper was rejected from the scale and incorporated into
the copper-rich layer, i.e. no diffusion of copper within the metal occurred. This
calculation shows the maximum copper-rich layer thickness that could form. Details of
this calculation have been placed in Appendix B.1. The predicted thicknesses have been
labelled in the following manner; for a 0.47wt%Cu-Fe alloy, the predicted thickness is
labelled 0.47Cu-Fe.

Results for the copper steels at various temperatures are summarised in Figure 5-16.
The error bars in this figure and those following represent the standard deviation in ACu.
There is significant variation in the results from the 0.47Cu steel. This is a reflection of
the degree of non-uniformity in the distribution of copper-rich phase at the interface.
For all temperatures the 0.47Cu steel produced more copper-rich phase than the 0.19Cu
steel. Obviously, an increase in the copper content of the steel resulted in the formation
of more copper-rich phase. The maximum amount of copper-rich phase was produced at

158
Enrichment Behaviour of Residual Elements during Oxidation

1150°C, while the least amount was produced at 1050°C. At 1050°C, the measured
amount of copper-rich phase for the 0.47Cu steel was far less than the maximum
predicted value, whereas at 1150 and 1250°C, the measured amounts were similar to the
maximum predicted. For the 0.19Cu steel, the measured amount was less than the
predicted value at 1050 and 1150°C, while no copper-rich phase formed at 1250°C.

The effect of adding nickel on the amount of copper-rich phase formed varied with
temperature (see Figure 5-17). At 1050°C, all copper and copper/nickel steels formed a
copper-rich phase at the scale/metal interface. In general, more copper-rich phase was
measured in the copper/nickel steels than in the copper steels. At 1150°C, a copper-rich
phase was found to form in all but the 0.17Cu-0.20Ni steel. The amount of copper-rich
phase measured in the 0.48Cu-0.13Ni and 0.46Cu-0.22Ni steels were similar to that in
the 0.47Cu steel. At 1250°C, a copper-rich phase was observed in the 0.48Cu-0.13Ni
and 0.46Cu-0.22Ni steels only after a considerably longer oxidation time than that
observed in the 0.47Cu steel. The amount of copper-rich phase was lower in these steels
than in the 0.47Cu steel. In general, at 1050 and 1150°C the measured amounts of
copper-rich phase were similar, whereas at 1250°C it was significantly lower than the
maximum predicted value.

The effect of silicon addition on the formation of copper-rich phase is illustrated in


Figure 5-18. At 1050°C, the results indicate little difference between the amount of
copper-rich phase measured in the 0.47Cu and 0.48Cu-0.52Si steels. At 1150°C, the
effect of silicon addition was to decrease the amount of copper-rich phase. At 1250°C, a
copper-rich phase was determined to have formed between 90 and 120 minutes of
oxidation for the 0.48Cu-0.52Si steel, while it formed between 30 and 60 minutes in the
0.47Cu steel. The amount of copper-rich phase was lower in the 0.48Cu-0.52Si steel
than in the 0.47Cu steel. In general, the amount of copper-rich phase formed in the
0.48Cu-0.52Si steel was significantly lower than the maximum predicted value at all
temperatures.

159
Enrichment Behaviour of Residual Elements during Oxidation

(a)
0.50
0.45
A Cu ( m 2/ m interface) 0.40
0.35 0.19Cu
0.30 0.47Cu
0.25 0.19Cu-Fe
0.20 0.47Cu-Fe
0.15
0.10
0.05
0.00
0 200 400 600 800 1000 1200 1400

Scale thickness (µ m)
(b)
3.5

3.0
A Cu ( m 2/ m interface)

2.5
0.19Cu
2.0 0.47Cu
1.5 0.19Cu-Fe
0.47Cu-Fe
1.0

0.5

0.0
0 200 400 600 800
Scale thickness (µ m)
(c)
1.600
1.400
A Cu ( m 2/ m interface)

1.200
1.000
0.47Cu
0.800
0.47Cu-Fe
0.600
0.400
0.200
0.000
0 100 200 300 400 500
Scale thickness (µ m)

Figure 5-16: Relative amount of copper-rich phase for copper steels at (a) 1050°C,
(b) 1150°C and (c) 1250°C.

160
Enrichment Behaviour of Residual Elements during Oxidation

(a)
8
7
A Cu ( m 2/ m interface) 0.47Cu
6
0.17Cu-0.20Ni
5 0.46Cu-0.22Ni
4 0.48Cu-0.13Ni
3 0.47Cu-Fe
0.17Cu-Fe
2
1
0
400 600 800 1000 1200

Scale thickness (µ m)
(b)
5
A Cu ( m 2/ m interface)

4
0.47Cu
3 0.48Cu-0.13Ni
0.46Cu-0.22Ni
2
0.47Cu-Fe

0
200 300 400 500 600 700

Scale thickness (µ m)
(c)
1.0
0.9
A Cu ( m 2/ m interface)

0.8
0.7 0.47Cu
0.6 0.48Cu-0.13Ni
0.5 0.46Cu-0.22Ni
0.4
0.47Cu-Fe
0.3
0.2
0.1
0.0
0 100 200 300 400 500

Scale thickness (µ m)

Figure 5-17: Relative amount of copper-rich phase for copper/nickel steels at (a)
1050°C, (b) 1150°C and (c) 1250°C.

161
Enrichment Behaviour of Residual Elements during Oxidation

(a)
1.0
0.9
A Cu ( m 2/ m interface)
0.8
0.7
0.47Cu
0.6
0.48Cu-0.52Si
0.5
0.47Cu-Fe
0.4
0.3
0.2
0.1
0.0
200 400 600 800 1000 1200
Scale thickness (µ m)
(b)
3.5

3.0
A Cu ( m 2/ m interface)

2.5
0.47Cu
2.0
0.48Cu-0.52Si
1.5 0.47Cu-Fe
1.0

0.5

0.0
400 450 500 550 600 650 700

Scale thickness (µ m)
(c)
0.030

0.025
A Cu ( m 2/ m interface)

0.020
0.47Cu
0.015 0.48Cu-0.52Si
0.47Cu-Fe
0.010

0.005

0.000
0 50 100 150 200

Scale thickness (µ m)

Figure 5-18: Relative amount of copper-rich phase for copper/silicon steel at (a)
1050°C, (b) 1150°C and (c) 1250°C.

162
Enrichment Behaviour of Residual Elements during Oxidation

5.3 Discussion

5.3.1 Copper-rich Phase Morphology and Distribution

5.3.1.1 Copper-rich Phase at Scale/Metal Interface

The copper-rich phase was found to be distributed non-uniformly at the scale/metal


interface and existed as discrete particles and/or as a semi-continuous layer (see Section
5.2.1). If a continuous, uniform copper-rich layer were to form, the following conditions
would need to be satisfied:

• Metal consumption would need to be uniform, leading to uniform enrichment of


residual elements and a flat scale/metal interface.
• Either all residual elements are rejected from the scale, or occlusion occurs
uniformly. That is, the amount of residual elements accumulated at the
scale/metal interface must be uniform.
• Diffusion of residual elements back into the metal is not favoured at particular
areas, and hence the diffusion is uniform. That is, the amount of residual
elements leaving the scale/copper-rich phase interface must be uniform.

Clearly, very few of these conditions were satisfied in the model steels investigated.
Many of the model steels had non-planar scale/metal interfaces, particularly the
copper/nickel and copper/silicon steels. The distribution of occluded particles was not
measured, but it is likely that occlusion did not occur uniformly across the scale/metal
interface as oxidation proceeded. Finally, diffusion of enriched elements is most likely
to be favoured at metal grain boundaries and so this process was also non-uniform. It is
therefore not surprising that the distribution and morphology of the copper-rich phase
was non-uniform.

In addition to the factors discussed above, wettability between the scale and metal
would also affect the distribution and morphology of the copper-rich phase. As
discussed in Section 2.2.3.1.4, there is no reported data of the wettability of iron oxides
with copper-rich phases. Generally, the wettability of oxides with liquid metals is low

163
Enrichment Behaviour of Residual Elements during Oxidation

[130,131]. Assuming this behaviour exists between liquid copper and FeO, one would
expect drops to be favoured over a continuous copper-rich film. However, dihedral
angle and sessile drop measurements of liquid copper and iron substrates between 900-
1300°C [114,119,120] suggest that liquid copper readily wets iron at these
temperatures. Since the copper-rich phase precipitates between the scale and metal, and
the wettability is expected to be very different between these two phases, it is unclear
from a wettability point of view how a copper-rich phase would appear. However, it can
be said that a copper-rich phase is more likely to wet and therefore penetrate the metal
than the scale.

5.3.1.2 Copper-rich Phase in Scale

In the case of the copper and copper/nickel steels, occluded particles were observed at
all temperatures. They were most frequently observed at 1150°C, and appeared more
numerous in the copper/nickel steels than the copper steels. It is generally accepted that
occlusion occurs when internal oxides coalesce and join with the external scale, leading
to engulfment of the enriched metal surface [88,106]. The images in Figure 5-19, taken
in different areas of the same specimen, support this mechanism. Figure 5-19a shows
copper-rich phase at the scale/metal interface with some internal oxides coalescing and
joining with the external scale. It would seem quite likely that as further scale growth
continues, both the internal oxides and copper-rich phase would become trapped and
incorporated into the scale. Figure 5-19b shows occluded particles in the scale.

Fukagawa and Fujikawa [128] proposed that occlusion occurs when the scale/metal
interface becomes irregular. They found that nickel addition and increased temperature
enhanced the irregularity of the interface. Akamatsu et al. [103] also found that nickel
addition increases the irregularity of the scale/metal interface while others have
indicated that nickel promotes internal oxidation [92,106]. Interface irregularity and
internal oxidation of iron are caused by the same phenomenon. The enrichment of
residual elements during oxidation reduces the iron activity in the subsurface region.
This provides a condition favourable for internal oxidation and/or breakdown of the
interface. As the scale layer grows inward, internal oxides contact the scale/metal
interface which further enhances its irregularity.

164
Enrichment Behaviour of Residual Elements during Oxidation

(a)

(b)

Figure 5-19: Evidence of (a) internal oxides coalescing and impinging with external
scale, causing (b) occluded particles (0.46Cu-0.22Ni oxidised for 15 minutes at
1150°C).

The amount of internal oxidation was measured for a selection of model steels under
various oxidation conditions. The measurement was conducted in a similar manner for
determining the amount of copper-rich phase; the surface area of the internal oxides
were measured using image analysis and expressed relative to the length of the
scale/metal interface analysed. The results are shown in Figure 5-20. It is clear that the
copper/nickel steels internally oxidised more readily than the copper steel. In addition,
internal oxidation occurred more readily at higher temperatures.

165
Enrichment Behaviour of Residual Elements during Oxidation

Surface area of internal oxides per


14

unit scale/metal interface length


12

10

( m 2/ m)
0.47Cu
8
0.17Cu-0.20Ni
6
0.46Cu-0.22Ni
4

0
60 mins. at 15 mins. at 90 mins. at
1050°C 1150°C 1250°C
Oxidation conditions

Figure 5-20: Amount of internal oxidation as a function of alloy composition and


oxidation conditions.

If it is assumed that the amount of internal oxidation affects the amount of occlusion,
one would expect to observe more occlusion (alternatively; less copper-rich phase at the
metal surface) in the copper/nickel steels than the copper steels and more occlusion at
higher temperatures. The image analysis results (see Section 5.2.3) showed that, in
general, the amount of copper-rich phase was unaffected by nickel addition at 1050 and
1150°C, whereas it was significantly lower at 1250°C. One notable exception was the
0.17Cu-0.20Ni steel at 1150°C in which no copper-rich phase was detected. Enrichment
of nickel would be expected to increase the copper solubility in austenite, leading to
more copper enriched within the metal during oxidation. However, this was not
observed, indicating that occlusion was responsible. Hence, occlusion was most
significant when the amount of internal oxidation was high, as was the case in the
copper/nickel steels oxidised at 1250°C. At the lower temperatures, the relationship
between the amount of internal oxidation and occlusion was less clear. Occlusion was
significant in the 0.17Cu-0.20Ni steel at 1150°C, but not in the 0.48Cu-0.13Ni and
0.46Cu-0.22Ni steels.

In the case of the copper/silicon steel, occluded particles were observed at all
temperatures, but were most evident at 1150°C. The particles were often associated with
the significant quantity of liquid subscale that formed at this temperature. Kajitani et al.

166
Enrichment Behaviour of Residual Elements during Oxidation

[107] proposed that the presence of a liquid fayalite subscale decreased the interfacial
energy between the scale and copper-rich phase, leading to its occlusion. However
occlusion was also observed when the subscale was solid. The mechanism by which this
occurred is similar to that in the copper and copper/nickel steels. The scale/metal
interface of the copper/silicon steel at 1050°C was rough and an internal oxide zone was
present. Copper-rich phase particles could become trapped between the inward moving
scale/metal interface and internal oxides. Further scale growth would lead to the copper-
rich phase and internal oxides becoming engulfed and incorporated into the scale.

In Section 5.2.1, the presence of copper-enriched particles at or near the outer scale
surface was reported. These were observed in the model steels oxidised in air at 1050
and 1150°C, but not in those oxidised in the CO2-N2 atmosphere at 1250°C. The rate of
scale growth was rapid in air and slow in the CO2-N2 atmosphere (see Table 4-1). A
copper-rich phase forms quickly when scale growth is rapid. It is thought that in the
early stages of oxidation in air, the copper-rich particles became engulfed by the
growing scale. Copper enrichment at the scale/gas interface has been observed in
another investigation. In their observations of a 4.54wt%Cu-Fe alloy oxidised in oxygen
at 700°C for 6 and 18 hours, Nanni and Gesmundo [53] identified a slight enrichment of
copper at the scale/gas interface using BSE imaging.

5.3.2 Residual Element Enrichment

The results in Table 5-3 show that the concentration of copper in the copper-rich phase
decreases with increasing temperature. With reference to the iron-copper phase diagram
(see Figure 5-21), it can be seen that the copper concentration of the γ + Liq. / Liq.
phase boundary does decrease with increasing temperature. However, the copper
concentration in the copper-rich phase cannot fully be defined by the γ + Liq. / Liq.
phase boundary alone. In fact, it would be expected to vary between the phase boundary
concentrations being set by the γ + Liq. / Liq. and the Liq. / Liq. + FeO equilibrium
concentrations. Unfortunately, no data is available for the effect of temperature on the
Liq. / Liq. + FeO equilibrium.

167
Enrichment Behaviour of Residual Elements during Oxidation

1300

1250

Temperature (°C)
γ + Liq. Liq.
1200

1150

1100

1050 γ+ε ε

1000
85 90 95 100
Atomic Percent Copper

Figure 5-21: Iron-rich end of iron-copper phase diagram calculated using Thermo-
Calc ver. M [126] and an Fe-Cu-Ni-Sn database [144].

The presence of a copper-rich phase along metal grain boundaries suggests that it was
molten at the time of penetration. This was observed at all temperatures for most model
steels. The melting point of the copper-rich phase predicted from the iron-copper phase
diagram is 1096°C [57]. As copper-rich phase was observed to have penetrated grain
boundaries at 1050°C, it is obvious that its melting point was significantly lower than
1096°C. As indicated in Table 5-3 - Table 5-5, the copper-rich phase contains a
significant amount of tin. The presence of tin is known to decrease the melting point of
the copper-rich phase [70]. This is illustrated by reference to the copper-tin phase
diagram (see Figure 5-22). The degree to which the presence of tin decreases the
melting point of the copper-rich phase is summarised in Table 5-6. A 3at% tin addition
within the copper-rich phase is predicted to decrease the melting point below 1050°C.

168
Enrichment Behaviour of Residual Elements during Oxidation

1100
1050

Temperature (°C)
1000 Liq.
950
900
850
800
ε
750
β
700
0 5 10 15 20
Atomic Percent Tin

Figure 5-22: Calculated copper-rich end of copper-tin phase diagram using


Thermo-Calc [126] and an Fe-Cu-Ni-Sn database [144].

Table 5-6: Melting point of the copper-rich phase in a copper-tin alloy predicted
using Thermo-Calc [126] and an Fe-Cu-Ni-Sn database [144]

Tin composition in Melting point of


Cu-rich phase (at%) Cu-rich phase (°C)
0 1085
1 1073
3 1042
5 1006
10 904

In Section 5.2.2.2 it was observed that the enrichment depth of copper within the metal
at 1050 and 1150°C was in the order of 1-2µm. However, at 1250°C the enrichment
depth was significantly higher (5-20µm). Similar observations were observed for the
copper/nickel and copper/silicon steels. These observations can be discussed with
reference to Figure 5-23. At the lower temperatures the diffusivity of copper is
decreased, favouring smaller penetration depths. In addition, the rate of metal
consumption was far higher under these conditions than that at 1250°C. As a result, the
motion of the scale/metal interface was rapid compared to the diffusion of copper away
from the enriched region into the metal at the lower temperatures. Hence, the enriched
copper was more likely to be incorporated into a copper-rich layer at the lower
temperatures. This would result in less copper being enriched within the metal and

169
Enrichment Behaviour of Residual Elements during Oxidation

therefore smaller penetration depths. This result is important because it illustrates that
significantly more copper was incorporated within the metal under the conditions at
1250°C than those at 1050 and 1150°C. When this observation is combined with the
fact that occlusion became more obvious at higher temperatures, conditions which
inhibit the formation of a copper-rich phase at the scale/metal interface may be
identified.

(a) Oxidation at 1050/1150°C in air: Scale Metal


• More metal consumed
• Lower diffusivity of copper in
metal

Cu-rich phase

(b) Oxidation at 1250°C in CO2-N2:

Concentration
• Less metal consumed
• Higher diffusivity of copper in
metal

Distance

Figure 5-23: Schematic copper concentration profiles when oxidised in (a) air at
1050°C/1150°C and (b) CO2-N2 at 1250°C.

The measured compositions of the copper-rich phase of the copper steels were
summarised in Table 5-3. These compositions have been projected onto Fe-Cu-Sn
isotherms calculated using Thermo-Calc [126]. The measured compositions were found
to lie close to the γ + Liq. / Liq. phase boundary, and a typical example is illustrated in
Figure 5-24. Considering that the alloys investigated were steels and not comparably
simpler Fe-Cu-Sn alloys, the composition can be considered to lie on the phase
boundary. This being the case, it would be logical to determine the composition at the γ
/ γ + Liq. phase boundary, defined by the tie-line that the measured composition lies
upon in the Fe-Cu-Sn isotherm. This was used to predict the copper solubility limit in
austenite as it could not be measured directly by EPMA in the present work due to
errors associated with reporting compositions at phase boundaries. The predicted
solubility limits, along with those predicted from the iron-copper phase diagram have
been summarised in Table 5-7. The difference between the solubility limit predicted

170
Enrichment Behaviour of Residual Elements during Oxidation

from the Fe-Cu-Sn isotherm and that from the iron-copper phase diagram is not
significant.

Liq1 + Liq2
50

in
40

tT
en
30

rc
Pe
α + γ + Liq1
ic 20
om
At

10
γ + Liq1 Liq2
α
0
0 20 40 60 80 100

γ Atomic Percent Copper

Figure 5-24: Composition of copper-rich phase (S) in 0.47Cu steel at 1150°C


projected on an Fe-Cu-Sn isotherm calculated using Thermo-Calc [126] and an Fe-
Cu-Ni-Sn database [144].

Table 5-7: Predicted solubility limit of copper (wt%) in austenite for 0.47Cu steel

Temperature Using experimental results on an From the iron-copper


(°C) Fe-Cu-Sn isotherm phase diagram
1050 6.1%Cu-93.8%Fe 6.2%Cu
1150 7.6%Cu-92.4%Fe 8.0%Cu
1250 8.9%Cu-91.1%Fe 9.5%Cu

Only one copper-rich phase composition was reported for the copper/silicon steel (see
Table 5-5). It was very similar in magnitude to those measured in the copper steels. This
was not the case for the copper/nickel steels. The copper-rich phase developed by these
steels were found to contain a significant quantity of nickel (see Table 5-4). These
compositions have been projected onto calculated Fe-Cu-Ni isotherms by neglecting the
tin component. As before, the measured compositions were found to lie near the γ + Liq.
/ Liq. phase boundary (see Figure 5-25). Using the same method adopted for the 0.47Cu
steel, the copper solubility limit in austenite was estimated using the experimentally-
determined copper-rich phase composition and the calculated isotherm (see Table 5-8).
The result at 1050°C is not significantly different from that obtained for the copper

171
Enrichment Behaviour of Residual Elements during Oxidation

steel. However, the results at higher temperatures suggest that the copper solubility limit
is greatly increased. As such, the iron-copper phase diagram cannot be used to predict
the solubility limit in the copper/nickel steels at 1150 and 1250°C. An increase in the
solubility limit would be expected to affect copper-rich phase formation. However, the
likelihood of copper-rich phase formation is determined by both thermodynamic (the
copper solubility in austenite) and kinetic (the relative rates of oxidation and copper
back-diffusion) factors. It is suspected that the change in solubility limit would not
greatly affect copper-rich phase formation when the oxidation rate is rapid compared to
copper back-diffusion. These aspects will be investigated using diffusion modelling in
Chapter 6.

50
el
ick

40 γ1
tN
en

30 γ1 + γ2 + Liq.
rc
Pe
ic

20
om
At

10 γ1 + Liq. Liq.

0
0 20 40 60 80 100

Atomic Percent Copper

Figure 5-25: Composition of copper-rich phase (S) of 0.46Cu-0.22Ni steel at


1150°C projected on an Fe-Cu-Sn isotherm calculated using Thermo-Calc [126]
and an Fe-Cu-Ni-Sn database [144].

Table 5-8: Predicted solubility limit of copper (wt%) in austenite for 0.46Cu-
0.22Ni steel

Temperature Using experimental results on an


(°C) Fe-Cu-Ni isotherm
1050 8.1%Cu-78.8%Fe-13.1%Ni
1150 14.3%Cu-59.6%Fe-14.3%Ni

1250 13.5%Cu-68.4%Fe-18.1%Ni

Note that no composition was measured at 1250°C. This result was based on an
estimated copper-rich phase composition of 80at%Cu-7%Ni.

172
Enrichment Behaviour of Residual Elements during Oxidation

Concentration profiles near the copper-rich phase/metal interface of copper/nickel steels


(see Figure 5-10) revealed that nickel was most highly enriched at this interface. This
can easily be explained with reference to the Fe-Cu-Ni phase diagram, for example
Figure 5-25. Due to the slope of the tie-lines, the nickel concentration within the metal
at the γ / γ + Liq. phase boundary is generally much higher than that existing within the
liquid at the γ + Liq. / Liq. phase boundary. Therefore nickel enrichment within the
metal can reach much higher levels than that accumulated within the copper-rich phase.

5.3.3 Amount of Copper-rich Phase

5.3.3.1 Initial Considerations

In Section 5.2.3, the measured amount of copper-rich phase was compared against a
predicted thickness. In some cases, the measured amount was found to vary
considerably from the predicted thickness. Before discussing the results it is worth
commenting on the assumptions used to calculate the predicted thickness. The estimate
assumed that all copper is rejected from the scale during its formation. This is a
seemingly reasonable assumption as the solubility of copper in FeO is very small (see
Figure 5-26). Copper is therefore assumed to enrich at the metal surface and precipitate
as a copper-rich phase once its limited solubility in the metal has been exceeded.

To simplify the calculation, it was initially assumed that the solubility limit be exceeded
immediately. The experimental results indicated that a copper-rich phase indeed formed
rapidly when oxidation took place in air at 1050 and 1150°C. For example, significant
quantities were observed in the 0.47Cu steel after only 10 minutes oxidation in air at
1150°C. However, when oxidation was carried out in CO2-N2 at 1250°C, the formation
of a copper-rich phase was delayed for a considerable amount of time. There are two
factors behind this phenomenon. Firstly, the rate of copper enrichment was decreased
compared to the situation at the lower temperatures due to the lower oxidation rate in
CO2-N2. Secondly, more copper was incorporated in the metal due to the deeper
diffusion penetration depths. This observation was discussed qualitatively with
reference to Figure 5-23. As a result, one would expect the predicted copper-rich phase
thickness calculated for the situation at 1250°C to be overestimated.

173
Enrichment Behaviour of Residual Elements during Oxidation

Figure 5-26: Fe-Cu-O isotherm at 1200°C [17].

Lastly, all rejected copper was assumed to be incorporated within a uniform and
continuous copper-rich layer. As discussed in Section 5.3.1.1, the morphology and
distribution of the copper-rich phase was not uniform. The measured quantity was
actually an average obtained from an exhaustive analysis of multiple microstructural
images (see Section 3.3). In fact, the copper-rich phase was not only present at the metal
surface, but also as occluded particles (see Section 5.3.1.2). Hence, differences between
the measured and predicted amount of copper-rich phase give an indirect measure of the
degree of occlusion.

Before the results are discussed, there is one further point to consider. Scale separation
occurred readily in all steels oxidised at 1050 and 1150°C in air (see Section 4.2.1). In
order to properly relate the level of residual element enrichment and the amount of
consumed metal, analysis was confined to adhered scale layers. The observed non-
uniform scale growth at 1150°C indicated that scale separation and healing occurred
during reaction (see Section 4.2.3). There is no evidence to suggest that similar
behaviour occurred when oxidation was carried out in CO2-N2 at 1250°C.

174
Enrichment Behaviour of Residual Elements during Oxidation

The significance of scale separation and healing is that some of the copper-rich layer
would have dissolved while scale formation had ceased during separation. To
investigate this, an analytical equation was developed that described the time taken for a
known thickness of copper-rich layer to completely dissolve by diffusion into the metal:

2
π  L.C Cu
Cu − rich

t= 
( )

Metal / Cu − rich
(5.1)
D − C Cu
o
 2 C Cu 

where t is the time taken for dissolution of a copper-rich layer of thickness L, CCuCu-rich
is the copper concentration in the copper-rich phase, CCuMetal/Cu-rich is the copper
solubility limit and CCuo the bulk concentration of copper in the metal. A derivation of
this equation has been placed in Appendix B.2.

The measured results for copper-rich phase accumulation from the copper steels at
1050°C were significantly lower than the maximum predicted values (see Figure 5-16).
For example, ACu was measured as 0.287µm for the 0.47Cu steel after 2 hours oxidation
in air. For an equivalent scale thickness, L was calculated as 2.36µm; a difference of
2.07µm. Taking CCuCu-rich and CCuMetal/Cu-rich to equal 96.69 and 6.16 wt%Cu,
respectively (estimated from the iron-copper system using Thermo-Calc [126]), CCuo to
equal 0.47 wt%Cu, and D to equal 3.35×10-12 cm2/s [145], a 2.07µm layer was
calculated to take 805.8 hours to dissolve. Hence, it is obvious that the difference
between the observed and predicted copper-rich layer thickness cannot be accounted for
by scale separation and dissolution. The time taken for dissolution of various
thicknesses at various temperatures was calculated and has been summarised in Figure
5-27 using data summarised in Table 5-9.

175
Enrichment Behaviour of Residual Elements during Oxidation

Time taken for dissolution (hrs)


8
7
6
5 2 micron layer
4 1 micron layer
3 0.5 micron layer
2
1
0
1100 1150 1200 1250 1300
Temperature (°C)

Figure 5-27: Time taken for dissolution of various copper-rich layer thicknesses as
a function of temperature (CCuo = 0.45wt%).

Table 5-9: Relevant parameter values for dissolution calculation

Temperature (°C) D (cm2/s) † CMetal/Cu-rich (wt%) ☼ CCu-rich/Metal (wt%) ☼


1100 9.6×10-12 7.3 97.0
1150 2.3×10-11 8.0 96.1
1200 5.1×10-11 8.8 95.0
1250 1.1×10-10 9.5 93.6
2.2×10-10 130010.2 91.6

Copper diffusivity in austenite data from [140].

Calculated using Thermo-Calc [126] for an iron-copper alloy.

5.3.3.2 Measured Results

In the case of the copper steels, the measured amount of copper-rich phase from image
analysis results was far less than the maximum predicted value at 1050°C. At 1150 and
1250°C, the measured amounts were similar to the maximum predicted. As discussed
previously (see Section 5.3.3.1), this difference cannot be accounted for by scale
separation or copper-rich layer dissolution. Excluding the possibility that there was less
copper in these steels than measured, the rejected copper must be occluded within the
scale. Occluded particles were observed at this temperature, although they appeared
176
Enrichment Behaviour of Residual Elements during Oxidation

more numerous at 1150 and 1250°C. A quantitative measure of the amount of copper
occluded within the scale was not obtained. However, mass balance measurements were
performed to determine the amount of copper rejected due to metal consumption (using
scale thickness measurements) and that accumulated ahead of the scale (using
microprobe analysis). The amount of copper rejected due to metal consumption (A1),
assuming that none is accommodated within the scale, is given by:

A1 = l ⋅ N Cu ° (5.2)

where NCuo is the copper concentration in the original steel and l is the thickness of
metal consumed.

The amount of copper accumulated ahead of the scale (A2) was measured by integrating
the copper concentration profile near the scale/metal interface. Only one measurement
was performed for a copper steel at 1050°C. The result is summarised in Table 5-10,
where A2(Cu-rich) and A2(Metal) denotes the amount of copper accumulated within the
copper-rich layer and metal phases, respectively. However, since A2 is greater than A1,
the amount of accumulated copper is higher than that possible and therefore the result
must be rejected. As the bulk of the accumulated copper is contained within the copper-
rich layer, it is possible that a localised increase in the copper-rich layer thickness led to
this anomalous result. As a result, the fact that there were such large differences
between the measured and maximum predicted copper-rich layer thickness for the
copper steels at 1050°C remains unexplained.

Table 5-10: Mass balance results for 0.47Cu steel oxidised for 30 minutes at
1050°C

A1 (mol Cu/cm2) A2 (mol Cu/cm2)


A2(Cu-rich): 3.2×10-5
1.6×10-5 ± 5.0×10-7 A2(Metal): 1.8×10-7
Total: 3.2×10-5

The addition of nickel was found to have a marked effect on the formation of a copper-
rich phase at both 1150 and 1250°C. At 1150°C, no copper-rich phase was observed or

177
Enrichment Behaviour of Residual Elements during Oxidation

measured at the scale/metal interface of the 0.17Cu-0.20Ni steel. Only small amounts of
copper-rich phase were measured for the 0.19Cu steel at the same temperature. Mass
balance results for the 0.17Cu-0.20Ni steel oxidised at 1150°C are summarised in Table
5-11. The results indicate that a considerable quantity of copper is unaccounted for at
1150°C and must be located within the scale. Therefore occlusion was significant and
removed a sufficient amount of copper-rich phase so that none was observed at the
scale/metal interface.

Table 5-11: Mass balance results for 0.19Cu and 0.17Cu-0.20Ni steels at 1150 and
1250°C

Condition A1 (mol Cu/cm2) A2 (mol Cu/cm2)


0.17Cu-0.20Ni steel oxidised for 15
4.3×10-6 ± 1.2×10-6 6.2×10-7 ± 2.3×10-7 †
minutes at 1150°C
0.19Cu steel oxidised for 90 minutes
3.1×10-6 ± 2.3×10-7 4.5×10-6 ☼
at 1250°C
0.17Cu-0.20Ni steel oxidised for 90
5.1×10-6 ± 7.0×10-7 2.3×10-6 ☼
minutes at 1250°C

Average of two concentration profiles

Result from a single concentration profile

At 1250°C, a copper-rich phase was not observed at the scale/metal interface of the
0.19Cu and 0.17Cu-0.20Ni steels. The mass balance results are summarised in Table
5-11. Although only one EPMA scan was made on each at 1250°C, it will be assumed
that the fractional error in A2 was the same as in the measurement made at 1150°C. The
measured value for A2 was greater than A1 for the 0.19Cu steel, but may be within
experimental error. No conclusion is possible. The result for the 0.17Cu-0.20Ni steel
indicates that occlusion was significant and removed a sufficient amount of copper-rich
phase so that none was observed at the scale/metal interface.

While a copper-rich phase was observed at the scale/metal interface of the 0.48Cu-
0.13Ni and 0.46Cu-0.22Ni steels at 1250°C, it took a considerably longer time to appear
than in the case of the 0.47Cu steel. Relevant mass balance results are summarised in
Table 5-12. The result for the 0.47Cu steel indicates more copper accumulated than
predicted. It is possible that this result is not typical due to the unusually large amount
of copper measured within the metal. For other samples oxidised for 90 minutes at

178
Enrichment Behaviour of Residual Elements during Oxidation

1250°C, the amount of copper measured within the metal varied between 2×10-6 and
4.5×10-6 mol Cu/cm2. If the amount of copper in the metal were of this magnitude, A2
would be approximately equivalent to A1. As a result, occlusion is thought not to have
been significant in this case. However, the results for the 0.46Cu-0.22Ni steel indicate a
significant quantity of copper to have been occluded into the scale. It can be concluded
that occlusion and the reduced amount of metal consumption in the copper-nickel steel
contributed to a copper-rich phase taking considerably longer to form.

Table 5-12: Mass balance results for 0.47Cu and 0.46Cu-0.22Ni steels at 1250°C

Condition A1 (mol Cu/cm2) A2 (mol Cu/cm2)


A2(Cu-rich): 1.2×10-5
0.47Cu steel oxidised for 90
1.6×10-5 ± 8.8×10-7 A2(Metal): 1.1×10-5
minutes at 1250°C
Total: 2.2×10-5 †
0.46Cu-0.22Ni steel oxidised for 90
8.4×10-6 ± 7.4×10-7 1.9×10-6 †
minutes at 1250°C

Result from a single concentration profile

The effect of adding silicon was significant at 1150 and 1250°C. In each of these cases,
the 0.48Cu-0.52Si steel formed less copper-rich phase at the interface than the 0.47Cu
steel. In general, the amount of copper-rich phase in the 0.48Cu-0.52Si steel determined
from the image analysis results was far lower than the maximum predicted values.
Relevant mass balance results are summarised in Table 5-13. The results at 1150°C for
both steels show an average measured A2 result higher than A1. However, the
measurement error in the result for the 0.47Cu steel is very large. In the case of the
0.48Cu-0.52Si steel, the error in the A2(Cu-rich phase) measurement can be estimated
from the standard deviation in the measured thickness of the copper-rich phase of this
sample. This was 1.6µm after 15 minutes oxidation, which is equivalent to 1.1×10-5 mol
Cu/cm2. Since both sets of analysis have large errors, the difference between A2 and A1
cannot be considered significant. However, the result for the 0.48Cu-0.52Si steel at
1250°C is conclusive and indicates that occlusion made a significant contribution.
Hence, the mass balance results do not indicate why the amount of copper-rich phase
was reduced upon silicon addition at 1150°C, but do suggest that occlusion was
significant at 1250°C.

179
Enrichment Behaviour of Residual Elements during Oxidation

Table 5-13: Mass balance results for 0.47Cu and 0.48Cu-0.52Si steels

Condition A1 (mol Cu/cm2) A2 (mol Cu/cm2)


A2(Cu-rich): 2.6×10-5 ± 2.2×10-5
0.47Cu steel oxidised for
1.7×10-5 ± 1.1×10-6 A2(Metal): 2.8×10-6 ± 2.7×10-6
15 minutes at 1150°C
Total: 2.9×10-5 ± 2.2×10-5 †
0.48Cu-0.52Si steel A2(Cu-rich): 2.0×10-5
oxidised for 15 minutes at 1.8×10-5 ± 1.8×10-6 A2(Metal): 5.4×10-7
1150°C Total: 2.1×10-5 ☼
0.47Cu steel oxidised for
See Table 5-12 See Table 5-12
90 minutes at 1250°C
0.48Cu-0.52Si steel
oxidised for 90 minutes at 6.3×10-6 ± 1.0×10-6 1.2×10-6 ☼
1250°C

Average of five concentration profiles

Result from a single concentration profile

5.4 Summary and Conclusions

The copper-rich phase was found to be distributed non-uniformly at the scale/metal


interface and existed as discrete particles and/or as a semi-continuous layer. The
formation of a continuous, uniform copper-rich layer is not expected since the
scale/metal interface was not smooth, the accumulation of residual elements at the
interface was likely to be non-uniform (due to effects of occlusion), and diffusion of
enriched elements into the metal from the interface is likely to be non-uniform. The
likelihood of poor wettability between the copper-rich phase and FeO would also favour
discrete particle formation.

Occluded particles were observed in all copper and copper/nickel steels at all
temperatures. Microstructural observations supported that occlusion occurs when
copper-rich phase particles become entrapped between an inward migrating irregular
scale/metal interface and internal oxides. As further scale growth continues, both the
copper-rich phase particles and internal oxides become incorporated into the scale.
Occlusion was most significant when the amount of internal oxidation was high, as was
the case in the copper/nickel steels oxidised at 1250°C.

Occlusion appeared significant in the copper/silicon steel and was especially evident at
1150°C when significant amounts of liquid subscale were formed. Particles were readily

180
Enrichment Behaviour of Residual Elements during Oxidation

observed associated with the liquid subscale that had penetrated deep within the scale at
this temperature. As proposed by Kajitani et al. [107], it is believed that the liquid
subscale decreases the interfacial energy between the scale and the copper-rich phase.
Copper-rich phase particles are then able to penetrate the scale and become occluded.
Some occlusion was also evident when the subscale was solid. It is believed that the
occlusion mechanism in this case is similar to that occurring in the copper and
copper/nickel steels, i.e. increased scale/metal interface irregularity and levels of
internal oxidation leading to entrapment and engulfment by the inward migrating
external scale.

Oxidation conditions were found to have a significant effect on the enrichment of


copper within the model steels. When oxidation took place in air at 1050 and 1150°C,
significant quantities of copper-rich phase were observed for most steels. Under these
conditions, the oxidation rate (and thus enrichment of residual elements) was rapid
compared to the diffusion of residual elements back into the metal. The solubility limit
was exceeded rapidly and copper-rich phase accumulated at the scale/metal interface or
was occluded into the scale. When oxidation was carried out in CO2-N2 at 1250°C, the
oxidation rate was considerably lower than that at the lower temperatures. Along with
the reduction in residual element enrichment, more copper could be incorporated within
the metal. These conditions contributed to the fact that the copper-rich phase was not
observed for a significant amount of time; and in some cases not at all. As a result,
oxidation conditions may be used to inhibit the formation of a copper-rich phase at the
scale/metal interface.

Copper solubility limits were predicted from ternary phase diagrams using experimental
results and revealed that the iron-copper phase diagram could be used to predict the
solubility limit for the copper and copper/silicon steels in the range 1050-1250°C. It
could not be used between 1150 and 1250°C for the copper/nickel steels, as significant
nickel enrichment led to an increase in the copper solubility within this temperature
range. It is suspected that the effect of an increase in solubility limit with nickel addition
on copper-rich phase formation only becomes significant when the oxidation rate and
copper diffusion within the metal are similar. This will be investigated using diffusion
modelling in Chapter 6.

181
Enrichment Behaviour of Residual Elements during Oxidation

The addition of nickel and silicon were found to have a marked effect on the amount of
copper-rich phase measured at the scale/metal interface at 1150 and 1250°C. For
example, no copper-rich phase was detected at 1150°C in the 0.17Cu-0.20Ni steel. It
was concluded that the level of occlusion in this steel was sufficient to remove all
copper-rich phase from the interface. At 1250°C, smaller quantities of copper-rich phase
were observed after a considerably longer oxidation time in the 0.48Cu-0.13Ni and
0.46Cu-0.22Ni steels compared to the 0.47Cu steel. Once again, occlusion was found to
be significant in the copper/nickel steels. In the case of silicon addition, less copper-rich
phase was measured at the interface of the 0.48Cu-0.52Si steel compared to the 0.47Cu
steel at 1150 and 1250°C. Although the results were inconclusive at 1150°C, occlusion
was found to be significant in the 0.48Cu-0.52Si steel at 1250°C.

It is therefore apparent that both oxidation conditions and alloying additions may be
used to reduce or completely suppress the formation of a copper-rich phase at the
interface. It should be noted that promoting occlusion by either liquid subscale
formation or through increased internal oxidation is likely to be detrimental with respect
to scale adhesion. Both nickel [84,146] and silicon [85,86] have been shown to decrease
the descalability of steel. If the occlusion mechanism were to be used to inhibit surface
hot shortness, one would be compromising the descalability properties of the steel.

182
CHAPTER
SIX
Modelling of Residual Element Enrichment

6.1 Introduction

This chapter is concerned with modelling the enrichment of copper during oxidation of
copper-bearing steels. In the two previous chapters, the oxidation behaviour and
enrichment of residual elements of a set of model steels was reported. Results from
these investigations will be used as parameters in models and to test the validity of
predictions. A model that accurately predicts whether a copper-rich phase forms, and if
so, the amount that forms would be very useful. For example, it could be used to
investigate means of minimising copper-rich phase formation. Alternatively, it could act
as a warning system, highlighting alloy compositions and/or processing conditions
likely to cause problems relating to surface hot shortness.

6.2 Development of an Analytical Model

6.2.1 Initial Considerations

Consider an iron-copper alloy undergoing isothermal oxidation. As iron is selectively


oxidised, copper will enrich within the metal at the scale/metal interface. Since the
solubility of copper in FeO is negligible, copper will continue to accumulate at the
interface. In terms of a diffusion model, this situation can be described as a moving
interface problem. The scale/metal interface moves into the metal, and the copper
concentration at the interface (and within the subsurface region of the metal) varies with
time. To simplify the problem, one would assume that the system is semi-infinite and
consider one-dimensional diffusion occurring across a planar interface (see Figure 6-1).
Under these conditions, the copper concentration C at position x is described by Fick’s
second law:

183
Modelling of Residual Element Enrichment

∂C i ~ ∂ 2 C i
= Di (6.1)
∂t ∂x 2

~
where Ci and Di are the copper concentration and interdiffusion coefficient in phase i,

respectively; x the distance from the original scale/metal interface and t the time. Eqn.
~
(6.1) is valid only when Di is not a function of concentration. For an ideal
~
substitutional binary alloy A-B, D is defined as:

~
D = X B DA + X A DB (6.2)

where Di is the intrinsic diffusion coefficient and Xi the molar fraction for species i,
respectively.

C
Scale Metal

CMetal/Scale

Co

x
0 l

Figure 6-1: Schematic representation of copper concentration profile at the metal


surface prior to copper-rich phase formation. C represents concentration and x the
distance from the original metal surface. Other parameters are described in the
text. In all cases, ‘Metal’ refers to the iron/copper alloy.

In order to solve Fick’s second law, one initial condition and two boundary conditions
are required. In this case:

184
Modelling of Residual Element Enrichment

C = Co, x ≥ 0, t = 0
C = CMetal/Scale, x = l, t > 0 (where CMetal/Scale = f{t})
C = Co, x = ∞, t > 0

There appears to be no analytical solution to Eqn. (6.1) for the initial and boundary
conditions listed above, i.e. for a moving interface problem where the interfacial
concentration varies with time. Hence, no analytical model can be developed to describe
copper enrichment prior to copper-rich phase formation. However, solutions are readily
available where the interfacial concentrations are fixed. In the current problem, fixed
interfacial concentrations are encountered when a copper-rich phase is present, i.e. at
the scale/copper-rich phase and copper-rich/metal interface. Therefore, an analytical
model could be developed if it is assumed that a copper-rich phase forms immediately
upon oxidation. Such an assumption may be valid for the case of oxidation in air. For
example, significant quantities of copper-rich phase were observed for most model
steels after 10 minutes oxidation at 1150°C. In contrast, the formation of a copper-rich
phase was delayed for extended periods of time when oxidation took place in CO2-N2.
Hence, an analytical model assuming immediate copper-rich phase formation may be
applicable to the results obtained from oxidation in air, but not to those from oxidation
in CO2-N2.

6.2.2 Analytical Model Assuming Immediate Copper-rich Phase


Formation

Consider an iron-copper alloy undergoing isothermal oxidation. It is assumed that a


copper-rich phase forms immediately upon oxidation and that the copper solubility in
the scale is negligible. A planar, semi-infinite system is considered where diffusion
occurs only in one dimension. A schematic showing the copper concentration profile
after a period of oxidation is shown in Figure 6-2.

185
Modelling of Residual Element Enrichment

C
Scale Cu-rich phase Metal

CCu-rich/Scale
CCu-rich/Metal

JScale/Cu-rich JCu-rich/Metal
Metal/Cu-rich
C
/Metal/Cu-rich
JCu-rich/Scale J

Co

CScale/Cu-rich x
0 x1 x2

Figure 6-2: Schematic representation of copper concentration profile at the metal


surface when a copper-rich phase has formed.

The diffusion of copper in the copper-rich layer and metal is defined by Eqn. (6.1).
Within the copper-rich layer (x1 ≤ x ≤ x2), the boundary conditions are:

CCu-rich = CCu-rich/Scale, x = x1, t > 0


CCu-rich = CCu-rich/Metal, x = x2, t > 0

The solution to the boundary conditions for x1 ≤ x ≤ x2 is:

C Cu − rich / Metal − C Cu − rich / Scale   x  


 - erf (λ1 )
CCu − rich = C Cu − rich / Scale + erf (6.3)
erf (λ 2 ) − erf (λ1 ) 
  2 DCu − rich t 
 

x1 x2
where λ1 = and λ 2 = . A derivation for Eqn. (6.3) has been
2 DCu− rich t 2 DCu− rich t

placed in Appendix C.1.

Within the metal (x ≥ x2), the initial and boundary conditions are:
186
Modelling of Residual Element Enrichment

CMetal = Co, x ≥ 0, t = 0
CMetal = CMetal/Cu-rich, x = x2, t > 0

The solution to Eqn. (6.1) for x ≥ x2 is:

C o − C Metal / Cu − rich  x 
C Metal = C o − erfc  (6.4)
erfc(λ 2α ) 
 2 DMetal t

DCu −rich
where α = . A derivation for Eqn. (6.4) has been placed in Appendix C.2.
DMetal

With reference to Figure 6-2, a mass balance at the copper-rich phase/metal interface
(i.e. at x = x2) requires that the diffusion flux in the copper-rich phase to the interface
minus the flux in the metal away from the interface equals the rate at which copper is
added to the copper-rich phase:

∂x 2 Cu − rich / Metal
J Cu − rich / Metal − J Metal / Cu − rich =
∂t
C ( − C Metal / Cu − rich ) (6.5)

The flux terms in Eqn. (6.5) can be evaluated as follows:

∂C Cu − rich
J Cu −rich / Metal = − DCu − rich
∂x x = x2

( )
Cu − rich / Scale
(6.6)
DCu − rich C − C Cu − rich / Metal
=
πt [erf (λ 2 ) − erf (λ1 )]exp(λ 2 )2

∂C Metal
J Metal / Cu − rich = − DMetal
∂x x = x2

( )
Metal / Cu − rich
(6.7)
DMetal C − Co
=
πt erfc(λ 2α ) exp(λ 2α )2

187
Modelling of Residual Element Enrichment

∂   x  A  x2 
since  Aerf    = exp −  , where A is a constant.
∂x   2 Dt  πDt  4 Dt 

∂x 2 DCu− rich
Substituting Eqns. (6.6) and (6.7) into (6.5), and noting that = λ 2 , yields:
∂t t

A1 A2
− = A3 λ 2
[erf (λ 2 ) − erf (λ1 )]exp(λ 2 )
(6.8)
erfc(λ 2α ) exp(λ 2α )
2 2

(
where A1 = C Cu −rich / Scale − C Cu − rich / Metal ) DCu − rich
π

(
A2 = C Meta // Cu −rich − C o ) DMetal
π

(
A3 = C Cu − rich / Metal − C Metal / Cu −rich ) DCu −rich

Assuming that the consumption of metal is described according to a parabolic rate law:

(x1 )2 = k cp t (6.9)

where k cp is the parabolic rate constant in terms of metal consumption. As a result:

1 k cp
λ1 = (6.10)
2 DCu − rich

No analytical solution exists for Eqn. (6.8) and it must be solved graphically:

A1 A2
f (λ 2 ) = − − A3 λ 2
[erf (λ 2 ) − erf (λ1 )]exp(λ 2 )
(6.11)
erfc(λ 2α ) exp(λ 2α )
2 2

A plot of f(λ2) vs. λ2 can be used to graphically determine the solution for λ2 (see Figure
6-9).

188
Modelling of Residual Element Enrichment

6.3 Development of a Numerical Model

6.3.1 Initial Considerations

A numerical method was used to overcome the limitations imposed in modelling copper
enrichment using analytical methods. The purpose of this model was to determine
whether a copper-rich layer forms, and if so, the time taken and its thickness. Once
again, a simplified situation was analysed; i.e. copper enrichment in an iron-copper
alloy where one-dimensional movement of planar interfaces was considered. The
numerical model was divided into two distinct stages; the transient stage describing
enrichment of copper prior to the formation of a copper-rich layer, and the fixed
boundary condition stage describing copper enrichment when a copper-rich layer is
present.

Numerical modelling of diffusional processes involves numerical approximations of


Fick’s laws. The most common numerical technique to solve differential equations
associated with diffusional transport has been finite differences [147]. In general, the
technique involves the following steps:

• Establishing a grid across the region of material for which diffusion occurs. Each
grid point has a specific concentration associated with it. In many cases a uniform
grid spacing is used, although finer or coarser grid spacings may be used where the
concentration gradient is anticipated to be steep or shallow, respectively [147].
• Replacement of Fick’s laws with finite difference approximations. These are based
on Taylor series expansions [148]
• Definition of initial and boundary conditions.
• Calculation of a solution through an iterative process.

The first order, central-difference equation to Fick’s first law is given as [147]:

 C i j+1 − C i j−1 
J i = − D
j

 (6.12)
 2∆x 

189
Modelling of Residual Element Enrichment

where ∆x is the spacing between grid points and where i refers to the grid number and j
the time increment or iteration.

Finite difference approximations of Fick’s second law are given in an explicit or


implicit representation. An explicit representation requires all concentrations at time
step j to calculate those at j + 1, whereas an implicit representation requires both known
and unknown values. There are various advantages and disadvantages associated with
each representation [147,149]. The explicit method is simpler in terms of programming
but is slower compared to other numerical techniques [149]. This results from a
restriction on the size of the time step, which may lead to more iterations. For strongly
time-dependent boundary conditions, small time increments may be required. In this
situation, an explicit method may be superior as the use of small increments negates the
advantages of using an implicit method [147]. An explicit representation was chosen in
the current study in view of its simplicity and since time-dependent boundary conditions
would be encountered during the transient stage of the model.

The explicit representation of the terms in Fick’s second law is given as [149]:

j
∂C C i j +1 − C i j
= (6.13)
∂t i ∆t

and

j
∂ 2C C i j+1 − 2C i j + C i j−1
= (6.14)
∂x 2 i (∆x )2

For the case where D is not a function of C, substitution of Eqns. (6.14) and (6.13) into
(6.1) yields:

C i j +1 − C i j  C j − 2C i j + C i j−1 
= D i +1 

∆t  (∆x )2 

or

190
Modelling of Residual Element Enrichment

∆t
C i j +1 = C i j + D (
C i +j 1 − 2C i j + C i −j 1 ) (6.15)
(∆x ) 2

where the superscript j refers to the current iteration at time t, and j + 1 the next iteration
at t + ∆t. With the explicit representation, Ci j +1 can be calculated from the known

concentrations, C i j , at time i. An initial concentration profile is declared and Eqn.

(6.15) can be used to calculate new profiles, where the previously calculated are used as
a known profile for the current time step. The concentration grid at time t and the new
concentration at time t+∆t are shown schematically in Figure 6-3.

j
Ci-1j+1
C1 Cij+1
Ci+1j+1 Time
iteration
t+∆t
Ci-1j
Cij C j
i+1 Time
iteration
∆x t
0 1 2 i-1 i i+1 N-1 N N+1

Figure 6-3: Schematic of the concentration/space grid used in the finite difference
technique.

6.3.2 Stability Criterion for Convergence of Explicit Finite Difference


Method

Eqn. (6.15) can be rearranged to express the following:

191
Modelling of Residual Element Enrichment

 2 D∆t  ∆t
C i j +1 = C i j 1 − +D
2 
(
C i +j 1 + C i j−1 ) (6.16)
 (∆x )  (∆x ) 2

 2 D∆t 
The quantity 1 −  should not have a negative value as this may result in negative
2 
 (∆x ) 
values for C i j or lead to oscillations during the iterative process. As a result, a stability

criterion is obtained by declaring the quantity to be larger than zero. The criterion can
be expressed so that ∆t is limited by ∆x and D:

∆t ≤
(∆x )2 (6.17)
2D

Romig et al. [149] commented that it is good practice to keep ∆t at about 1 of the
10
limit set by Eqn. (6.17). The following stability criterion was used to determine ∆t in the
current model:

∆t = 0.04 ⋅
(∆x )2 (6.18)
D

6.3.3 The Fixed Grid Method

The enrichment of copper during oxidation is a moving interface problem. Moving


interfaces are treated using either a fixed or movable grid method. In the latter, the
locations of grid points are shifted according to the moving interface. A systematic
correction of the concentration value of each grid point is required at the end of each
iteration. Lee and Oh [150] found that the correction can result in an unrealistic
concentration profile, leading to serious mass balance errors. Lee and Oh [150] and Lee
[151] developed a fixed grid finite difference method that minimised the mass balance
error associated with a moving interface. Lee [151] also developed a method to treat the
formation of thin layers (less than two inter-grid distances). Such a treatment is
applicable to the early stages of copper-rich layer formation. As a result, a fixed grid
finite difference method was applied in the current model.

192
Modelling of Residual Element Enrichment

Figure 6-4 schematically illustrates a concentration profile about the interface existing
between two phases, A and B. The grid numbers are identified as m-1, m, m+1 and m+2
etc., with the interface lying between the mth and (m+1)th grid point. The position of the
interface between grid points is denoted using a fractional parameter p, where 0 ≤ p ≤ 1:

p=
(position of interface) − (position of m th grid ) (6.19)
∆x

C A B

C A/B

C B/A

p∆x

m-1 m m+1 m+2 m+3


x

Figure 6-4: Schematic concentration profile about the interface of two phases, A
and B.

The relevant equations required to evaluate Fick’s second law using explicit finite
difference relationships based on equidistant grids were summarised in Eqns. (6.12),
(6.13) and (6.14). However, in a fixed grid method where the interface moves through
the grid, the interface may not necessarily fall upon a grid point. Since the distance
between the interface and neighbouring grid points are not equal to ∆x, special
∂ 2C ∂C
expressions are required for and at and near the interface. These are based on a
∂x 2
∂x
Lagrangian interpolation. The generalised relationships for a three-point Lagrangian
interpolation are [152]:

193
Modelling of Residual Element Enrichment

1 ∂ 2C C i −1 Ci
= +
2 ∂x i ( x i −1 − x i )( xi −1 − x i +1 ) ( x i − x i −1 )(x i − xi +1 )
2

(6.20)
C i +1
+
(xi +1 − x i −1 )( xi +1 − x i )

and

∂C
= l i'−1 ( x)C i −1 + l i' ( x)C i + l i'+1 ( x)C i +1 (6.21)
∂x interface

where
(x − xi ) + (x − xi +1 ) ' (x − xi +1 ) + (x − xi −1 )
l i'−1 ( x) = , l i ( x) = ,
(xi −1 − xi )(xi −1 − xi +1 ) (xi − xi −1 )(xi − xi +1 )
(x − xi −1 ) + (x − xi )
l i'+1 ( x) =
(xi +1 − xi −1 )(xi +1 − xi )

and Ci and xi are the concentration and distance at grid i, respectively.

Referring to the notation used in Figure 6-4:

∂ 2C 2  C m −1 C m C A/ B 
=  − +  (6.22)
∂x 2 m (∆x )2 1 + p p p (1 + p ) 

∂ 2C 2  C B/ A C C 
=  − m +1 + m + 2  (6.23)
∂x 2 m +1 (∆x )2  (1 − p )(2 − p ) 1 − p 2 − p 

∂C A 1  pC m −1 (1 + p )C m (1 + 2 p )C A / B 
=  − +  (6.24)
∂x interface
∆x  1 + p p p (1 + p ) 

∂C B 1  (2 p − 3)C B / A (2 − p )C m +1 (1 − p )C m + 2 
=  + −  (6.25)
∂x interface
∆x  (1 − p )(2 − p ) 1− p 2− p 

194
Modelling of Residual Element Enrichment

where CA and CB is the concentration within phase A and B, respectively, and CA/B and
CB/A is the concentration in A at the interface and B at the interface, respectively.

Eqns. (6.22) to (6.25) develop singularities at p = 0 and 1. Zhou and North [153] and
Lee and Oh [150] adopted a procedure to avoid this by ignoring the affected grid point
when the interface was near it. The procedure used by Lee and Oh [150] is used in the
current model. A critical value, p*, is introduced where 0 < p* < 0.5. The mth grid point
is ignored when p < p* and the (m+1)th grid point ignored when p > 1- p*. When this
occurs, a linear concentration profile was assumed to exist between the (m-1)th or
(m+2)th grid point and the interface, respectively. When p < p*, the second derivative of
the concentration at the (m-1)th grid point and the concentration gradient at the interface
in A must be modified [150]:

∂ 2C 2  C m − 2 C m −1 C A/ B 
=  − +  (6.26)
∂x 2 m −1 (∆x )2  2 + p 1 + p (1 + p )(2 + p )

∂C A 1  (1 + p )C m − 2 (2 + p )C m −1 (3 + 2 p )C A / B 
= − +
 (1 + p )(2 + p ) 
(6.27)
∂x interface
∆x  2 + p 1+ p

When p > 1- p*, the following expressions for the second derivative of the
concentration at the (m+2)th grid point and the concentration gradient at the interface in
the B must be modified [150]:

∂ 2C 2  C B/ A C C 
= 2 
− m+ 2 + m+3  (6.28)
∂x 2 m+ 2 (∆x )  (2 − p )(3 − p ) 2 − p 3 − p 

∂C B 1  (2 p − 5)C B / A (3 − p )C m + 2 (2 − p )C m + 3 
=  + −  (6.29)
∂x interface
∆x  (2 − p )(3 − p ) 2− p 3− p 

195
Modelling of Residual Element Enrichment

6.3.4 The Initial and Boundary Conditions

The initial condition requires that all grid points be assigned a concentration value, i.e.
C i1 for i = 1 to N. This may be as simple as a uniform concentration throughout the

sample or one where a prior concentration profile exists prior to the diffusion analysis
[147]. For the current model, C i1 = Co where Co is the bulk concentration of copper

within the alloy.

The next step requires boundary conditions to be defined so that new boundary
conditions ( C1j +1 and C Nj +1 ) can be calculated. The first boundary condition defines the
concentration at the innermost grid point, N. In the current model a constant copper
concentration is set, being equal to the bulk concentration of copper within the alloy, Co
(i.e. CN = Co). This boundary condition is appropriate for samples that are infinitely
thick in comparison to the diffusion distance during the time of oxidation. It assumes
that none of the diffusing species reaches grid point N. This is equivalent to defining a
zero flux boundary and is described by using an imaginary grid point N + 1, such that:

C Nj +1 = C Nj −1 (6.30)

A description of the other boundary condition depends on whether a copper-rich layer is


absent or present, i.e. whether the model is in the ‘transient’ or ‘fixed boundary’
condition stage. A description of each case follows.

6.3.5 Transient Condition Stage

During the early stages of oxidation, the copper concentration at the scale/metal
interface and within the subsurface region of the metal is expected to vary with time. A
number of steps were involved in determining the concentration profile for each
iteration:

1. Determine the thickness of metal consumed during the particular iteration, ∆l.
2. Determine the amount of copper rejected due to metal consumption, Ai.

196
Modelling of Residual Element Enrichment

3. Set the concentration in the metal at the scale/metal interface, CMetal/Scale, and
calculate the profile within the metal.
4. Determine the amount of copper accumulated within the metal, Aii.

The concentration profile for a particular iteration was determined by varying CMetal/Scale
(and recalculating the concentration within the metal based on this concentration) until
the amount of copper rejected equalled that accumulated during the iteration. This is
represented schematically in Figure 6-5. During iteration j+1, the scale/metal interface
j+1
advances a distance of ∆l . The mass change due to motion of the interface and
diffusion in the metal can be calculated by determining the difference in the area
between the thick dashed line (concentration profile at j) and the thick solid line
(concentration profile at j+1). If all copper is rejected from the growing scale, the mass
change should be zero, i.e. Ai = Aii. To limit the computation time, a mass balance was
assumed when the difference between Ai and Aii was less than 0.05% of Ai.

C CMetal/Scale, j+1

Aii

C mj ++11

Ai C mj ++12

CMetal/Scale, j

∆l j+1 x

Figure 6-5: Schematic change in concentration profile after movement of the


scale/metal interface. The thick dashed line represents the concentration profile at
j while the thick solid line represents the concentration profile at j+1. If all copper
is rejected, Ai = Aii.

197
Modelling of Residual Element Enrichment

The relationship used to determine the concentration at the grid point adjacent to the
scale/metal interface depended on the value of p. For example, for p < p*, Eqn. (6.23)
replaced the usual finite difference expression, Eqn. (6.14):

j
∂C
j
∂ 2C
=D 2
∂t m +1 ∂x m +1

2 D  C Metal / Scale, j C m +1 C m + 2 
j +1
C m +1 −C j
m +1
j j
=  − + 
∆t (∆x )2  (1 − p )(2 − p ) 1 − p 2 − p 
2 D∆t  C Metal / Scale, j C m +1 C m + 2 
j j
C mj ++11 = 2 
− +  + C m +1
j
(6.31)
(∆x )  (1 − p )(2 − p ) 1 − p 2 − p 

where CMetal/Scale,j is the concentration of the previous concentration profile (i.e. of


iteration j) at the scale/metal interface after its displacement for the current iteration (i.e.
j + 1). In this example, the concentration at the (m+2)th, (m+3)rd etc. grid points were
determined using the usual finite difference relationship based on equidistant grid
spacing, Eqn. (6.15).

The method used to calculate Ai can be illustrated with reference to Figure 6-6. After
j+1
iteration j+1, ∆l of metal is consumed. The amount of copper rejected during j+1 is
thus equivalent to the sum of areas AI, AII and AIII. The general form of the amount of
copper rejected at the scale/metal interface during j+1 is thus:

l j +1
Ai = ∫ j C j ⋅ dx (6.32)
l

198
Modelling of Residual Element Enrichment

C
CMetal/Scale,j

Cmj +1

Cmj + 2
AI C mj + 3 C mj + 4
AII
AIII

lj l j+1
x

Figure 6-6: Schematic copper concentration profile at j and position of the


scale/metal interface at j + 1 due to consumption of ∆l j+1 of metal.

The integral in Eqn. (6.32) can be calculated by summing the areas of the trapezia that
are defined by the concentration profile at time step j and position of the scale/metal
interface at time step j + 1. For example, from Figure 6-6:

∆x j
AII =
2
(
C m +1 + C mj + 2 ) (6.33)

∆l was determined using the following method: assuming that the rate of metal
consumption obeys an initial linear law followed by a parabolic rate law, the position of
the scale/metal interface can be determined by the following relationships:

For t < to: l = k cl t (6.34)

For t > to: [ ]


l = k cp t − (t o − t gradient ) − k cp t gradient + l o (6.35)

where to is the transition time between linear and parabolic metal consumption, lo is the
thickness consumed at to, and k cl and k cp are the rate constants for linear and parabolic

199
Modelling of Residual Element Enrichment

metal consumption, respectively. tgradient defines the time at which the slope of the

relationship l 2 = k cp t equals the slope of the linear region, l = k cl t :

k cp
t gradient =
( )
4 k cl
2 (6.36)

As a result, Eqn. (6.35) is defined so that the slope of the parabolic portion of the curve
is equal to that of the linear curve at to (see Figure 6-7). This assures a smooth transition
from linear to parabolic metal consumption.

Linear region Parabolic region

l l = k cl t
l 2 = kcpt

C2
lo l 2 = k cp (t − C1 ) − C 2

tgradient to t

C1

Figure 6-7: Schematic of method used to determine the amount of metal consumed,
l, at time, t. C1 and C2 are constants, remaining parameters are defined in the text.

6.3.6 Fixed Boundary Condition Stage

Once the concentration at the interface exceeds the solubility limit for copper in the
metal, the numerical procedure was adapted to account for the formation of a copper-
rich layer. For the iteration in which the solubility limit was exceeded, the profile was
recalculated with the concentration at the interface fixed at the solubility limit. The
amount of copper defined by the difference between the total amount rejected due to

200
Modelling of Residual Element Enrichment

metal consumption and that accumulated within the metal was assumed to form as a
copper-rich layer. The mass balance criteria is illustrated in Figure 6-8. To limit the
computation time, a mass balance was assumed when the difference between A1 and A2
was less than 0.05% of A1.

C
Scale Metal
Copper-rich phase

A2

A1
Co

x
0

Figure 6-8: Schematic illustrating the mass balance criteria when a copper-rich
phase is present; the amount of copper rejected, A1, must equal that accumulated
ahead of the scale, A2.

A procedure was developed that varied the position of the copper-rich/metal interface
until the amount of copper accumulated ahead of the scale equalled that rejected. At
each attempt, the concentration profile within the copper-rich phase and metal were
recalculated. As described previously, grid points close to an interface (defined by p*)
were ignored. When the copper-rich phase contained no grid points, a linear profile was
assumed, defined by the concentration at the scale/copper-rich interface, CCu-rich/Scale,
and at the copper-rich/metal interface, CCu-rich/Metal.

This stage of the model also accounted for the possibility that the amount of copper
rejected may not be sufficient to sustain a copper-rich layer. For this situation the
copper-rich layer was assumed to dissolve and a similar procedure was adopted to that
outlined in the transient condition stage, i.e. the concentration at the scale/metal

201
Modelling of Residual Element Enrichment

interface was varied until a mass balance was achieved (where the interface
concentration is less than the solubility limit).

A complication arose when calculating the concentration profile within the copper-rich
phase. As one would expect, the diffusivity within the liquid copper-rich phase is much
higher than in the metal. The stability criterion of the numerical technique requires ∆t to
be extremely small when it is based on the diffusivity within the copper-rich phase. As
indicated in Table 6-1, these values are too small and led to unreasonable computational
times. Simulations conducted using the ∆t value of the metal indicated that numerical
instabilities developed when calculating the concentration within the copper-rich layer.
However, because diffusion in the liquid phase is so much faster than in the metal, it
need not be calculated for present purposes. Instead, the accumulation of the layer will
be predicted on the basis of mass balance using an average composition defined by:

C Cu −rich / Scale + C Cu − rich / Metal


C = (6.37)
2

where CCu-rich/Scale and CCu-rich/Metal are concentrations in the copper-rich layer determined
by a condition of local equilibrium at the scale/copper-rich and copper-rich/metal
interfaces, respectively. If there were no grid points between the interfaces, a linear
concentration profile was assumed. As a result, ∆t was based only upon conditions in
the metal regardless of whether a copper-rich layer was present.

Table 6-1: Variation in the time step when based on the diffusivity within the
copper-rich phase and metal (∆x = 0.1µm)

Temperature DCu-rich DMetal ∆t (s)


2 † 2 ☼
(°C) (cm /s) (cm /s) Based on DCu-rich Based on DMetal
-5 -12
1100 4.2×10 9.6×10 9.6×10-8 0.42
-5 -11 -8
1200 5.3×10 5.0×10 7.6×10 0.08
-5 -10 -8
1300 6.5×10 2.2×10 6.1×10 0.02

Assumed that the interdiffusion coefficient is approximated by the self-diffusivity of
liquid copper. Data taken from [154].

Calculated from Eqn. (6.2) for a 0.47wt%Cu-Fe alloy using DCu and DFe data from
[140].

202
Modelling of Residual Element Enrichment

6.4 Application of the Analytical Model

One of the major assumptions made in the development of the analytical model outlined
in Section 6.2.2 was that a copper-rich phase formed immediately upon oxidation. This
is obviously not applicable to the model steels oxidised in the CO2-N2 environment
since a copper-rich phase did not form for an extended period of time. However, a
copper-rich phase formed rapidly when oxidation took place in air. As a result,
parameters relevant to the oxidation in air at 1150°C were used to test the model.

It is more correct for the concentration parameters used in the equations in Section 6.2.2
be expressed in terms of mol. species/cm3 phase instead of atomic or mass fraction.
Concentration measured as an atomic fraction was converted to that in terms of mol.
species/cm3 using the following relationship:

Ci
Ni = (6.38)
Vm

where Ni and Ci are the concentration of species i expressed in mol. species/cm3 and
atomic fraction, respectively; and Vm is the molar volume of the phase under
consideration. The molar volume can be determined from the following equation:

j MW j
Vm = (6.39)
ρj

where MWj and ρj are the molar weight and density, respectively of phase j. The density
of the copper-rich phase was approximated closely by the density of liquid copper
[155]:

ρ Cu = 8.000 − 0.801 × 10 −3 (T − 1356 ) (6.40)

where density is expressed in g/cm3 and the temperature, T, in K. Taking the molecular
weight of copper as 63.546g/mol, VmCu-rich = 7.997cm3/mol at 1150°C. The molar

203
Modelling of Residual Element Enrichment

volume of the metal was approximated closely by the value for pure iron, i.e. VmMetal =
7.096cm3/mol.

Table 6-2 lists the parameters used in the model. The variation of f(λ2) vs. λ2 is shown
in Figure 6-9. No solution can be obtained as f(λ2) develops singularities for values of
λ2 less than approximately 0 and greater than 0.0042. A singularity develops for λ2 >
0.0042 since λ2α > 2.8 and for z ≥ 2.8, erfc(z) ≈ 0. Parameter values relevant to the
conditions of oxidation at 1050°C in air were also substituted in Eqn. (6.11). However
the result was relatively unchanged, i.e. a solution could not be obtained. It is therefore
apparent that a solution for Eqn. (6.11) cannot be obtained using parameter values
relevant to this investigation.

Table 6-2: Parameter values chosen for testing at 1150°C

Parameter Value Details


Co 4.13×10 -3
Equivalent to a 0.47wt%Cu-Fe alloy
Predicted using Thermo-Calc [126] for an iron-copper
CMetal/Cu-rich 7.12×10-2
alloy
Predicted using Thermo-Calc [126] for an iron-copper
CCu-rich/Metal 0.9563
alloy
CCu-rich/Scale 0.98 Estimated value for Fe-Cu-O system
DMetal 2.28×10-11cm2/s Estimated interdiffusion coefficient in metal †
Estimated interdiffusion coefficient in copper-rich
DCu-rich 4.71×10-5cm2/s
phase [154] ☼
Parabolic rate constant estimated from 0.47Cu steel
k cp 9.4×10-7cm2/s
oxidised at 1050°C in air ◊

Calculated using Eqn. (6.2), where DCu = 2.27×10-11cm2/s [140] and DFe = 3.98×10-
11
cm2/s [141].

Assumed that the interdiffusion coefficient is approximated by the self-diffusivity
of liquid copper.

No rate constants were calculated at 1150°C. This value was estimated from the
0.47Cu steel rate constant at 1050°C, assuming that the temperature dependence
can be described by an Arrhenius relationship (see Appendix C.3).

204
Modelling of Residual Element Enrichment

-1.42E-04
-1.44E-04
-1.46E-04
-1.48E-04
2)

-1.50E-04
f(

-1.52E-04
-1.54E-04
-1.56E-04
-1.58E-04
-0.001 0 0.001 0.002 0.003 0.004 0.005
λ 2

Figure 6-9: Variation of f(λ2) with λ2 using parameters outlined in Table 6-2.

6.5 Application of the Numerical Model

6.5.1 Initial Simulations

The numerical model was developed using Microsoft VisualBasic 6 to run on


Windows98. An Intel Celeron 433MHz with 128MHz RAM was used to perform all
simulations. The input concentration variables for the model were expressed in terms of
weight fraction and converted to mol. species/cm3 phase using Eqns. (6.38) and (6.39).
The density of the copper-rich layer was approximated by the value for pure copper.
When the phase was liquid, Eqn. (6.40) was used, while a value of 8.96g/cm3 was used
when solid. The interdiffusion coefficient of the metal was calculated using Eqn. (6.2).
The interdiffusion coefficient of the copper-rich layer was approximated closely by the
value for pure copper. The input scaling rate constants were expressed in terms of scale
thickness (k'). They were converted to values expressed in terms of the thickness of
metal consumed (kc) assuming that only an FeO scale formed (see Appendix A.2).

The first test of the validity of the numerical model was to simulate a relatively simple
condition; the enrichment of copper in the metal prior to the formation of a copper-rich
layer. Model parameters were chosen to simulate the behaviour of the 0.47Cu steel after
30 minutes oxidation at 1250°C. The scaling rate parameter determined from TGA (see

205
Modelling of Residual Element Enrichment

Table 4-2) could not be used as the predicted scale thickness differed from that
measured in this sample. Using the TGA-determined figure would have resulted in a
larger quantity of rejected copper. As a result, the scaling rate parameter used was
calculated from the average scale thickness of the sample. The average scale thickness
of the 0.47Cu steel after 30 minutes oxidation at 1250°C was 71.3µm. Since parabolic
kinetics were not established until approximately 60 minutes, a linear rate constant was
calculated from the average scale thickness (i.e. 3.96×10-6 cms-1).

A further point to consider is that this sample had a significant quantity of copper
occluded. The mass balance analysis was discussed in Section 5.3.3.2. The results for
this sample are summarised in Table 6-3. Since A1 > A2, a significant quantity of copper
was occluded. Assuming that the fraction of copper released by the oxidation process
that passes into the scale remains constant, the concentration of copper occluded into the
scale can be expressed as:

A1 − A2
N occluded = (6.41)
l

where Noccluded is the concentration of copper occluded into the scale and l is the
thickness of metal consumed. In this case, Noccluded = 3.9×10-4mol/cm3, which is
equivalent to Coccluded = 2.81×10-3.

Table 6-3: Mass balance results for 0.47Cu steel oxidised for 30 minutes at 1250°C
(where A1 and A2 are the amount of copper rejected due to metal consumption and
that accumulated within the metal, respectively)

A1 (mol Cu/cm2) A2 (mol Cu/cm2)


2.3×10-6 ± 1.7×10-7 7.5×10-7

In order for the numerical model to treat occlusion, the procedure used to determine the
amount of copper rejected, Eqn. (6.32), was modified to read:

Ai = ∫ j
l j +1

l
(C j
)
− C occluded ⋅ dx (6.42)

206
Modelling of Residual Element Enrichment

The parameter values chosen for simulation are summarised in Table 6-4. Figure 6-10
illustrates the predicted and measured concentration profiles. Some of the measured data
points near the scale/metal interface have been omitted due to their limited accuracy.
The accuracy is decreased since the electron probe has a sampling diameter of about 1-
2µm, and in the region of a phase boundary adjacent phases are sampled
simultaneously. Taking this into consideration, a more accurate prediction was obtained
when DMetal = 3.0×10-10cm2s-1. The magnitude of the increase of the diffusivity
compared to the literature-determined value is not unreasonable, and this value was
adopted in subsequent simulations at 1250°C.

Table 6-4: Parameter values chosen to simulate 0.47Cu steel oxidised for 30
minutes at 1250°C

Parameter Value Details


Co 4.13×10 -3
Copper concentration of original steel
Coccluded 2.81×10-3 Copper concentration occluded into the scale
DMetal 1.07×10-10cm2s-1 Interdiffusion coefficient of metal †
Linear rate constant estimated from 0.47Cu steel
kl ' 3.96×10-6cms-1
oxidised for 30 minutes at 1250°C
N 401 Total number of grid points
∆x 0.25µm Grid spacing
p* 0.2 Critical parameter for position of interface

Calculated using Eqn. (6.2), where DCu = 1.07×10-10 cm2s-1 [140] and DFe = 1.92×10-
10
cm2s-1 [141].

207
Modelling of Residual Element Enrichment

7
DMetal = 1.07×10-10cm2s-1
6 DMetal = 3.0×10-10cm2s-1
DMetal = 6.0×10-10cm2s-1
Concentration (at%)
5 × Measured

0
-1 0 1 2 3 4 5 6 7 8 9 10
Distance from scale/metal interface (µ m)

Figure 6-10: Predicted and measured copper concentration profile for 0.47Cu steel
oxidised for 30 minutes at 1250°C.

6.5.2 Simulations of Time Taken for Precipitation of the Copper-rich


Phase

6.5.2.1 Simulations at 1250°C

The next stage in testing the validity of the numerical model was to compare the
predicted and observed (see Table 5-1) time taken for the precipitation of copper-rich
phase in the model steels. The behaviour of the model steels at 1250°C is most relevant,
since copper-rich phase formation was delayed for an extended period of time. In the
case of the 0.47Cu steel, a copper-rich phase commenced formation between 30 and 60
minutes oxidation. To obtain a more accurate picture, further discontinuous oxidation
tests were performed in 8.6%CO2-N2 at 1250°C. The samples were cross-sectioned and
polished and the scale/metal interface was examined for the presence of a copper-rich
phase. Copper-rich phase formation was found to have started between 40 and 50
minutes oxidation. The parameter values used in the model simulation are summarised
in Table 6-5. The model predicted that a copper-rich layer commenced formation at
44.6 minutes. Thus, agreement between the predicted and observed result was achieved.

208
Modelling of Residual Element Enrichment

Table 6-5: Parameter values chosen to simulate commencement of copper-rich


phase formation in 0.47Cu steel oxidised at 1250°C

Parameter Value Details


Co 4.13×10 -3
Copper concentration of original steel
Coccluded 2.81×10-3 Copper concentration occluded into the scale †
Predicted solubility limit of copper in metal for
CMetal/Cu-rich 7.92×10-2
0.47Cu steel (see Table 5-7)
Copper-rich phase composition of 0.47Cu steel (see
CCu-rich/Metal 0.8922
Table 5-3)
CCu-rich/Scale 0.98 Estimated value for Fe-Cu-O system
DMetal 3.0×10-10cm2s-1 Interdiffusion coefficient of metal †
Estimated interdiffusion coefficient in copper-rich
DCu-rich 5.90×10-5cm2/s
phase at 1250°C [154]
Linear rate constant estimated from scale thickness
kl ' 4.5×10-6 cms-1 data of 0.47Cu steel oxidised for 30, 40 and 50
minutes at 1250°C
kp ' Measured parabolic rate constant for 0.47Cu steel (see
3.3×10-7cm2s-1
Table 4-2)
to 57.5 minutes Transition time between linear and parabolic oxidation
N 301 Total number of grid points
∆x 0.5µm Grid spacing
p* 0.2 Critical parameter for position of interface

Determined in Section 6.5.1.

In the case of the copper/nickel steels, further oxidation tests defined the time taken for
the commencement of copper-rich phase formation in the 0.46Cu-0.22Ni steel as
between 60 and 70 minutes. The parameter values used in the model simulation are
summarised in Table 6-6. The model predicted that a copper-rich layer commenced
formation at 47.8 minutes, significantly ahead of the observed result. However, the
measured scale thickness of the sample reacted for 60 minutes indicated that it oxidised
at a slightly lower rate than the other samples. Using a linear rate constant calculated
using the average scale thickness of the 60 minute-oxidised sample (i.e. k l ' = 4.34×10-
6
cms-1), the model predicted that a copper-rich phase would not form at all, clearly at
variance with actuality. Because other oxidation tests with the 0.46Cu-0.22Ni steel
yielded reaction rates more similar to the 70 minute-oxidised sample, it is concluded
that the 60 minute experiment produced spurious results. Since the 30 and 70 minute-
oxidised samples oxidised at a similar rate, and no copper-rich phase was observed in

209
Modelling of Residual Element Enrichment

the former, the model prediction of 47.8 minutes is in agreement with the observed time
interval of 30 to 70 minutes.

Table 6-6: Parameter values chosen to simulate commencement of copper-rich


phase formation in 0.46Cu-0.22Ni steel oxidised at 1250°C

Parameter Value Details


Co 4.04×10 -3
Copper concentration of original steel
Coccluded 3.17×10-3 Copper concentration occluded into the scale †
Predicted solubility limit of copper in metal for
CMetal/Cu-rich 0.1208
0.46Cu-0.22Ni steel (see Table 5-8)
Predicted copper-rich phase composition of 0.46Cu-
CCu-rich/Metal 0.7915
0.22Ni steel (see Table 5-8)
CCu-rich/Scale 0.98 Estimated value for Fe-Cu-O system
DMetal 3.0×10-10cm2s-1 Interdiffusion coefficient of metal ☼
Linear rate constant required to form measured scale
thickness of 0.46Cu-0.22Ni steel oxidised for 70
kl ' 6.72×10-6cms-1
minutes, assuming scale growth can be described by
Eqns. (6.34) and (6.35)
kp ' Measured parabolic rate constant for 0.46Cu-0.22Ni
2.4×10-7cm2s-1
steel (see Table 4-2)
to 59.5 minutes Transition time between linear and parabolic oxidation
N 801 Total number of grid points
∆x 0.5µm Grid spacing
p* 0.2 Critical parameter for position of interface

Determined using Eqn. (6.41) and values from Table 5-12.

Determined in Section 6.5.1.

It is interesting to note that the rate of copper enrichment due to metal consumption
must decrease once scale growth becomes controlled by diffusion through the scale (i.e.
parabolic oxidation). If a copper-rich phase has not commenced formation prior to the
transition between linear and parabolic oxidation, it should not form at all. Since most
model steels had a transition time of approximately 60 minutes (see Table 6-7), it is
therefore surprising that copper-rich phase formation was not observed for some steels
till after this time. However, it is possible that variations in the degree of occlusion
could delay the appearance of a copper-rich phase for oxidation times greater than the
transition time. Since the relationship between the amount of occlusion and oxidation
time was not investigated, this possibility cannot be verified. The simple assumption

210
Modelling of Residual Element Enrichment

made here that occlusion leads to a constant concentration of copper in the scale has
proved to be acceptable as a first approximation.

Table 6-7: Transition time (to) between linear and parabolic oxidation for model
steels at 1250°C

Steel to (minutes)
0.19Cu 41.5
0.47Cu 57.5
0.17Cu-0.20Ni 41.5
0.48Cu-0.13Ni 57.5
0.46Cu-0.22Ni 59.5
0.48Cu-0.52Si 67.5

In the case of the copper/silicon steel, a copper-rich phase was not observed for the 90
minute-oxidised sample but was for the 120 minute-oxidised sample. The parameter
values used in the model simulation are summarised in Table 6-8. Under these
conditions no copper-rich phase was predicted to form. The mass balance results
discussed in Section 5.3.3.2 indicated that occlusion was significant in this steel at
1250°C. As discussed above, the occlusion behaviour of this steel may explain the
discrepancy between the predicted and observed result.

211
Modelling of Residual Element Enrichment

Table 6-8: Parameter values chosen to simulate commencement of copper-rich


phase formation in 0.48Cu-0.52Si steel oxidised at 1250°C

Parameter Value Details


Co 4.22×10 -3
Copper concentration of original steel
Coccluded 3.17×10-3 Copper concentration occluded into the scale †
Predicted using Thermo-Calc [126] for an iron-copper
CMetal/Cu-rich 8.46×10-2
alloy ☼
Predicted using Thermo-Calc [126] for an iron-copper
CCu-rich/Metal 0.9278
alloy ☼
CCu-rich/Scale 0.98 Estimated value for Fe-Cu-O system
DMetal 3.0×10-10cm2s-1 Interdiffusion coefficient of metal ◊
Linear rate constant required to form measured scale
thickness of 0.48Cu-0.52Si steel oxidised for 120
kl ' 2.15×10-6cms-1
minutes, assuming scale growth can be described by
Eqns. (6.34) and (6.35)
kp ' Measured parabolic rate constant for 0.48Cu-0.52Si
2.0×10-7cm2s-1
steel (see Table 4-2)
to 67.5 minutes Transition time between linear and parabolic oxidation
N 301 Total number of grid points
∆x 0.5µm Grid spacing
p* 0.2 Critical parameter for position of interface

Determined using Eqn. (6.41) and values from Table 5-13.

The composition of the copper-rich phase and the solubility limit of copper in the
metal were not measured at 1250°C.

Determined in Section 6.5.1.

The 0.19Cu and 0.17Cu-0.20Ni steels were not observed to form a copper-rich phase at
1250°C. The parameters chosen to simulate the oxidation of the 0.19Cu steel at 1250°C
are summarised in Table 6-9. Under these conditions no copper-rich phase was
predicted to form. Appropriate parameter values for the 0.17Cu-0.20Ni steel also
indicated that a copper-rich phase would not form. As a result, these predictions are in
agreement with observation.

212
Modelling of Residual Element Enrichment

Table 6-9: Parameter values chosen to simulate oxidation of the 0.19Cu steel at
1250°C

Parameter Value Details


Co 1.67×10 -3
Copper concentration of original steel
Coccluded 0 Approximation; no relevant data analysed at 1250°C
Predicted solubility limit of copper in metal for
CMetal/Cu-rich 7.92×10-2
0.47Cu steel (see Table 5-7)
Copper-rich phase composition of 0.47Cu steel (see
CCu-rich/Metal 0.8922
Table 5-3)
CCu-rich/Scale 0.98 Estimated value for Fe-Cu-O system
DMetal 3.0×10-10cm2s-1 Interdiffusion coefficient of metal †
Linear rate constant estimated from scale thickness
kl ' 3.15×10-6 cms-1
data of 0.19Cu steel oxidised for 30 minutes
N 301 Total number of grid points
∆x 0.5µm Grid spacing
p* 0.2 Critical parameter for position of interface

Determined in Section 6.5.1.

6.5.2.2 Simulations at 1050 and 1150°C

Results from Chapter 5 indicated that the copper-rich phase formed rapidly when
oxidation took place at 1050 and 1150°C in air. Observation of samples after the
smallest oxidation time (10 minutes at 1150°C) revealed significant quantities of
copper-rich phase for most model steels. The predicted time for precipitation of the
copper-rich layer is summarised in Table 6-10, in which the parameter values used are
summarised in Table 6-11 and Table 6-12. The results indicate precipitation is almost
instantaneous at both 1050 and 1150°C. The lower oxidation rates observed in the
copper/nickel and copper/silicon steels had no significant effect on the precipitation
time. This is due to the rate of copper enrichment being much higher than that of back-
diffusion. At the lower oxidation rates the rate of copper enrichment was still much
more rapid than back-diffusion, leading to rapid precipitation times.

213
Modelling of Residual Element Enrichment

Table 6-10: Predicted time for precipitation of the copper-rich layer (seconds)

Steel 1050°C 1150°C


0.19Cu 0.3 0.044
0.47Cu 0.3 0.044
0.17Cu-0.20Ni 0.3 0.658
0.46Cu-0.22Ni 0.3 0.044
0.48Cu-0.52Si 0.3 0.044

Table 6-11: Parameter values chosen to simulate commencement of copper-rich


phase formation at 1050°C (N = 1501, ∆x = 0.05µm, p* = 0.3, DMetal = 3.35×10-
12
cm2s-1 † and CCu-rich/Scale = 0.98)

Steel Co Coccluded CMetal/Cu-rich CCu-rich/Metal k p ' (cm2s-1) ††


0.19Cu 1.67×10-3 0 5.41×10-2 ☼ 0.9222 ◊ 1.4×10-6
0.47Cu 4.13×10-3 0 5.41×10-2 ☼ 0.9222 ◊ 1.3×10-6
0.17Cu-0.20Ni 1.49×10-3 0 7.20×10-2 Ø 0.8056 ◊◊ 9.9×10-7
0.46Cu-0.22Ni 4.04×10-3 0 7.20×10-2 Ø 0.8056 ◊◊ 6.6×10-7
0.48Cu-0.52Si 4.22×10-3 0 5.50×10-2 †◊ 0.9625 †◊ 2.9×10-7

Calculated using Eqn. (6.2), where DCu = 3.35×10-12 cm2s-1 [145] and DFe = 2.80×10-
12
cm2s-1 [140].
††
See Table 4-2.

Predicted solubility limit of copper in metal for 0.47Cu steel (see Table 5-7).

Copper-rich phase composition of 0.47Cu steel (see Table 5-3).
Ø
Predicted solubility limit of copper in metal for 0.46Cu-0.22Ni steel (see Table 5-8).
◊◊
Copper-rich phase composition of 0.46Cu-0.22Ni steel (see Table 5-4).
†◊
Predicted using Thermo-Calc [126] for an iron-copper alloy.

214
Modelling of Residual Element Enrichment

Table 6-12: Parameter values chosen to simulate commencement of copper-rich


phase formation at 1150°C (N = 1501, ∆x = 0.05µm, p* = 0.3, DMetal = 2.28×10-
11
cm2s-1 † and CCu-rich/Scale = 0.98)

Steel Co Coccluded CMetal/Cu-rich CCu-rich/Metal k p ' (cm2s-1) ††


0.19Cu 1.67×10-3 0 6.75×10-2 ☼ 0.9133 ◊ 3.1×10-6
0.47Cu 4.13×10-3 0 6.75×10-2 ☼ 0.9133 ◊ 3.0×10-6
0.17Cu-0.20Ni 1.67×10-3 ◊† 1.49×10-3 * 0.1281 Ø 0.7656 ◊◊ 2.2×10-6
0.46Cu-0.22Ni 4.04×10-3 0 0.1281 Ø 0.7656 ◊◊ 1.5×10-6
0.48Cu-0.52Si 4.22×10-3 0 7.11×10-2 †◊ 0.9563 †◊ 6.4×10-7

Calculated using Eqn. (6.2), where DCu = 2.27×10-11cm2/s [140] and DFe = 3.98×10-
11
cm2/s [141].
††
No rate constants were calculated at 1150°C. These values were estimated from
measured values at 1050°C assuming that the temperature dependence can be
described by an Arrhenius relationship (see Appendix C.3).

Predicted solubility limit of copper in metal for 0.47Cu steel (see Table 5-7).

Copper-rich phase composition of 0.47Cu steel (see Table 5-3).
Ø
Predicted solubility limit of copper in metal for 0.46Cu-0.22Ni steel (see Table 5-8).
◊◊
Copper-rich phase composition of 0.46Cu-0.22Ni steel (see Table 5-4).
†◊
Predicted using Thermo-Calc [126] for an iron-copper alloy.
נ
Measured using EPMA.
* Determined using Eqn. (6.41) and values from Table 5-11.

Increasing the solubility limit of copper in the metal had little affect on the predicted
time. For example, when an arbitrary value of CMetal/Cu-rich = 0.3 was chosen, no change
in the predicted times occurred. Values as high as CMetal/Cu-rich = 0.9 did alter the
predicted times, but were still less than 1 second. This is a significant result as it
suggests that the addition of elements to alter the copper solubility limit in austenite will
have little effect on inhibiting the formation of a copper-rich phase under the conditions
studied at 1050 and 1150°C. However at 1250°C, model predictions suggested that
varying the copper solubility limit could have a significant effect on copper-rich phase
formation. For example, the model predicted that no copper-rich phase would form in
the 0.47Cu steel if the copper solubility limit in the metal had a value of CMetal/Cu-rich =
0.3 (all other parameters having the same value as summarised in Table 6-5). Recall
that when the practical value of 7.92×10-2 was used, a copper-rich layer commenced
formation at 44.6 minutes. Hence, changes in the solubility limit do have a significant
effect under the conditions studies at 1250°C, but not at 1050/1150°C. This is clearly a
kinetic effect due to differences in DMetal and k p ' . The activation energy for copper

215
Modelling of Residual Element Enrichment

diffusion in austenite (280.1kJ/mol [140]) is considerably higher than that for the
oxidation of low carbon steel in oxygen-rich atmospheres (126.8kJ/mol [61]). It follows
from the activation energies that the ratio k p ' D Metal decreases with increasing

temperature. At 1250°C, k p ' D Metal will be less than at 1050/1150°C. Because

oxidation in CO2-N2 is slower than in air, the actual ratio k p ' D Metal will be even lower.

In fact, it is so low that copper precipitation can be avoided under some circumstances.
This occurs when the rate of back-diffusion is high enough to remove most of the
copper accumulating at the metal surface so that the solubility limit is not exceeded.

The model was also used to predict the limiting copper concentration below which a
copper-rich phase would not precipitate during oxidation. Using model parameters
relevant to the 0.47Cu steel (see Table 6-11 and Table 6-12), the model predicted a limit
of 8.79×10-7 (0.0001wt%) at both 1050 and 1150°C. This value appears far too low
considering that copper concentrations ranging from 0.15 to 0.3wt% are commonly
quoted [77,80,91,101] as limits below which defects arising from surface hot shortness
will not occur. However, the simulation used to predict the limiting concentration
ignored occlusion. If occlusion were present, the limiting concentration would be
expected to rise. Mass balance measurements for the copper and copper/silicon steels at
1050 and 1150°C indicated that the amount of copper accumulated ahead of the scale
was comparable with that rejected due to metal consumption (see Section 5.3.3.2). It
was therefore concluded that occlusion was not significant under these conditions. Mass
balance analysis indicated that occlusion was significant in the copper/nickel steels at
these temperatures. For example, Coccluded = 1.49×10-3 (0.17wt%) for the 0.17Cu-0.22Ni
steel at 1150°C. It can be assumed that steels (of similar nickel concentration)
containing copper concentrations less than 0.17wt% would not form a copper-rich
phase. However, care needs to be taken in predicting limiting copper concentrations
based on the amount occluded. As will be discussed in Section 6.5.3, the level of
occlusion was found to vary with oxidation conditions. More research is required to
properly quantify the effect of oxidation time, temperature and alloying additions on the
amount of copper-rich phase removed into the scale by occlusion. This would allow
limiting copper concentrations to be determined more accurately.

216
Modelling of Residual Element Enrichment

The only steel in which a copper-rich phase was not observed at the scale/metal
interface when oxidation took place in air was the 0.17Cu-0.22Ni steel at 1150°C (see
Table 5-1). Using appropriate parameter values (see Table 6-13), the model predicted
that a copper-rich phase commenced formation rapidly (after 7.5 seconds). However,
after an extended oxidation period the predicted thickness was very small. For example,
after 30 minutes oxidation the model predicted a thickness of 0.07µm. Such a small
thickness would not have been detected using the methods used in the current study.
Obviously the large amount of occlusion measured in this sample contributed greatly to
the small copper enrichment predicted.

Table 6-13: Parameter values chosen to simulate oxidation of 0.17Cu-0.22Ni steel


at 1150°C

Parameter Value Details


Co 1.67×10-3 Copper concentration of original steel †
Coccluded 1.49×10-3 Copper concentration occluded into the scale ☼
Predicted solubility limit of copper in metal for
CMetal/Cu-rich 0.1281
0.46Cu-0.22Ni steel (see Table 5-8)
Copper-rich phase composition of 0.46Cu-0.22Ni steel
CCu-rich/Metal 0.7656
(see Table 5-4)
CCu-rich/Scale 0.98 Estimated value for Fe-Cu-O system
DMetal 2.28×10-11cm2s-1 Interdiffusion coefficient of metal ◊
Estimated interdiffusion coefficient in copper-rich
DCu-rich 4.71×10-5cm2/s
phase [154] ††
kp ' Parabolic rate constant estimated from 0.17Cu-0.22Ni
2.2×10-6cm2s-1
steel oxidised at 1050°C in air Ø
N 2001 Total number of grid points
∆x 0.2µm Grid spacing
p* 0.2 Critical parameter for position of interface

Measured using EPMA.

Determined using Eqn. (6.41) and values from Table 5-11.

Calculated using Eqn. (6.2), where DCu = 2.27×10-11cm2/s [140] and DFe = 3.98×10-
11
cm2/s [141].
††
Assumed that the interdiffusion coefficient is approximated by the self-diffusivity of
liquid copper.
Ø
No rate constants were calculated at 1150°C. This value was estimated from the
0.17Cu-0.22Ni steel rate constant at 1050°C, assuming that the temperature
dependence can be described by an Arrhenius relationship (see Appendix C.3).

217
Modelling of Residual Element Enrichment

6.5.3 Simulations of Predicted Copper-rich Layer Thickness

Before the results of this section are discussed, a comment should be made on the way
the results of the predicted copper-rich layer thickness are presented. As a consequence
of the numerical procedure employed, the predicted thickness varied with time. The
degree of variation could be minimised by choosing a finer grid spacing, and hence a
smaller time step. However, the need for minimising this variation was balanced against
the increased computation time. As a result, grid spacings were chosen that provided an
acceptable accuracy (a relative error which was generally no greater than 5% of the
predicted copper-rich layer thickness) and total computation time (no greater than 1
hour using an Intel Celeron 433MHz with 128MB RAM running Windows98).
Predictions of copper-rich layer thicknesses as a function of time had the typical
appearance of Figure 6-11. In subsequent discussion, these are represented by a line of
best fit.

0.66
Copper-rich layer thickness ( m)

0.65
0.64
0.63
0.62
Predicted
0.61
Line of best fit
0.6
0.59
0.58
0.57
0.56
100 102 104 106 108 110
Oxidation time (minutes)

Figure 6-11: Predicted copper-rich layer thicknesses (simulation using parameter


values from Table 6-5).

In the previous section, it was shown that predictions of the time taken for the
commencement of copper-rich phase formation usually agreed with observations. In this
section, predicted and measured copper-rich layer thicknesses as a function of oxidation
time were compared. The first example, the oxidation of the 0.47Cu steel at 1150°C, is a

218
Modelling of Residual Element Enrichment

simplified case. Firstly, the copper-rich phase formed quite often as a semi-continuous
layer and its distribution at the scale/metal interface was generally uniform. Secondly,
mass balance analysis indicated that the amount of copper-rich phase occluded was
negligible. The parameter values chosen and the results of the simulation are
summarised in Table 6-14 and Figure 6-12, respectively. The predicted and measured
copper-rich layer thickness compare very favourably. Although there appears to be
some discrepancy between the predicted and measured concentration profile, this can be
accounted for by the lack of accuracy in the measured profile when the microprobe
beam sampled adjacent phases.

Table 6-14: Parameter values chosen to simulate 0.47Cu steel oxidised for 60
minutes at 1150°C

Parameter Value Details


Co 4.13×10 -3
Copper concentration of original steel
Coccluded 0 Copper concentration occluded into the scale
Predicted solubility limit of copper in metal for
CMetal/Cu-rich 6.75×10-2
0.47Cu steel (see Table 5-7)
Copper-rich phase composition of 0.47Cu steel (see
CCu-rich/Metal 0.9133
Table 5-3)
CCu-rich/Scale 0.98 Estimated value for Fe-Cu-O system
DMetal 2.28×10-11cm2s-1 Interdiffusion coefficient of metal †
Estimated interdiffusion coefficient in copper-rich
DCu-rich 4.71×10-5cm2/s
phase [154] ☼
kp ' Parabolic rate constant estimated from 0.47Cu steel
3.0×10-6 cm2s-1
oxidised at 1050°C in air ◊
N 3001 Total number of grid points
∆x 0.2µm Grid spacing
p* 0.2 Critical parameter for position of interface

Calculated using Eqn. (6.2), where DCu = 2.27×10-11cm2/s [140] and DFe = 3.98×10-
11
cm2/s [141].

Assumed that the interdiffusion coefficient is approximated by the self-diffusivity of
liquid copper.

No rate constants were calculated at 1150°C. This value was estimated from the
0.47Cu steel rate constant at 1050°C, assuming that the temperature dependence can
be described by an Arrhenius relationship (see Appendix C.3).

219
Modelling of Residual Element Enrichment

(a)
3.5

Copper-rich layer thickness ( m)


3

2.5

2 Predicted
1.5 Measured

0.5

0
0 5 10 15 20 25 30
Oxidation time (minutes)

(b)
100
90
80
Concentration (at%)

70
60
Predicted
50
Measured
40
30
20
10
0
0 2 4 6
Distance from scale/metal interface (µ m)

Figure 6-12: Predicted and measured (a) copper-rich layer thickness and (b)
copper concentration profile after 15 minutes oxidation, for the 0.47Cu steel
oxidised at 1150°C.

The second example, the oxidation of the 0.47Cu steel at 1250°C, is slightly more
complicated as a copper-rich phase did not form for an extended period of time. In
addition, mass balance analysis indicated that occlusion was significant. The parameter
values chosen are summarised in Table 6-5, where ∆x was modified to 1µm. The
measured and predicted copper-rich layer thickness are summarised in Figure 6-13. Two
predicted thicknesses are shown, one using the Coccluded value of Table 6-5 and the other
where Coccluded = 0. Use of the parameter values of Table 6-5 accurately predicted the

220
Modelling of Residual Element Enrichment

time taken for commencement of copper-rich phase formation (see Section 6.5.2.1) and
the thickness of the 60 minute-oxidised sample. However, they underestimate the
copper-rich layer thickness for the 90 and 120 minute-oxidised samples. A more
accurate result for the longer oxidation times was obtained when no occlusion was used
in the simulation. However, under these conditions the copper-rich layer was predicted
to commence formation much earlier than was observed. These results indicate that the
level of occlusion decreases at longer oxidation times.

1.6
Copper-rich layer thickness ( m)

1.4
1.2
1 Measured
0.8 No occlusion

0.6 Occlusion

0.4
0.2

0
0 20 40 60 80 100 120
Oxidation time (minutes)

Figure 6-13: Predicted and measured copper-rich layer thickness of 0.47Cu steel at
1250°C.

The third example, the oxidation of the 0.46Cu-0.22Ni steel at 1250°C, was another
situation where the copper-rich phase formed as discrete particles and occlusion was
significant. The parameter values chosen are summarised in Table 6-6, with the
following modifications: N = 301 and ∆x = 1µm. The results of the simulation are
summarised in Figure 6-14a. It is difficult to determine whether the prediction is
sensible considering the large variation in the measured results. Only a very small
amount of copper-rich phase was detected at 90 minutes oxidation, whereas a
significant amount was measured at 120 minutes oxidation. As indicated in Figure
6-14b, the Coccluded value chosen has a significant effect on the predicted thickness. The
Coccluded value summarised in Table 6-6 was based on mass balance analysis from one
sample. Although this value was useful in predicting the time taken for the

221
Modelling of Residual Element Enrichment

commencement of copper-rich phase formation, it may not be relevant for simulations


that include the development of a copper-rich layer. As with the results of the 0.47Cu
steel oxidised at 1250°C, these results indicate that the level of occlusion decreases at
longer oxidation times. Prior to copper-rich phase formation, a solid region enriched in
copper is in contact with the scale. When a copper-rich layer forms, a liquid phase is
now in contact with the scale. As a result, the energetics at the interface would be
expected to change. Hence, it is not surprising that the level of occlusion also changes.

(a) 0.3
Copper-rich layer thickness ( m)

0.25

0.2
Predicted
0.15
Measured
0.1

0.05

0
0 20 40 60 80 100 120
Oxidation time (minutes)

(b)
1
Copper-rich layer thickness ( m)

0.9
0.8
0.7
0.6 Predicted (no
occlusion)
0.5
Measured
0.4
0.3
0.2
0.1
0
0 20 40 60 80 100 120
Oxidation time (minutes)

Figure 6-14: Predicted and measured copper-rich layer thickness for (a) occlusion
and (b) no occlusion of the 0.46Cu-0.22Ni steel at 1250°C.

222
Modelling of Residual Element Enrichment

6.5.4 Other Simulations

In Chapter 5 it was postulated that the effect of an increase in copper solubility limit in
the metal with nickel addition on copper-rich phase formation may only become
significant when the oxidation rate and copper diffusion within the metal are similar.
Some of the simulations already presented add evidence to this claim. From Section
6.5.2.2, altering the solubility limit made no difference in the time taken for the
commencement of copper-rich layer formation under the conditions investigated at 1050
and 1150°C. This was not the case under the conditions studied at 1250°C, as model
predictions indicated that altering the solubility limit could significantly affect copper-
rich phase formation. The reason for this alteration is the change in the relative
difference between the rate of copper enrichment and copper diffusion into the metal
(back-diffusion). The magnitude of this change is illustrated in Table 6-15.

Table 6-15: Predicted amount of copper dissolved in metal after 15 minutes


oxidation of 0.47Cu steel (parameter values summarised in Table 6-5 and Table
6-14, where Coccluded = 0)

Amount of copper dissolved in


Oxidation conditions metal relative to the total amount
accumulated ahead of the scale (%)
1150°C in air 0.5
1250°C in CO2-N2 93.9

At 1050 and 1150°C the difference is large, i.e. copper enrichment is much more rapid
than back-diffusion. Changes in the solubility limit do not significantly affect the
copper-rich layer formation since the capacity to remove the enriched copper by back-
diffusion is limited. However, at 1250°C, back-diffusion is more similar to the rate of
copper enrichment. As a result, a higher solubility limit would increase the time taken
for the commencement of copper-rich layer formation. If high enough, a copper-rich
layer may not form at all. The effect of temperature on the back-diffusion is discussed
later in this section (see Figure 6-16 and the preceding text).

As discussed in Section 2.3, the Bergman/West [108] and Bilby [71,132] models
showed the most promise in treating copper enrichment during oxidation. The only

223
Modelling of Residual Element Enrichment

other known model treating copper enrichment during oxidation, the Lewis model [94],
completely ignored diffusion. Bergman and West [108] compared predictions between
the two models (see Figure 2-51) and summarised the parameter values used in the
comparison (see Table 2-10). Using these values, the current model produced results
nearly identical to that of the Bergman/West model (see Figure 6-15). The reason why
the thickness of the Bilby prediction decreases at higher temperatures was attributed to
the effect of back-diffusion [132]. In their use of the model, Hewitt and Meadows [132]
stated that provided the scaling rate is not too high, the rate of removal by back-
diffusion will exceed the rate of enrichment. It is for this reason that the copper-rich
layer thickness first decreases and then does not form at higher temperatures.

1.8
Copper-rich layer thickness ( m)

1.6
1.4
1.2
Current
1
Bergman/West
0.8
Bilby
0.6
0.4
0.2
0
600 700 800 900 1000 1100 1200
Oxidation temperature (°C)

Figure 6-15: Comparison between various models of the copper-rich layer


thickness after 1 hour oxidation.

Table 6-16 compares the magnitude of the diffusivity in the metal and the scaling rate
constant used by Bergman and West [108]. This indicates that the diffusivity in the
metal is orders of magnitude smaller than the scaling rate. It is therefore not surprising
that the current model predicts an increasing copper-rich layer thickness with
temperature. It is not clear from Bergman and West [108] whether the same parameter
values were used in the Bilby model prediction. The Bilby model behaviour at higher
temperatures could be a result of the use of a lower scaling rate, higher diffusivity, or a
combination of both. For example, the Bilby model predicted that a copper-rich layer

224
Modelling of Residual Element Enrichment

does not form at 1200°C after 1 hour oxidation. Using parameter values suggested by
Bergman and West [108], the current model would require a DMetal value of 5×10-8cm2/s
to arrive at a similar result. This is considerably higher than the literature value, DMetal =
5.1×10-11cm2/s (calculated using Eqn. (6.2), where DCu = 5.06×10-11 cm2s-1 [140] and
DFe = 8.98×10-11cm2s-1 [141]). It is therefore likely that the Bilby model prediction did
not use the same parameter values summarised by Bergman and West [108].

Table 6-16: Comparison between copper diffusivity in metal and scaling rate
constants using data from [108]

Temperature (°C) DMetal (cm2/s) k cp (cm2/s) k p ' (cm2/s) †


900 1.5×10-13 8.4×10-8 2.6×10-7
1000 1.4×10-12 3.3×10-7 1.0×10-6
1100 9.6×10-12 1.1×10-6 3.3×10-6
1200 5.1×10-11 2.9×10-6 9.1×10-6

Conversion assuming that only an FeO scale forms (see Appendix A.2).

Back-diffusion is one mechanism that is commonly used to explain the reduction or


disappearance of copper-rich phase at the scale/metal interface at higher temperatures
[69,71,87,105]. The comparison between the current, Bilby and Bergman/West models
(see Figure 6-15) indicated that under those conditions, the current model predicts that
back-diffusion will have little effect at higher temperatures. This trend was investigated
further by performing simulations using scaling rate constants obtained in simulated
reheat furnace atmospheres [67]. Rate constants and other relevant parameter values are
summarised in Table 6-17 and Table 6-18. The results of the simulation for a 0.47Cu
steel oxidised for 1 hour and assuming that no copper is occluded is summarised in
Figure 6-16. As with the previous example, the predicted copper-rich layer thickness
increases with temperature. As indicated in Figure 6-16b, the relative amount of copper
accumulated within the metal increased with temperature. However, its contribution was
very small, i.e. less than 2% of the total amount accumulated ahead of the scale for all
temperatures investigated. Hence, these results appear to indicate that under the degree
of oxidation occurring in reheat furnaces, back-diffusion plays little part in reducing the
amount of enriched copper. However, in order to accurately predict copper enrichment
in a reheat furnace, one would need to know how the oxidation conditions (temperature

225
Modelling of Residual Element Enrichment

and gas composition) vary throughout the furnace. The mixing of combusted products
in a direct fired reheat furnace would be expected to vary and therefore the oxidation
potential of the furnace atmosphere would also vary. A decrease in the oxidation
potential would increase the relative contribution made by back-diffusion and therefore
aid in suppressing copper-rich phase formation.

Table 6-17: Reported ( k pp ) and converted ( k p ' ) parabolic rate constants from

A1006 oxidised in a simulated reheat atmosphere (112% theoretical combustion


air) [67]

Temperature (°C) k pp (g2/cm4s) k p ' (cm2/s) †


1000 3.1×10-7 1.9×10-7
1100 7.3×10-7 4.5×10-7
1200 1.5×10-6 9.5×10-7
1300 2.9×10-6 ☼ 1.8×10-6

Conversion assuming that only an FeO scale forms (see Appendix A.2).

Value estimated from Lee [67] using an Arrhenius relationship.

Table 6-18: Relevant parameter values for simulations using reheat oxidation
conditions

Temperature (°C) DMetal (cm2/s) † CMetal/Cu-rich (wt%) ☼ CCu-rich/Metal (wt%) ☼


1000 1.4×10-12 5.2 97.4
1100 9.6×10-12 7.3 97.0
1200 5.0×10-11 8.8 95.0
1300 2.2×10-10 10.2 91.6

Calculated from Eqn. (6.2) for a 0.47wt%Cu-Fe alloy using DCu and DFe data from
[140].

Calculated using Thermo-Calc [126] for an iron-copper alloy.

226
Modelling of Residual Element Enrichment

(a)
2.5

Copper-rich layer thickness ( m)


2

1.5

0.5

0
950 1000 1050 1100 1150 1200 1250 1300 1350
Temperature (°C)
(b)
Amount of copper dissolved in metal

2
accumulated ahead of scale (%)
relative to the total amount

0
1000 1100 1200 1300
Temperature (°C)

Figure 6-16: (a) Predicted thickness and (b) relative amount of copper
accumulated within metal from simulations using parameter values summarised in
Table 6-17 and Table 6-18 (0.47Cu-Fe oxidised for 1 hour, no occlusion).

As mentioned previously, back-diffusion is commonly quoted as contributing to the


observed reduction of the amount of copper-rich phase formed at higher temperatures.
In addition, ever since the work of Melford [87] it has generally been accepted that the
rate of back-diffusion would overtake that of enrichment at high temperatures [108].
However, the results of this investigation indicate that back-diffusion becomes
significant only under reduced scaling rate conditions. The only mechanism that could
contribute significantly under rapid scaling conditions at higher temperatures is that of

227
Modelling of Residual Element Enrichment

occlusion. More research is required to properly quantify the effect of oxidation time,
temperature and alloying additions on the amount of copper-rich phase removed into the
scale by occlusion. With a greater understanding of this phenomenon, the treatment of
occlusion in the current numerical model could be modified. This would provide a more
powerful tool in simulating the enrichment of copper during oxidation.

6.6 Summary and Conclusions

A numerical model was developed which successfully predicted enrichment of copper,


both prior to and after copper-rich phase formation. The conditions investigated at
1250°C presented a situation where the copper-rich phase did not form for an extended
period of time. Predictions of the time for commencement of copper-rich phase
formation were in close agreement with observation. The conditions investigated at
1050 and 1150°C differed markedly in the sense that a copper-rich phase was observed
to form rapidly. Under these conditions the model predicted that a copper-rich phase
formed almost instantaneously. The overriding factor in the two sets of behaviour was
the difference between the rate of copper enrichment and back-diffusion. At 1050 and
1150°C, the difference was found to be large. The high oxidation rate led to copper
being enriched much more rapidly than it could be removed by back-diffusion, causing
rapid copper-rich layer formation. In contrast, the lower oxidation rate at 1250°C (in
CO2-N2) allowed more copper to be accommodated within the metal and hence to have
a significant affect on copper-rich layer formation.

Agreement between the predicted and observed copper-rich layer thickness varied.
When the copper enrichment was simple; i.e. rapid formation, little or no occlusion, and
uniform distribution, agreement was excellent. This was not the case when occlusion
was significant and the copper-rich phase distribution was non-uniform. When this
occurred, the measured thickness often varied non-uniformly with time. It was proposed
that the amount of occlusion may vary with oxidation time and whether a copper-rich
phase was present or not. For example, results at 1250°C indicated that the level of
occlusion decreased at longer oxidation times. The model handled occlusion in a
simplified manner, assuming that the level was fixed and independent of time. Hence

228
Modelling of Residual Element Enrichment

discrepancies between the predicted and measured thickness could be accounted for by
the simplified way the model treats occlusion.

Model predictions indicated that under the conditions studied at 1050 and 1150°C,
increasing the copper solubility limit in austenite would not significantly affect copper-
rich phase formation. This is significant because it suggests that copper-rich phase
formation cannot be influenced by attempting to modify the solubility limit under rapid
oxidation conditions. This relates to the finding that the amount of enriched copper
accommodated within the metal via back-diffusion was very limited under these
conditions. The situation was quite different under the reduced scaling rate conditions
investigated at 1250°C. Predictions indicated that increasing the solubility limit could
increase the time taken for commencement of copper-rich phase formation to the point
that it does not form at all.

The model indicated that the commonly held belief that the rate of back-diffusion would
overtake that of enrichment at high temperatures could only occur under reduced scaling
rate conditions. When oxidation is rapid, e.g. in air or in reheat furnaces operated at
high air/fuel ratios, back-diffusion is not significant due to the rapid rate of copper
enrichment.

229
CHAPTER
SEVEN
Cracking Behaviour of Model Steels

7.1 Introduction

This chapter is concerned with characterising the cracking behaviour of the model steels
using a simulated hot working test. In previous chapters, an understanding of the
oxidation and residual element enrichment behaviour of the model steels was
developed. The aim of this chapter is to investigate the cracking behaviour arising from
oxidation conditions where the enrichment behaviour is understood.

Cylindrical specimens of the model steels (10mm diameter × 110mm) were hot
compressed after oxidation, at a high strain rate using a Gleeble 3500. The possible
number of tests was limited due to the small amount of material available for specimen
machining. Hence, only a selection of the model steels was available for testing.
Specimens of A1006 were used as a reference. Oxidation temperatures ranged between
1000 and 1250°C, while deformation temperatures ranged between 900 and 1100°C.
The specimens were cross-sectioned along the transverse plane of the mid-span bulge
zone (see Figure 3-5). Crack characteristics (amount of cracking, average and maximum
crack depth) were measured on the polished cross-sectioned samples using an optical
microscope and image analysis software. Details of testing conditions and quantitative
analysis methods are discussed in Section 3.4.

7.2 Results

7.2.1 General Points

Figure 7-1 illustrates a typical example of a specimen after hot compression testing. A
scale layer developed about the hot zone of the sample during the oxidation stage of the

230
Cracking Behaviour of Model Steels

test but most spalled during the compression stage. A thin scale layer remained after
compression. No cracks were visible on the surface of the specimens.

1cm
Figure 7-1: Typical appearance of a specimen after hot compression.

Some initial tests failed when specimens were melted during compression. The Gleeble
unit heats the specimen via electrical resistance and the temperature is monitored using
a thermocouple spot-welded to the centre of the specimen. As indicated previously,
most of the scale layer detached during the compression stage of the test. On some
occasions the thermocouple also became detached with the scale. As the Gleeble unit
attempted to maintain the required temperature, the rapid cooling of the detached scale
layer and its lack of response to the application of thermal power led to large amounts of
power being directed at the specimen. Ultimately, this led to the specimen melting. To
avoid this, thermal power was turned off prior to compression. In order to obtain the
required compression temperature, the hold time prior to compression was varied
according to the required temperature and the observed cooling rate. The logged
temperature from specimens in which the thermocouple maintained contact with the
specimen surface indicated a cooling rate of 33.1°C/s. For all but a few tests, excellent
agreement between the nominal and observed compression temperature was achieved
(see Figure 7-2).

The free span length of each specimen (i.e. the portion of the specimen able to deform)
was approximately 48mm. To determine the width of the hot zone, a preliminary test
was conducted where thermocouples were attached at 5mm and 10mm from the centre
of the specimen. The specimen was heated under vacuum (2.3×10-1 Torr) to avoid scale
formation and possible thermocouple detachment. The specimen was held at
temperatures ranging from 900 to 1250°C. The results indicated that the hot zone was
reasonably narrow, as the drop in temperature was approximately 42 and 106°C at 5 and
10mm, respectively, from the centre of the sample.

231
Cracking Behaviour of Model Steels

Compression
stage
1056

1054
Temperature (°C)

1052

1050

1048

1046
331.05 331.1 331.15 331.2
Time (s)

Figure 7-2: Temperature-time profile of specimen where the thermocouple


remained in contact with the sample surface during compression (nominal
compression temperature of 1050°C).

Little variation in the final maximum diameter in the mid-span bulge region was
measured for all alloys in all tests conducted. The strain developed in compressed
cylindrical samples is characterised by the hoop strain, ε:

Df
ε = ln (7.1)
Do

where Do is the initial diameter of the specimen and Df is the final maximum diameter of
the specimen in the mid-span bulge zone. ε varied between 0.20-0.30 for all samples.

7.2.2 Appearance of Cracks

Observation of the cross-sectioned specimens revealed that cracking was observed in


many of model steels tested. No cracks were observed in A1006 specimens under any
testing condition. As a result, the cracks observed in the model steels must have been
due to the presence of a copper-rich phase. For all but a few specimens, the distribution

232
Cracking Behaviour of Model Steels

of cracks about the circumference was found to be non-uniform and localised (see
Figure 7-3).

Resin

Metal 200µm

Figure 7-3: Optical image illustrating typical example of cracking developed about
circumference of mid-span bulge region (0.19Cu oxidised for 10 minutes at 1150°C
and compressed at 1050°C).

More detailed images of the form of cracking are presented in Figure 7-4. A thin scale
layer (not greater than 30µm in thickness) was observed about the specimen surface and
within cracks. As most of the scale layer spalled during compression, it is believed that
this adhered scale formed after spallation and/or during cooling.

In regions where cracking occurred, a copper-rich phase was often observed along crack
surfaces (see Figure 7-4b) and at the scale/metal interface (see Figure 7-4c). In regions
where cracking was not present, a copper-rich phase was rarely observed at the
scale/metal interface. In addition, a copper-rich phase was found to penetrate deep
within the metal about cracks. Although penetration of the metal was observed from
oxidation tests (see Section 5.2.1), the depth did not exceed approximately 50µm. The
maximum crack depth measured was approximately 400µm. Even at this depth, a
copper-rich phase was occasionally observed about the crack tip and penetrating into the
metal ahead of the crack.

233
Cracking Behaviour of Model Steels

(a)

(b)

(c)

Figure continued overleaf

234
Cracking Behaviour of Model Steels

(b)

(c)

Figure 7-4: BSE images of cracks and copper-rich phase penetration along (a)
sample surface, with higher magnification images along (b) crack tip and (c)
sample surface (0.47Cu oxidised for 5 minutes at 1250°C and compressed at
1050°C).

7.2.3 Measurement of Crack Characteristics

The hot compression testing mainly involved examining the behaviour of a selection of
model steels under standard compression conditions after oxidation at 1050, 1150 and

235
Cracking Behaviour of Model Steels

1250°C. Although the number of specimens was limited, there was an opportunity to
investigate two important questions: how does cracking vary with compression
temperature, and, how much oxidation is required for cracking?

Figure 7-5 summarises cracking characteristics as a function of compression


temperature. Note that the amount of cracking, Acrack, was determined by measuring the
surface area of cracking per unit scale/metal interface length (see Figure 3-8). The error
bars in this figure and those following represent the standard deviation in the Acrack
measurement. In classifying these results it was noted that Acrack < 1µm2/µm for
specimens with rare cracks, while Acrack > 10µm2/µm where cracks were observed on all
regions of the samples surface. Cracking was observed for temperatures as low as
900°C. The average crack depth varied little with compression temperature, however,
there was a maximum in Acrack at 1000°C. Apart from there being significantly more
cracking in this sample, microstructural analysis revealed no difference between it and
the others. As such, this observation remains unexplained.

236
Cracking Behaviour of Model Steels

(a) 180
160

Average crack depth ( m)


140
120
100
80
60
40
20
0
850 900 950 1000 1050 1100 1150

Compression temp. (°C)

(b) 16
14
12
A crack ( m 2/ m)

10
8
6
4
2
0
850 900 950 1000 1050 1100 1150
Compression temp. (°C)

Figure 7-5: Effect of compression temperature on (a) average crack depth and (b)
amount of cracking (0.47Cu steel oxidised at 1150°C for 10 minutes).

Figure 7-6 summarises cracking characteristics as a function of oxidation temperature.


Only limited tests could be performed due to a lack of specimens. Significant cracking
was observed for oxidation times as short as 1 minute. Interestingly, the amount of
cracking after 10 minutes oxidation was significantly lower than after 1 minute.
Microstructural analysis revealed no obvious differences between the samples.

237
Cracking Behaviour of Model Steels

(a) 160
140

Average crack depth ( m)


120
100
80
60
40
20
0
1 10
Oxidation time (minutes)

(b) 10
9
8
A crack ( m 2/ m)

7
6
5
4
3
2
1
0
1 10
Oxidation time (minutes)

Figure 7-6: Effect of oxidation time on (a) average crack depth and (b) amount of
cracking (0.47Cu oxidised at 1150°C and compressed at 1050°C).

The remaining hot cracking tests involved examination of the effect of steel
composition on cracking behaviour. The tests were conducted under standard oxidation
and compression conditions, the details of which were summarised in Section 3.4 (see
Table 3-3). Note that all compression was carried out at 1050°C. Table 7-1 summarises
under which conditions cracking occurred. All model steels were found to crack after
oxidation at 1050°C. At 1150°C, only the 0.17Cu-0.20Ni and 0.48Cu-0.52Si steels were
free of cracking. At 1250°C, only the 0.47Cu and 0.48Cu-0.52Si steels had cracks.

238
Cracking Behaviour of Model Steels

Table 7-1: Conditions under which cracking was observed in the model steels (yes -
3, no - 2)

Oxidation Temperature (°C)


Steel
1050 1150 1250
0.19Cu 3 3 2
0.47Cu 3 3 3
0.17Cu-0.20Ni 3 2 2
0.48Cu-0.13Ni 3 3 2
0.46Cu-0.22Ni 3 3 2
0.48Cu-0.52Si 3 2 3

Figure 7-7 summarises the cracking behaviour of the copper steels. Under conditions
where cracking occurred for both steels, no significant difference between the steels
could be measured within the rather large experimental scatter. However, the amount of
cracking in the 0.47Cu steel was significantly higher than in the 0.19Cu steel.
Microstructural observations revealed no indication of the cause in the reduction in the
amount of cracking in the 0.47Cu steel at 1150°C. The average crack depth of the
0.47Cu steel increased at 1250°C. Note that the maximum crack depth measured at
1250°C was 401µm, while at 1050 and 1150°C it was 194 and 207µm, respectively.

239
Cracking Behaviour of Model Steels

(a) 300

Average crack depth ( m)


250

200
0.19Cu
150
0.47Cu
100

50

0
1050 1150 1250
Oxidation temp. (°C)

(b) 8
7
6
A crack ( m 2/ m)

5
0.19Cu
4
0.47Cu
3
2
1
0
1050 1150 1250
Oxidation temp. (°C)

Figure 7-7: (a) Average crack depth and (b) amount of cracking of copper steels.

Figure 7-8 summarises the cracking behaviour of the copper/nickel steels. The 0.47Cu
steel was included to provide a reference. Under conditions where cracking was
observed (see Table 7-1), the amount of cracking was lower in the copper/nickel steels
than in the 0.47Cu steel at 1050°C, but similar at 1150°C.

240
Cracking Behaviour of Model Steels

(a) 300

Average crack depth ( m)


250

200 0.17Cu-0.20Ni
0.48Cu-0.13Ni
150
0.46Cu-0.22Ni
100 0.47Cu

50

0
1050 1150 1250
Oxidation temp. (°C)

(b) 8

7
6
A crack ( m 2/ m)

5 0.17Cu-0.20Ni
0.48Cu-0.13Ni
4
0.46Cu-0.22Ni
3 0.47Cu
2
1

0
1050 1150 1250
Oxidation temp. (°C)

Figure 7-8: (a) Average crack depth and (b) amount of cracking of copper/nickel
steels.

Figure 7-9 summarises the cracking behaviour of the copper/silicon steel. As before, the
0.47Cu steel was included to provide a reference. The addition of silicon was found to
reduce both the average crack depth and the amount of cracking at 1250°C. Although
not obvious in Figure 7-9b, a small value of Acrack was measured for the 0.48Cu-0.52Si
steel at 1250°C. No cracking was observed in the copper/silicon steel at 1150°C. At
1050°C, the average crack depth was similar, while the amount of cracking was higher
in the copper/silicon steel. Microstructural observations of the cracked samples tested at
1050°C revealed discrete copper-rich phase particles distributed along most of the metal
surface in the 0.48Cu-0.52Si steel, whereas this phase was sparsely distributed in the

241
Cracking Behaviour of Model Steels

0.47Cu steel. As a result, there was more copper-rich phase available for cracking in the
0.48Cu-0.52Si steel at this temperature.

(a) 300
Average crack depth ( m)

250

200
0.48Cu-0.52Si
150
0.47Cu
100

50

0
1050 1150 1250
Oxidation temp. (°C)
(b) 12

10
A crack ( m 2/ m)

8
0.48Cu-0.52Si
6
0.47Cu
4

0
1050 1150 1250
Oxidation temp. (°C)

Figure 7-9: (a) Average crack depth and (b) amount of cracking of copper-silicon
steels.

242
Cracking Behaviour of Model Steels

7.3 Discussion

7.3.1 Possible Limitations of the Hot Cracking Test

Before discussing the results from the hot cracking tests, some possible limitations of
the investigation are considered. It was evident from Chapter 4 that the scales formed in
air spalled readily during oxidation. In addition, detachment followed by rehealing
occurred. In the hot cracking tests, most of the scale layer detached during compression.
Since it did not remain intact, there was no way of knowing whether the scale had
remained completely adhered during the oxidation stage of the test. If detachment
during oxidation did occur, one would expect less copper-rich phase to be precipitated
and this would be expected to affect the cracking behaviour. In addition, this would
introduce an error when the results were interpreted in terms of the amounts of copper-
rich phase reported in Chapter 5. It should be noted, however, that the hot cracking
specimens would be less prone to scale detachment due to their larger size compared to
the oxidation samples investigated in Chapters 4 and 5.

Another possibility is that some copper-rich phase could have remained with the scale
layer as it detached during compression. If this were to occur, the amount of copper-rich
phase available for penetration would decrease. This phenomenon was not investigated,
but it should not be discounted as a possibility.

7.3.2 Mechanism of Crack Formation

The cracking observed in the hot cracking tests was clearly due to the presence of a
copper-rich phase. This conclusion follows from the results in Chapter 5 which showed
that A1006, which cannot form a copper-rich phase, never cracks under the test
conditions. Conversely, steels which formed a copper-rich phase invariably cracked. In
addition, microstructural observations of the hot cracking test specimens revealed an
obvious relationship between the presence of copper-rich phase and cracks: in general,
the presence of copper-rich phase at the metal surface was accompanied by nearby
cracking. That is to say, copper-rich phase at the metal surface was rarely observed
without the presence of cracks nearby. Also, copper-rich phase was often observed
along crack surfaces and penetrating within the metal adjacent to the cracks. Although
243
Cracking Behaviour of Model Steels

penetration of copper-rich phase was observed in samples from oxidation tests, the
penetration was far deeper and more common in the samples after hot compression.

Proposed mechanisms of crack formation were summarised in Section 2.2.2.3.2. The


formation of cracks from surface hot shortness is commonly referred to as a form of
liquid metal embrittlement (LME). There are many models describing LME [156], but
perhaps the most applicable to surface hot shortness is the grain boundary penetration
model. It is proposed that crack initiation is preceded by stress assisted diffusion of
embrittling atoms along grain boundaries [156]. Kajitani et al. [107] considered the
interfacial energy required for crack initiation in austenite with and without a copper-
rich phase present. Their results indicated that the interfacial energy is greatly reduced
in the presence of a copper-rich phase.

It is therefore apparent that cracks developed in the following manner: During


oxidation, the precipitation of copper-rich phase at the metal surface was followed by
some penetration of this phase along metal grain boundaries. However, in the presence
of stress, the amount and depth of penetration was greatly increased. Cracks initiated
because the presence of copper-rich phase along grain boundaries considerably lowered
the energy required for crack formation. The observation that in most cases the cracking
distribution was non-uniform is explained by recalling that the distribution of copper-
rich phase at the scale/metal interface after oxidation was found to be non-uniform.

In Section 2.2.2.3.2, two mechanisms describing crack development in surface hot


shortness were reviewed. The main difference between the two was whether an applied
stress was required for copper-rich phase penetration along metal grain boundaries. The
results from this study indicate that some penetration occurs during oxidation. However,
the amount and depth of penetration is small compared to that occurring during hot
working. Obviously, applied stress greatly enhances penetration. The relevant question
is whether the degree of penetration during oxidation contributes significantly to
cracking during hot working. The results are summarised in Figure 7-10. Only relevant
data was included in the analysis, i.e. results from 30 minute-oxidised steels at 1050°C
and 10 minute-oxidised steels at 1150°C (since oxidation in the hot cracking tests was
performed for 30 and 10 minutes at 1050 and 1150°C, respectively). Results at 1250°C
were not included since hot cracking tests were not carried out in the CO2-N2

244
Cracking Behaviour of Model Steels

atmosphere. Figure 7-10 indicates that the average copper-rich phase penetration depth
has little affect on the average crack depth. Hence, the degree of penetration occurring
during oxidation is not significant.

120
Average crack depth ( m)

100

80
1050°C
60
1150°C
40

20

0
0 5 10 15 20 25
Average Cu-rich phase penetration depth (µ m)

Figure 7-10: Variation in average crack depth with average copper-rich phase
penetration depth after oxidation.

7.3.3 Variation of Cracking Behaviour with Testing Conditions

The effect of varying the compression temperature on the cracking behaviour was
summarised in Figure 7-5. Significant cracking was observed for temperatures as low as
900°C. Analysis of the specimen temperature revealed that it reached 934°C just prior
to compression, after which the thermocouple was lost due to scale separation.
Microstructural investigation of the specimen revealed that the copper-rich phase was
liquid at the time of compression. The effect of tin concentration on the melting point of
the copper-rich phase was summarised in Table 5-6. The average tin content of the
copper-rich phase for all model steels was approximately 3at%. Using this value, a
melting point of 1042°C is predicted. Assuming that the tin content of the copper-rich
phase in the hot cracking specimens was equivalent, the copper-rich phase should have
solidified prior to compression. The results from Table 5-6 suggest that a copper-rich
phase containing 10at% tin would melt at 904°C. Other investigators [70,103] have

245
Cracking Behaviour of Model Steels

indicated that the copper-rich phase can remain molten to a temperature of 900°C or
lower.

The effect of varying the oxidation time on the cracking behaviour was summarised in
Figure 7-6. Significant cracking was observed for oxidation times as short as 1 minute.
In fact, the amount of cracking (Acrack) observed for this specimen was surprisingly high,
considering the lower value obtained for the 10 minute-oxidised test. No oxidation
experiments were performed for times less than 10 minutes at 1150°C. The numerical
model developed in Chapter 6 predicted a copper-rich phase thickness of 0.37µm after 1
minute oxidation, while the measured result after 10 minutes was 1.33µm. As will be
shown in the following section, an increase in the amount of copper-rich phase leads to
more cracking. One would expect the amount of cracking at 10 minutes oxidation to be
similar to or higher than the result at 1 minute. Hence the result for one of these
specimens must be anomalous. Some possible causes were discussed in Section 7.3.1.
More rigorous testing would be required to fully elucidate the relationship between
oxidation time and cracking behaviour.

7.3.4 Variation of Cracking Behaviour with Alloy Composition

Table 7-1 summarised which model steels formed cracks. The conditions under which a
copper-rich phase formed during oxidation were summarised in Table 5-1. Considering
the results at 1050 and 1150°C initially (since the oxidation and hot cracking
experiments at 1250°C were conducted in different atmospheres), most model steels
that were found to form a copper-rich phase in the oxidation experiments went on to
form cracks during the hot cracking tests. The only exception was the 0.48Cu-0.52Si
steel at 1150°C, which was found to form a copper-rich phase during oxidation
experiments but did not produce any cracks. For this steel, the amount of cracking at
1250°C was very small, while at 1050°C it was considerable. These results may be
attributed to the state of the subscale and its effectiveness in promoting occlusion.
Occluded particles were most evident when the subscale was liquid, i.e. at 1150 and
1250°C (see Section 5.2.2.4). Higher levels of occlusion, leading to less copper-rich
phase available for cracking, would be expected to reduce the amount of cracking. The

246
Cracking Behaviour of Model Steels

fact that no cracks were observed at 1150°C may indicate that occlusion was very
effective or that the result is an anomaly.

The hot cracking results at 1250°C cannot be directly compared to those from the
oxidation experiments since oxidation was carried out in a different atmosphere. To aid
in the interpretation of these results, a selection of the model steels (0.47Cu, 0.17Cu-
0.20Ni, 0.46Cu-0.22Ni and 0.48Cu-0.52Si) were oxidised in flowing air at 1250°C for 5
minutes. The same procedure as used in the discontinuous oxidation experiments was
adopted (see Section 3.2). Rare copper-rich phase particles were observed at the
scale/metal interface in the 0.46Cu-0.22Ni and 0.48Cu-0.52Si steels. Occluded particles
were common in both of the copper/nickel steels. Unfortunately, the scale of the 0.47Cu
steel had separated during reaction. Based on the observations of the other steels, a
copper-rich phase would have been expected to form at the scale/metal interface. In
view of these observations, one would have expected both the 0.48Cu-0.13Ni and
0.46Cu-0.22Ni steels to form cracks, albeit a small amount. Their absence may be a
result of the mechanisms discussed in Section 7.3.1.

Having established that the majority of steels that formed a copper-rich phase at the
scale/metal interface cracked during the hot cracking tests, the next logical point to
investigate is the relationship between the amount of copper-rich phase formed at the
scale/metal interface and the amount of cracking measured. Figure 7-11 summarises the
relationship between ACu and ACrack for all model steels. Results at 1250°C were not
included since hot cracking tests were not carried out in the CO2-N2 atmosphere. All
cracking tests were performed at the same compression temperature (i.e. 1050°C). Once
again, only relevant results were included in the analysis, i.e. ACu results from 30
minute-oxidised steels at 1050°C and ACu results from 10 minute-oxidised steels at
1150°C (since oxidation in the hot cracking tests was performed for 30 and 10 minutes
at 1050 and 1150°C, respectively). In general the amount of cracking was found to
increase with the amount of copper-rich phase measured at the scale/metal interface.
This result seems logical in view of the cracking mechanism discussed in Section 7.3.2.
An increase in the amount of copper-rich phase at the scale/metal interface would be
expected to result in the penetration of more grain boundaries and/or penetration to a
greater depth. This is discussed in more detail below.

247
Cracking Behaviour of Model Steels

A Crack ( m 2/ m)
3
1050°C
1150°C
2

0
0 0.5 1 1.5 2
2
A Cu (µ m / µ m)

Figure 7-11: Variation in the amount of cracking, ACrack, with the amount of
copper-rich phase measured at the scale/metal interface, ACu.

The relationship between the amount of copper-rich phase formed at the scale/metal
interface, the copper-rich phase penetration depth after oxidation and the cracking depth
is summarised in Figure 7-12. As before, all measurements were performed on 30
minute-oxidised steels at 1050°C and 10 minute-oxidised steels at 1150°C. Figure 7-12a
indicates that the average copper-rich phase penetration depth after oxidation does not
vary greatly with the amount formed. This result is expected if penetration along grain
boundaries is diffusion controlled. Under these conditions, there would be a limit to the
penetration depth, regardless of the amount of copper-rich phase present.

248
Cracking Behaviour of Model Steels

(a)
25

Average Cu-rich phase penetration


20

depth ( m) 15
1050°C
1150°C
10

0
0 0.5 1 1.5
2
A Cu (µ m /µ m)
(b)
120
Average crack depth ( m)

100

80
1050°C
60
1150°C
40

20

0
0 0.5 1 1.5
2
A Cu (µ m /µ m)

Figure 7-12: Variation in (a) the average copper-rich phase penetration depth and
(b) average crack depth with the amount of copper-rich phase formed at the
scale/metal interface.

In general, the average crack depth also does not vary greatly with the amount of
copper-rich phase formed (see Figure 7-12b). The strain developed in all compressed
samples remained relatively constant. If the penetration depth during compression is
controlled by the amount of compression, the penetration depth and average crack depth
would be expected to remain relatively constant. From Figure 7-10 it was shown that
the average copper-rich phase penetration depth during oxidation has little affect on the
average crack depth. In light of the cracking mechanism discussed previously, these

249
Cracking Behaviour of Model Steels

results indicate that an increase in the amount of copper-rich phase does not
significantly increase the depth of grain boundary penetration during oxidation or the
depth of cracking. Since the amount of cracking increases with the amount of copper-
rich phase formed, it can be concluded that increases in the latter result in more grain
boundary penetrations causing a larger number of cracks.

In the case of the copper steels, ACrack was higher in the 0.47Cu steel compared to the
0.19Cu steel, while the average crack depth did not vary significantly (see Figure 7-7).
Measurements of ACu for the copper steels (see Section 5.2.3) indicated that
significantly more copper-rich phase formed in the 0.47Cu steel compared to the
0.19Cu steel. Hence these results agree with the overall finding that ACrack increases
with ACu, while the average crack depth remains relatively unaffected.

Under conditions where cracking was found to occur, ACrack was lower in the
copper/nickel steels compared to the 0.47Cu steel at 1050°C while it was similar at
1150°C (see Figure 7-8). Recalling the results presented in Section 5.2.3, the effect of
nickel addition on ACu varied with temperature. For steels that were found to form a
copper-rich phase, ACu was higher at 1050°C for the copper/nickel steels compared to
the 0.47Cu steel, while it was similar at 1150°C. In view of the relationship between ACu
and ACrack (summarised in Figure 7-11), one would expect ACrack to be higher in the
copper/nickel steels at 1050°C. This assumes that copper enriched in a similar manner
in the oxidation samples and the hot cracking specimens. It is possible that the
copper/nickel steel specimens tested at 1050°C suffered from either of the anomalies
discussed in Section 7.3.1.

The effect of silicon addition on cracking behaviour was summarised in Figure 7-9. The
copper/silicon steel produced significant cracking at 1050°C, while none was observed
at 1150°C and only a very small amount formed at 1250°C. In fact, ACrack was slightly
higher in the copper/silicon steel than in the 0.47Cu steel at 1050°C. The ACu results
summarised in Section 5.2.3 indicated that the amount of copper-rich phase formed at
1050°C was similar, while at 1150°C it was greatly reduced by the addition of silicon.
One would therefore expect ACrack be more similar for the 0.48Cu-0.52Si and 0.47Cu
steels at 1050°C than at 1150°C, as was in fact observed. However, microstructural
analysis of the cracked specimens revealed more copper-rich phase present at the metal

250
Cracking Behaviour of Model Steels

surface in the 0.48Cu-0.52Si steel at 1050°C. It is therefore apparent that the increased
ACrack observed at this temperature was a result of there being more copper-rich phase
available for penetration and cracking.

7.4 Summary and Conclusions

A test was devised which enabled successful investigation of cracking caused by surface
hot shortness. Specimens of A1006, which did not form a copper-rich phase during
oxidation, did not suffer from cracking in any of the conditions investigated.
Microstructural observations of the hot cracking test specimens revealed an obvious
relationship between the presence of cracks and copper-rich phase. In general, copper-
rich phase at the metal surface was rarely observed without the presence of cracks
nearby. In addition, copper-rich phase was often observed along crack surfaces and
penetrating within the metal adjacent cracks. Although some penetration of copper-rich
phase along metal grain boundaries was observed from oxidation tests, the penetration
was far deeper and more common in the samples after hot compression. The depth of
penetration after oxidation was found not to significantly affect cracking depth. It was
proposed that cracks develop in the following manner: The formation of copper-rich
phase at the metal surface during oxidation is followed by some penetration of this
phase along metal grain boundaries. When stress is applied, the amount and depth of
penetration is greatly increased. Cracks are initiated since the presence of copper-rich
phase along grain boundaries considerably lowers the energy required for cracking.

For all but a few tests performed on the model steels, cracks were found to be
distributed non-uniformly and were localised in nature. This was explained by recalling
that the distribution of copper-rich phase at the scale/metal interface after oxidation was
found to be non-uniform. Since cracks only form in the presence of a copper-rich phase,
one would expect the cracking distribution to be related to the copper-rich phase
distribution at the metal surface.

Initial tests performed on the 0.47Cu steel revealed some interesting results. After
oxidation for 10 minutes at 1150°C, cracking was observed for compression
temperatures as low as 900°C. Significant quantities of tin in the copper-rich phase

251
Cracking Behaviour of Model Steels

would allow it to be molten at this temperature. When the oxidation time at 1150°C was
varied, significant cracking occurred for times as short as 1 minute. This result indicated
the high susceptibility of this steel to surface hot shortness under these conditions.

In almost all cases, steels which formed a copper-rich phase at the scale/metal interface
in oxidation experiments also cracked during hot compression tests. The amount of
cracking was found to increase with the amount of copper-rich phase measured at the
scale/metal interface. This was interpreted using the cracking mechanism described
earlier, i.e. an increase in the amount of copper-rich phase would be expected to result
in the penetration of more grain boundaries and/or penetration to a greater depth. Under
the conditions investigated, the average crack depth remained relatively constant when
the amount of copper-rich phase at the scale/metal interface varied. This is expected if
the copper-rich phase penetration depth of grain boundaries in the presence of stress is
controlled by the amount of stress. Hence, these findings indicate that an increase in the
amount of copper-rich phase at the interface results in the penetration of more grain
boundaries and the formation of more cracks.

Investigation of the cracking behaviour of the copper steels clearly indicated that an
increase in the copper concentration in the original steel results in more cracking. This
was attributed to the previous finding that the amount of cracking increased with the
amount of copper-rich phase formed at the scale/metal interface. The 0.19Cu steel was
far less susceptible to cracking: only small amounts of cracking were detected at 1050
and 1150°C, while none occurred at 1250°C.

Nickel addition was found to reduce the amount of cracking at 1050°C, but to have little
effect at 1150°C in the case of the 0.47Cu steel. In addition, no cracking was detected in
the 0.48Cu-0.13Ni and 0.46Cu-0.22Ni steels at 1250°C, or the 0.17Cu-0.20Ni steel at
1150 and 1250°C. Most of these observations were successfully interpreted in terms of
the amount of copper-rich phase measured from oxidation experiments. It can be
concluded that nickel addition reduced the amount of cracking at all temperatures, and
under some conditions, prevented cracking altogether, by suppressing or reducing
copper-rich phase formation.

252
Cracking Behaviour of Model Steels

The effect of adding silicon varied with temperature. At 1050°C, the amount of cracking
measured in the 0.48Cu-0.52Si steel was higher than in the 0.47Cu steel. No cracking
was measured in the copper/silicon steel at 1150°C while only a very small amount was
detected at 1250°C. Once again, these results could be explained in terms of the amount
of copper-rich phase formed during oxidation. At 1150 and 1250°C, the subscale
formed as a liquid and the amount of copper-rich phase was reduced in comparison to
the 0.47Cu steel. Hence, the hot cracking analysis indicated that silicon addition is
useful only when the subscale formed is liquid.

253
CHAPTER
EIGHT
Summary and Significance of Results

Defects arising from surface hot shortness are due to the formation of a copper-rich
phase during oxidation. There are a number of key factors which affect the likelihood
and extent of copper-rich phase formation. Kinetic (e.g. the oxidation and back-
diffusion rate) and thermodynamic (e.g. the copper solubility limit in austenite and
melting point of the copper-rich phase) factors greatly influence copper-rich phase
formation. Factors that influence the physical behaviour of the copper-rich phase are
also important; i.e. whether the copper-rich phase is present at the scale/metal interface
or occluded into the scale. This section will summarise the results of the investigation in
relation to these factors. In addition, the significance of results from the simulated hot
working test and diffusion modelling will also be discussed. Oxidation conditions and
alloying additions that reduce or eliminate cracking arising from surface hot shortness
will be identified.

Two oxidation conditions were used in this investigation: air at 1050/1150°C and CO2-
N2 at 1250°C. Both the oxidation and copper enrichment behaviour varied greatly
between these. Oxidation was rapid in air, and all steels oxidised according to a
parabolic growth rate law at 1050°C. Scale growth kinetics were not determined at
1150°C due to scale separation and healing occurring during oxidation. Oxidation was
comparatively slower in the CO2-N2 atmosphere; oxidation exhibited an initial linear
rate law which gradually transformed into a parabolic one. The copper steels oxidised at
a similar rate to iron. The copper/nickel steels and the copper/silicon steel oxidised more
slowly than iron, however the decrease was more pronounced in the latter. This was
attributed to the solid subscale formed at 1050°C, offering a barrier to diffusion and thus
lowering the oxidation rate. At the higher temperatures, large gaps were often observed
between the metal surface, subscale and scale layer. If the presence of a liquid subscale
promotes scale separation, the overall oxidation rate would decrease.

254
Summary and Significance of Results

Decreasing the oxidation rate of a copper-bearing steel would be of benefit in reducing


or eliminating the formation of a copper-rich layer. Additions of nickel and silicon were
found to reduce the oxidation rate, although the effect was stronger in the latter case.
However, these additions were also found to increase internal oxidation. As a result, the
scale/metal interfaces of the copper/nickel and copper/silicon steels were found to be
rougher than the copper steels. Rough scale/metal interfaces are known to increase the
adherence of scale [84,146], making descaling operations more difficult. As a
consequence, adherence problems have been observed in steels containing nickel
[84,146] and silicon [85,86] additions. Scale that is not removed may become rolled
into the product causing surface defects. Hence, the use of nickel and silicon to combat
the problem of surface hot shortness would need to be balanced against the effect these
additions may have on descalability.

Differences in the oxidation behaviour between the two oxidation conditions had a
significant effect on the copper enrichment behaviour. Significant quantities of copper-
rich phase were observed for most steels when oxidised in air at 1050 and 1150°C.
Simulations using the numerical model indicated that under these conditions, the rate of
copper enrichment was far higher than that of back-diffusion. As a consequence, the
copper solubility limit was exceeded almost instantaneously and the bulk of the
enriched copper accumulated within the copper-rich phase. Simulations indicated that
the reduced oxidation rates observed in the copper/nickel and copper/silicon steels could
not eliminate copper-rich phase formation, but did reduce the quantity formed. In
addition, increasing the copper solubility limit could not prevent copper-rich phase
formation. This is an important finding as it indicates that the strategy for inhibiting
copper-rich phase formation by increasing the solubility limit is not useful under
conditions of rapid oxidation.

Additions of nickel and silicon were found to be beneficial by reducing the amount of
copper-rich phase measured at the scale/metal interface. This occurred because nickel
and silicon addition promoted the occlusion of copper-rich phase into the scale.
Microstructural observations of oxidised samples revealed that occluded particles were
observed for all model steels at 1050 and 1150°C. However, they were more common at
1150°C, particularly in the copper/nickel and copper/silicon steels. In terms of reducing
the amount of copper-rich phase, nickel addition was beneficial at 1150°C in a steel

255
Summary and Significance of Results

containing a low concentration of copper (0.17wt%). No benefit was observed at


1050°C or at 1150°C (for the steels containing the higher concentration of copper).
Silicon addition proved useful at 1150°C when the subscale was liquefied, but not at
1050°C when it was solid.

The copper enrichment behaviour when oxidation took place in CO2-N2 was greatly
different to that occurring in air. The copper-rich phase was not observed for a
significant amount of time; and in some cases not at all. Simulations using the
numerical model indicated copper enrichment behaviour in CO2-N2 was much more
sensitive to changes in the original copper concentration, the oxidation rate and the
copper solubility limit. This was due to the rate of copper enrichment and back-
diffusion being more similar. For example, the model predicted the observed result that
the two model steels containing a low concentration of copper (0.19Cu and 0.17Cu-
0.20Ni) did not form a copper-rich phase. The copper concentration in these steels was
low enough that enrichment of copper did not exceed its solubility limit under these
conditions. A copper-rich phase was observed in the 0.48Cu-0.13Ni and 0.46Cu-0.22Ni
steels only after a considerably longer oxidation time than that observed in the 0.47Cu
steel. In addition, the amount of copper-rich phase measured in these steels was lower
than that in the 0.47Cu steel. These observations were due to a number of factors caused
by the presence of nickel: increased occlusion, increased copper solubility and the lower
oxidation rate measured in the copper/nickel steels compared to the copper steels. Less
copper-rich phase was measured in the 0.48Cu-0.52Si steel than the 0.47Cu steel.
Occlusion was found to be significant in the 0.48Cu-0.52Si steel. When this is
combined with the lower oxidation rate observed for this steel, less copper-rich phase is
expected.

The investigation indicated that the level of occlusion varied with steel composition and
oxidation conditions. The numerical model, which was based on diffusion and mass
balance analysis, could not predict occlusion. However, it was used to deduce how
occlusion varied by comparing predicted and measured copper-rich layer thickness.
Investigations at 1250°C revealed that the level of occlusion decreased at longer
oxidation times. It is not surprising that the level of occlusion changes with oxidation
time. Prior to copper-rich phase formation, a solid region enriched with copper is in

256
Summary and Significance of Results

contact with the scale. After formation, a liquid phase is in contact with the scale. The
energetics at the interface would be expected to change and therefore affect occlusion.

The simulated hot working tests revealed a number of important results. Cracks
developed in the following manner: the precipitation of copper-rich phase at the metal
surface is followed by some penetration of this phase along metal grain boundaries
during oxidation. The amount of penetration is small and was found not to significantly
influence cracking. During the compression stage of the test, secondary tensile forces
develop about the mid-span bulge zone of the specimen causing cracks. After
compression, the amount and depth of penetration greatly increased. The resulting
cracks develop since the presence of a copper-rich phase along grain boundaries
considerably lowers the energy required for crack formation. In most cases the crack
distribution about the sample’s circumference was non-uniform. This was a legacy of
the distribution of copper-rich phase during oxidation, which was generally found to be
non-uniform. The amount of cracking was found to increase with the amount of copper-
rich phase precipitated at the scale/metal interface during oxidation. Under the
conditions investigated, the average crack depth remained relatively constant when the
amount of copper-rich phase at the scale/metal interface varied. This is expected if
copper-rich phase penetration of grain boundaries is controlled by the amount of
compression. Hence, an increase in the amount of copper-rich phase at the interface
results in the penetration of more grain boundaries and the formation of more cracks.

It has been established that crack development in surface hot shortness requires the
copper-rich phase to be molten [69,101,102]. As a result, cooling hot shortness-affected
product in order to solidify the copper-rich phase prior to hot rolling may offer a
strategy in overcoming cracking. However, in the current study, cracks were found to
develop at relatively low compression temperatures. After 10 minutes oxidation at
1150°C in air, significant cracking was observed in the 0.47Cu steel when compressed
at 900°C. Microstructural observations revealed that the copper-rich phase was liquid
during compression. Having to cool affected product to below this temperature may not
be feasible in terms of roll wear and product properties.

The amounts of copper, nickel and silicon in the steels were found to affect cracking
behaviour. While significant amounts of cracking were measured in the 0.47Cu steel at

257
Summary and Significance of Results

all oxidation temperatures, only small amounts of cracking were detected at 1050 and
1150°C, while none occurred at 1250°C for the 0.19Cu steel. In general, nickel addition
reduced the amount of cracking at all temperatures; and under some conditions,
prevented cracking altogether by suppressing or reducing copper-rich phase formation.
Silicon reduced or completely suppressed cracking when the subscale formed was
liquid. The beneficial effects of nickel and silicon addition were related to their ability
to reduce the amount of copper-rich phase at the scale/metal interface and thus decrease
the amount of cracking.

In terms of strategies used to overcome or minimise surface hot shortness cracking,


nickel and silicon additions were found to be beneficial. However, beneficial effects
were found to be dependent upon oxidation conditions (oxidation potential and
temperature) and the original copper concentration of the steel. Under rapid oxidation
conditions, reduction in the oxidation rate (as observed with silicon and, to a lesser
extent, nickel addition) and increases in the copper solubility limit (nickel addition) did
not inhibit copper-rich phase formation. However, nickel (by promoting internal
oxidation) and silicon (due to the formation of a liquid subscale) addition promoted
occlusion and hence cracking was reduced or completely suppressed. The lower
oxidation potential of the CO2-N2 atmosphere illustrated the significant effect a lower
rate of copper enrichment can have. Under these conditions, a reduction of oxidation
rate and increase of the copper solubility limit were found to be beneficial. In addition,
occlusion was found to be significant in both the copper/nickel and copper/silicon
steels. More detailed research is required in quantifying the effect of oxidation time,
temperature and alloying (nickel and silicon) concentration on occlusion to properly
realise potential benefits from this mechanism.

This investigation has indicated that even small quantities of copper-rich phase at the
scale/metal interface can cause cracks. Preventing copper-rich phase formation is
obviously the key to overcoming the problem. To this extent, limiting the copper
enrichment rate through the use of a low oxidation potential atmosphere would be of
benefit. When the copper enrichment rate is rapid and copper-rich phase formation
cannot be avoided, occlusion is the only mechanism that can alleviate the problem.
Although cooling the steel prior to hot rolling so that the copper-rich phase is solid may
stop cracking, the low temperature required (especially if tin is present) is likely to

258
Summary and Significance of Results

render this strategy unfeasible due to deleterious effects on rolling equipment life and
product properties. Both nickel and silicon addition were found to promote occlusion,
however, increased occlusion is likely to adversely affect descalability by increasing
scale/metal interface roughness. In addition to the proposed research on occlusion, it
would also be of benefit to investigate the effect of occlusion on descalability. With a
better understanding of occlusion and how to promote it, and what levels can be
tolerated in terms of descalability, strategies could be developed to overcome surface
hot shortness when copper-rich phase formation is unavoidable.

259
REFERENCES
[1] N. Birks, G.H. Meier, Introduction to High Temperature Oxidation of Metals, Edward Arnold,
(1983).
[2] S.A. Bradford, “Fundamentals of Corrosion in Gases”, Metals Handbook, Ninth Edition, Volume
13, Corrosion, American Society for Metals, Metals Park, OH, 61-76 (1978).
[3] P. Kofstad, High-Temperature Oxidation of Metals, Wiley and Sons, New York, (1966).
[4] P. Kofstad, “Oxidation Mechanisms for Pure Metals in Single Oxidant Gases”, Proc. 6th
International Conference on High-Temperature Corrosion, Edited by R.A. Rapp, San Diego,
California, NACE, Houston, Texas, 123-138 (1981).
[5] D.L. Douglas, “Lattice and grain-boundary diffusion processes involved in the high-temperature
oxidation and sulfidation of metals and alloys”, Selected Topics in High Temperature Chemistry,
Defect Chemisty of Solids, Edited by O. Johannesen and A.G. Andersen, Elsevier, Amsterdam,
185-225 (1989).
[6] G. Yurek, “Mechanisms of Diffusion-Controlled High-Temperature Oxidation of Metals”,
Corrosion Mechanisms, Edited by F. Mansfield, Marcel Dekker, New York, USA, 397-447
(1987).
[7] C.Wagner, Z. Phys. Chem., B21, 25 (1933).
[8] L.Himmel, R.F. Mehl, C.E. Birchenall, “Self-Diffusion of Iron in Iron Oxides and The Wagners
Theory of Oxidation”, Trans. AIME, Journal of Metals, June, (1953).
[9] R.E. Carter, F.D. Richardson, Trans. AIME, 203, 336 (1955).
[10] W.J. Moore, B. Selikson, J. Chem. Phys., 19, 1539 (1951).
[11] P. Kofstad, R. Bredessen, “On the Use of The Wagner Model in Oxidation in Mixed Reactants”,
High Temperature Corrosion of Advanced Materials and Protective Coatings, Edited by Y.
Saito, B. Onay and T. Maruyama, Elsevier Science Publishers, 3-11 (1992).
[12] K. Sachs, C.W. Tuck, “Surface Oxidation of Steel in Industrial Furnaces”, Proc. Reheating for
Hot Working, Iron and Steel Institute, 1-17 (1968).
[13] A. Atkinson, “Transport processes during the growth of oxide films at elevated temperature”,
Reviews of Modern Physics, 57 [2], April, (1985).
[14] F. Ajersch, “Scale Formation in Steel Processing Operations”, Proc. 34th MWSP Conf., ISS-
AIME, 30, 419-437 (1993).
[15] D.J.M. Bevan, J.P. Shelton, J.S. Anderson, Chem. Soc., 1729 (1948).
[16] A. Bruckman, G. Simkovich, Corros. Sci., 12, 595 (1972).
[17] V. Raghavan, Phase Diagrams of Ternary Iron Alloys, Part 5, The Indian Institute of Metals,
Calcutta, (1989).
[18] J. Paidassi, “Oxidation of Iron in Air between 700 C and 1250 C”, Mem. Sci. Rev. Met., 54, 569-
585 (1957).
[19] W.W. Smeltzer, “The Oxidation of Iron in Carbon Dioxide-Carbon Monoxide Atmospheres”,
Transactions of the Metallurgical Society of AIME, 218, August, 674-681 (1960).
[20] F. Pettit, R. Yinger, J.B. Wagner Jr., “The Mechanism of Oxidation of Iron in Carbon Monoxide-
Carbon Dioxide Mixtures”, Acta Metallurgica, 8, 617-623 (1960).
[21] W.W. Smeltzer, “The kinetics of Wustite Scale Formation on Iron”, Acta Metallurgica, 8, June,
377-383 (1960).
[22] E.T. Turkdogan, W.M. McKewan, L. Zwell, “Rate of Oxidation of Iron to Wustite in Water-
Hydrogen Gas Mixtures”, The Journal of Physical Chemistry, 69 [1], 327-334 (1965).

260
References

[23] R. DeHoff, Thermodynamics in Materials Science, McGraw-Hill, (1993).


[24] J. Stringer, Corros. Sci., 10, 513-530 (1970).
[25] M.H. Davies, M.T. Simnad, C.E. Birchenall, “On the Mechanism and Kinetics of the Scaling of
Iron”, Journal of Metals, October, 889-896 (1951).
[26] H.J. Engell, F. Wever, Acta. Met., 5, 695-702 (1957).
[27] L.B. Pfeil, J. Iron Steel Inst., 119, 501 (1929).
[28] K. Hedden, G. Lehmann, Arch. Eisenh., 35, 839 (1964).
[29] A Rahmel, J. Tobolski, Corrosion Science, 5, 333 (1965).
[30] J.S. Sheasby, W.E. Boggs, E.T. Turkdogan, “Scale growth on steels at 1200 C: rationale of rate
and morphology”, Metal Science, 18, March, 127-136 (1984).
[31] D.P. Whittle, “Oxidation Mechanisms for Alloys in Single Oxidant Gases”, Proc. 6th
International Conference on High-Temperature Corrosion, Edited by R.A. Rapp, San Diego,
California, NACE, Houston, Texas, 171-183 (1981).
[32] B.D. Bastow, G.C. Wood, D.P Whittle, Oxidation of Metals, 16, 1 (1981).
[33] F. Gesmundo, Y. Niu, “The Criteria for the Transitions Between the Various Oxidation Modes
of Binary Solid-Solution Alloys Forming Immescible Oxides at High Oxidant Pressures”,
Oxidation of Metals, 50 [1/2], 1-26 (1998).
[34] F. Gesmundo, F. Viani, Y. Nui, Oxidation of Metals, 42, 409 (1994).
[35] C. Wagner, “Theoretical Analysis of the Diffusion Processes Determining the Oxidation Rate of
Alloys”, Journal of the Electrochemical Society, 99 [10], 369-380 (1952).
[36] C. Wagner, Z. Electrochem. Soc., 63, 772 (1959).
[37] R.A. Rapp, Corrosion, 21, 382 (1965).
[38] F. Gesmundo, F. Viani, Oxidation of Metals, 25, 269 (1986).
[39] R.A. Rapp, Acta. Metall., 9, 730 (1961).
[40] F. Gesmundo, P. Castello, F. Viani, “The Steady-State Corrosion Kinetics of Two-Phase Binary
Alloys Forming the Most-Stable Oxide”, Oxidation of Metals, 46, 383-398 (1996).
[41] C. Wagner, “Internal Oxidation of Cu-Pd and Cu-Pt Alloys”, Corrosion Science, 8, 889-893
(1968).
[42] D. Geneve, M. Confente, D. Rouxel, P. Pigeat, B. Weber, “Distribution Around the Oxide-
Substrate Interface of Alloying Elements in Low-Carbon Steel”, Oxidation of Metals, 51 [5/6],
(1999).
[43] A.U. Malik, D.P. Whittle, “Oxidation of Fe-C Alloys in the Temperature Range 600-850 C”,
Oxidation of Metals, 16 [5/6], December, 339-353 (1981).
[44] D. Caplan, G.I. Sproule, R.J. Hussey, M.J. Graham, Oxidation of Metals, 13, 255 (1979).
[45] D.J. McAdam, G.W. Geil, J. Res. Nat. Bur. Stand., 23, 63 (1939).
[46] C.A. Siebert, Trans. Amer. Soc. Met., 27, 75 (1939).
[47] E. Scheil, K. Kiwit, Arch. Eisenh., 9, 405 (1936).
[48] V. Lanteri, D. Huin, P. Drillet, D. Bouleau, P. Henry, H. Gaye, “Internal Oxidation of Fe-Si
Alloys in the Presence of External Scale”, Microscopy of Oxidation 3, Edited by S.B. Newcomb
and J.A. Little, Proc. Third International Conference on the Microscopy of Oxidation, 16-18
September, 535-550 (1996).
[49] M. Nagumo, Y. Hida, “Hot Shortness of Copper-Containing Steel in the Controlled Rolling
Process”, ATB Metallurgie, 23 [3], 10.1-10.19 (1983).
[50] M.I. Copeland, “Reducing Surface Hot Shortness of Copper-Containing Steels using Silicon
Additions and Controlled Reheating Conditions”, Bureau of Mines Report of Investigations,
[7936], 1-20 (1974).

261
References

[51] K. Sachs, “Variations in the Structure across the Thickness of the Scale on Nickel Steels”,
Journal of the Iron and Steel Institute, March, 348-357 (1957).
[52] W. Hatfield, J. Iron Steel Inst., 115, 483 (1927).
[53] P. Nanni, F. Gesmundo, “Scaling Behaviour of an Fe-4.54 wt% Cu Alloy in the Range 700 to
1000C Under 1 Atm Oxygen”, Corrosion, 36, 119 (1980).
[54] C.E. Birchenall, Oxidation of Metals and Alloys, Edited by D.L. Douglass, ASM, Metals Park,
Ohio, (1970).
[55] E. Brauns, A. Rahmel, Werst. Und Korr., 7, 448 (1956).
[56] C.S. Tedmon, Corros. Sci., 7, 525 (1967).
[57] Binary Alloy Phase Diagrams, Second Edition, Volume 2, Edited by T.B. Massalski, H.
Okamoto, P.R. Subramanian, L. Kacprzak, ASM International, (1990).
[58] C. Houper, M. Jallon, M. Confete, V. Lanteri, J.M. Jolivet, M. Guttmann, “Production of High
Quality Steels Using The Scrap/Electric Arc Furnace Route”, Proc. 79th Steelmaking
Conference, Pittsburgh Meeting, March 24-27, 79, 601-611 (1996).
[59] F.S. Pettit, J.B. Wagner (Jr.), “Transition from the Linear to the Parabolic Rate Law during the
Oxidation of Iron to Wustite in CO-CO2 Mixtures”, Acta Metallurgica, 12, January, 35-40
(1964).
[60] H.T. Abuluwefa, R.I.L. Guthrie, F. Ajersch, “Oxidation of Low Carbon Steel in Multicomponent
Gases: Part 1. Reaction Mechanisms during Isothermal Oxidation”, Metallurgical and Materials
Transactions A, 28A, August, 1633-1641 (1997).
[61] H. Abuluwefa, R.I.L. Guthrie, F. Ajersch, “The Effect of Oxygen Concentration on the
Oxidation of Low-Carbon Steel in the Temperature Range 1000 to 1250 C”, Oxidation of
Metals, 46 [5/6], 423-440 (1996).
[62] H.T. Abuluwefa, R.I.L. Guthrie, F. Ajersch, “Oxidation of Low Carbon Steel in Multicomponent
Gases: Part 11. Reaction Mechanisms during Reheating”, Metallurgical and Materials
Transactions A, 28A, August, 1643-1651 (1997).
[63] H. Abuluwefa, R.I.L. Guthrie, F. Mucciardi, “Scale Formation in a Walking-beam Steel Reheat
Furnace”, Proc. 34th MWSP Conf., ISS-AIME, 30, 453-468 (1993).
[64] E.A. Cook, K.E. Rasnussen, “Scale-free Heating of Slabs and Billets”, Iron and Steel Engineer
Yearbook, 175-182 (1970).
[65] K. Sachs, C.W. Tuck, “Scale Growth during Re-heating Cycles”, Werkstoffe und Korrosion, 21,
945-954 (1970).
[66] A.N. Minaev, V.M. Ol”Shanskii, M.M. Volkova, N.I. Shurova, “Scale formation in steels
exposed to gaseous fuel combustion products”, Steel in the USSR, 13, December, (1983).
[67] V.H.J Lee, “Scaling Behaviour of Steel in Simulated Reheat Furnaces Conditions”, Ph.D. Thesis,
University of New South Wales, May, (1997).
[68] H. Selenz, F. Oeters, “A contribution to the scaling of steel in technical flue gases”, Archiv Fur
Das Eisenhuttenwesen Steel Research, [5], 201-208 (1984).
[69] E.T. Stephenson, “Effect of Recycling on Residuals, Processing, and Properties of Carbon and
Low-Alloy Steels”, Metallurgical Transactions A, 14A [3], 343-353 (1983).
[70] D.A. Melford, “The Influence of Residual and Trace Elements on Hot Shortness and High
Temperature Embrittlement”, Phil. Trans. R. Soc. Lond., A 295, 89-103 (1980).
[71] A.J. Hartley, P. Eastburn, N. Leece, “Steelworks Control of Residuals”, Phil. Trans. R. Soc.
Lond., A 295, 45-55 (1980).
[72] ARC Funding Proposal prepared for this project by B. Gleeson, University of New South Wales,
(1996).
[73] M.I. Copeland, J.S. Howe, “Preventing Formation of Copper Alloys of Tin, Antimony, and
Arsenic on Steel Surfaces During Reheating to Reduce Hot Shortness”, Bureau of Mines Report
of Investigations, [8080], 1-22 (1975).

262
References

[74] T.Pham, “Current reheat furnace practices and the associated yield losses in BHP steel
operations”, The Broken Hill Proprietary Company Limited, January, (1994).
[75] H. Buchholtz, R. Pusch, Stahl und Eisen, 73, 204-212 (1953).
[76] K. Noro, M. Takeuchi, Y. Mizukami, “Necessity of Scrap Reclamation Technologies and
Present Conditions of Technical Development”, ISIJ International, 37 [3], 198-206 (1997).
[77] M. Wolf, H. Schwabe, “On Tramp Element Control in Electric Steelmaking”, Proc. 2nd
European Electric Steel Congress, Florence, Italy, 29 Sept.-1 Oct., 4.8/1 - 4.8/34 (1986).
[78] T.E. Dancy, P.R. Hastings, “Electric Furnace Steelmaking using Direct Reduced Iron for
Production of High Quality Flat Products, Rod and Bar”, Continuous Casting, Proc. Fourth
International Iron and Steel Congress, The Metals Society, London, 12-14 May, 15.1-15.14
(1982).
[79] B. Mintz, R. Abushosha, D.N. Crowther, “Influence of small additions of copper and nickel on
hot ductility of steels”, Materials Science and Technology, 11, May, 474-481 (1995).
[80] C. Marique, “Scrap Recylcing and Production of High Quality Steel Grades in Europe”, Proc.
79th Steelmaking Conference, Pittsburgh Meeting, March 24-27, 79, 613-619 (1996).
[81] T. Emi, O. Wijk, “Residuals in Steel Products - Impacts on Properties and Measures to Minimise
Them”, Proc. 79th Steelmaking Conference, Pittsburgh Meeting, March 24-27, 79, 551-565
(1996).
[82] D.T. Llewellyn, “Copper in steels”, Ironmaking and Steelmaking, 22 [1], 25-34.
[83] R. Rolls, “Scale adherence on steels”, Metal Treatment and Drop Forging, 28 [192], 346-348
(1961).
[84] T. Asai, T. Soshiroda, M. Miyahara, “Influence of Ni Impurity in Steel on the Removability of
Primary Scale in Hydraulic Descaling”, ISIJ International, 37 [3], 272-277 (1997).
[85] T. Fukagawa, H. Okada, Y. Maehara, “Mechanism of Red Scale Defect Formation in Si-added
Hot-rolled Steel Sheets”, ISIJ International, 34 [11], 906-911 (1994).
[86] Y. Ishii, A. Kodoi, I. Wakamatsu, “Improvements to the Surface Quality of High Silicon Content
Steel”, Proc. 34th MWSP Conf., ISS-AIME, 30, 447-450 (1993).
[87] D.A. Melford, “Surface Hot Shortness in Mild Steel”, Journal of The Iron and Steel Institute,
200 [4], 290-299 (1962).
[88] A. Nicholson, J.D. Murray, “Surface Hot Shortness in Low-carbon Steel”, Journal of The Iron
and Steel Institute, 203 [10], 1007-1018 (1965).
[89] M.I. Copeland, J.E. Kelley, “Reducing Surface Hot Shortness of Copper-bearing Steels”, Bureau
of Mines Report of Investigations, [7682], 1-19 (1972).
[90] K. Mayland, R.W. Welburn, A. Nicholson, “Influence of Microstructural and Compositional
Variables on Hot Working of Steel”, Metals Technology, 3 [8], 350-357 (1976).
[91] M.J.U.T. van Wijngaarden, G.P. Visagie, “The Effect of Residuals on the Presence of
Intergranular Surface Cracks on Continuously Cast Billets”, Proc. 79th Steelmaking Conference,
Pittsburgh Meeting, March 24-27, 79, 627-631 (1996).
[92] L. Habraken, J. Lecomte-Beckers, “Hot Shortness and Scaling of Copper-containing Steels”,
Copper in Iron and Steel, Edited by L. McDonald Schetky, John Waley & Sons, New York, 45-
81 (1982).
[93] M.M. Wolf, “Effects of Tramp Elements in Continuous Casting”, Ironmaking and Steelmaking,
12 [6], 284-301 (1985).
[94] P.J. Lewis, “Residual Enrichment Effects in Continuous Casting”, Ironmaking and Steelmaking,
[20], 126-133 (1993).
[95] M.M. Wolf, “Scale Formation and Descaling in Continuous Casting and Hot Rolling”, Iron and
Steel Maker, 27, January, 22 (2000).
[96] W.F. Savage, E.F. Nippes, R.P. Stanton, “Intergranular Attack of Steel by Molten Copper”,
Welding Journal, 57 [1], 9s-16s (1978).

263
References

[97] W.F. Savage, E.P. Nippes, M.C. Mushala, “Copper-Contamination Cracking in the Weld Heat-
Affected Zone”, Welding Journal, 57, May, 145s-152s (1978).
[98] W.F. Savage, E.P. Nippes, M.C. Mushala, “Liquid-Metal Embrittlement of the Heat-Affected
Zone by Copper Contamination”, Welding Journal, 57, August, 273s-245s (1978).
[99] G.F. Vander Voort, “Failures of Locomotive Axles”, Metals Handbook, Ninth Edition, Volume
11, Failure Analysis and Prevention, American Society for Metals, Metals Park, OH, 715-727
(1978).
[100] R.M. Fisher, A. Szirmae, “Embrittlement of Gun Steel by Copper”, The Metallurgical Society of
AIME, 253-263 (1984).
[101] Y. Kohsaka, C. Ouchi, “Hot Shortness of Copper Bearing High Strength Low Alloy Steels”, ATB
Metallurgie, 23 [3], 9.1-9.29 (1983).
[102] N. Imai, N. Komatsubara, K. Kunishige, “Effect of Cu and Ni on Hot Workability of Hot-rolled
Mild Steel”, ISIJ International, 37 [3], 224-231 (1997).
[103] S. Akamatsu, T. Senuma, Y. Takada, M. Hasebe, “Effect of nickel and tin additions on formation
of liquid phase in copper bearing steels during high temperature oxidation”, Materials Science
and Technology, 15, November, 1301-1307 (1999).
[104] C. Jonel, V. Leroy, T. Greday, L. Habraken, “Investigation of the Hot Shortness of Copper
Bearing Steel”, INCRA Report, (1962).
[105] D. Coutsouradis, V. Leroy, T. Greday, J. Lecomte-Beckers, “Review of Hot Shortness Problems
in Copper Containing Steels”, ATB Metallurgie, 23 [3], 7.1-7.24 (1983).
[106] G.L. Fisher, “The Effect of Nickel on the High-Temperature Oxidation Characteristics of
Copper-bearing Steels”, Journal of The Iron and Steel Institute, 207 [7], 1010-1016 (1969).
[107] T. Kajitani, M. Wakoh, N. Tokumitsu, S. Ogibayashi, S. Mizoguchi, “Influence of Heating
Temperature and Strain on Surface Crack in Carbon Steel Induced by Residual Copper”, Tetsu to
Hogane (Japanese), 81 [3], 29-34 (1995).
[108] A. Bergman, R. West, “Copper Enrichment During Reheating”, Scripta Metallurgica, 22 [5],
659-663 (1988).
[109] H. Suzuki, “Strain Rate Dependence of Cu Embrittlement in Steels”, ISIJ International, 37 [3],
250-254 (1997).
[110] A.R. Cox, J.M. Winn, “Scaling of plain and complex copper steels”, Journal of The Iron and
Steel Institute, 203, February, 175-179 (1965).
[111] N.S. Stoloff, “Liquid Metal Embrittlement”, Surfaces & Interfaces, II, Proc. 14th Sagamore
Army Materials Research Conference, Syracuse University Press, Syracuse, N.Y, 157-182
(1968).
[112] R.R. Hough, R. Rolls, “The High-Temperature Tensile Creep Behaviour of Notched, Pure Iron
Embrittled by Liquid Copper”, Scripta Metallurgica, 4, 17-24 (1970).
[113] R.R. Hough, R. Rolls, “Creep Fracture Phenomena in Iron Embrittled by Liquid Copper”,
Journal of Materials Science, 6 [12], 1493-1498 (1971).
[114] R.R. Hough, R.Rolls, “Copper Diffusion in Iron During High-Temperature Tensile Creep”,
Metallurgical Transactions, 2, 2471-2475 (1971).
[115] R.R. Hough, R. Rolls, “Some Factors Influencing the Effects of Liquid Copper on the Creep-
Rupture Properties of Iron”, Scripta Metallurgica, 8, 39-43 (1974).
[116] S. Seo, K. Asakura, K. Shibata, “Evaluation of Susceptibility to Surface Hot Shortness in Cu-
containing Steels by Tensile Test”, ISIJ International, 37 [3], 232-239 (1997).
[117] Metals Handbook, Volume 1, ASM, Metals Park, Ohio, (1961).
[118] C.S. Smith, “Grains, Phases and Interfaces: An Interpretation of Microstructure”, Transactions
AIME, [175], 15-48 (1948).
[119] L.H. Van Vlack, “Intergranular Energy of Iron and Some Iron Alloys”, Transactions AIME, 191,
March, 251-259 (1951).

264
References

[120] W.J.M. Salter, “Effects of Alloying Elements on Solubility and Surface Energy of Copper in
Mild Steel”, Journal of The Iron and Steel Institute, 204 [5], 478-488 (1966).
[121] Y. Zou, E.W. Langer, “A Study of the Formation and Penetration of the Molten Copper-rich
Phase in Iron with the Addition of Nickel and Tin”, Materials Science and Engineering, A110,
March, 203-208 (1989).
[122] K. Born, “Surface Defects in the Hot Working of Steel from Residual Copper and Tin”, Stahl
und Eisen, 73 [20], 1268-1277 (1953).
[123] N. Imai, N. Komatsubara, K. Kunishige, “Effect of Cu, Sn and Ni on Hot Workability of Hot-
rolled Mild Steel”, ISIJ International, 37 [3], 217-223 (1997).
[124] R.Y. Chen, R. Louey, D. Hobson, “Experimental Srudies on Two Fundamental Aspects of
Copper Hot Shortness”, BHP Technote, BHPR/GM/N/G/002, The Broken Hill Proprietary
Company Ltd., Melbourne Laboratories, Australia, January, (1997).
[125] D.A. Melford, Journal Iron and Steel Institute, 204, 495-496 (1966).
[126] B. Sundman, B. Jansson, J. Andersson, “The Thermo-Calc Databank System”, CALPHAD, 9 [2],
153-190 (1985).
[127] W.J.M. Salter, “Effect of Mutual Additions of Tin and Nickel on the Solubility and Surface
Energy of Copper in Mild Steel”, Journal of The Iron and Steel Institute, 207, December, 1619-
1623 (1969).
[128] T. Fukagawa, H. Fujikawa, “Effect of Small Amounts of Ni on Liquid-Cu Embrittlement in Hot-
Rolled Mild Steel After High-Temperature Oxidation”, Oxidation of Metals, 52, October, 177-
194 (1999).
[129] S. Seo, K. Asakura, K. Shibata, “Effects of 0.4% Si and 0.02% P Additions on Surface Hot
Shortness in 0.1% C-0.5% Mn Steels Containing 0.5% Cu”, ISIJ International, 37 [3], 240-249
(1973).
[130] B.J Keene, “Contact Angle and Work of Adhesion Between Nickel-Based Melts and Non-
Metallic Solids”, Natl. Phys. Lab. Rep., DMM(A)23, March, 120 (1991).
[131] S. Nakano, M. Ohtani, J. Jpn. Inst. Met., 34, 562 (1970).
[132] J. Hewitt, J. Meadows, “Effect of Reheating Conditions on Surface Quality of Steel Strip”, Iron
and Steel Inst. spec. Rep. no. 111, 55-60 (1968).
[133] Thermochemistry System version 5.1 IBM-PC Program Chemix, CSIRO Division of Mineral
Products, Australia, (1988).
[134] S.J.B. Reed, Electron Microprobe Analysis, Second Edition, Cambridge University Press,
Cambridge, 274-288 (1993).
[135] “Liquation Embrittlement Study Using Gleeble Systems”, Gleeble Systems Application Note,
(1997).
[136] R. Chen, personal communication, BHP Research Laboratories, Melbourne, (1999).
[137] R. Chen, personal communication, BHP Research Laboratories, Melbourne, (1997).
[138] D. Yuen, personal communication, BHP Research Laboratories, Port Kembla, (1999).
[139] J.W. Hinze, H.C. Graham, “The Active Oxidation of Si and SiC in the Viscous Gas-Flow
Regime”, Journal of The Electrochemical Society, 123 [7], 1066-1073 (1976).
[140] H. Oikawa, “Lattice Diffusion in Iron - A Review”, Tetsu-to-Hagané, 68, 1489-1497 (1982).
[141] S. Mrowec, Defects and Diffusion in Solids, An Introduction, Material Science Monographs, 5,
Elsevier Scientific Publishing Company, New York, (1980).
[142] O. Kubaschewski, C.B. Alcock, Metallurgical Thermochemistry, Fifth Edition, Pergamon Press,
(1979).
[143] J. Park, C.J. Altstetter, Met. Trans. A, 18A, 43 (1987).
[144] F. Imada, T. Jodai, M. Hasebe, N. Murakami, T. Kobayashi, CAMP-ISIJ, 7, 886 (1994).

265
References

[145] G. Salje, M. Feller-Kniepmeier, “The Diffusion and Solubility of Copper in Iron”, Journal of
Applied Physics, 48 [5], 1833-1839 (1977).
[146] J. Goodyear, D. Wood, A. Stephenson, A. Stacey, “Some Effects of Metallic Residuals on the
Performance of Rod in a Wire Mill”, Wire Journal, 10 [12], 70-75 (1977).
[147] J.A. Nesbitt, “Numerical Modeling of High-Temperature Corrosion Processes”, Oxidation of
Metals, 44 [1/2], August, 309-338 (1995).
[148] R.W. Hornbeck, Numerical Methods, Quantum Publishers, New York, (1975).
[149] A.D. Romig (Jr.), N.Y. Pehlivanturk, O.T. Inal, “Modelling of Diffusion Processes: Numerical
Solutions to the Diffusion Equation”, Diffusion Analysis and Appications, Edited by A.D. Romig
(Jr.) and M.A. Dayananda, The Minerals, Metals and Materials Society, Chicago, Illinois, 45-77
(1988).
[150] B. Lee, K. Oh, “Numerical Treatment of the Moving Interface in Diffusional Reactions”, Z.
Metallkd., 87 [3], 195-204 (1996).
[151] B. Lee, “Numerical Procedure for Simulation of Multicomponent and Multi-Layered Phase
Diffusion”, Metals and Materials (South Korea), 5 [1], February, 1-15 (1999).
[152] J. Crank, The Mathematics of Diffusion, Second Edition, Oxford University Press, (1997).
[153] Y. Zhou, T.H. North, Modelling Simul. Mater. Sci. Eng., 1, 505 (1993).
[154] J. Henderson, L. Yang, Trans. Met. Soc. AIME, 221, 72 (1961).
[155] Smithels Metals Reference Book, Seventh Edition, Edited by E.A. Brandes and G.B. Brook,
Butterworth Heinemann, (1998).
[156] P.J.L. Fernandes, D.R.H. Jones, “Mechanism of liquid metal induced embrittlement”,
International Materials Reviews, 42 [6], 251-261 (1997).
[157] Fossil Fuel Combustion: A Source Book, Edited by W. Bartok and A.F. Sarofim, Wiley, New
York, (1991).
[158] Y.K. Rao, Stoichiometry and Thermodynamics of Metallurgical Processes, Cambridge
University Press, 140 (1985).
[159] D.R. Gaskell, An Introduction to Transport Phenomenon in Materials Engineering, Macmillan
Publishing Company, New York, (1991).
[160] C.R. Wilke, J. Chem. Phys., 18, 517-519 (1950).

266
APPENDIX
A
Appendix A.1 Composition in simulated reheat atmosphere
The following section outlines a calculation determining the composition of the
atmosphere in a reheat furnace using natural gas for combustion. Table A-1 lists the
composition of a typical natural gas.

Table A-1: Composition of natural gas [157]

Species Volume %
CO2 0.3
CH4 91.3
C2H6 1.5
C3H8 0.7
N2 5.4
Other hydrocarbons 0.8

Stoichiometric combustion (or 100% combustion) occurs when the mixture of fuel and
air contains just sufficient oxygen for complete combustion of the fuel. For combustion
of natural gas, the following reactions must be considered:

CH 4 + 2O 2 → CO 2 + 2H 2 O (A.1)

C 2 H 6 + 3 .5O 2 → 2CO 2 + 3H 2 O (A.2)

C 3 H 8 + 5O 2 → 3CO 2 + 4H 2 O (A.3)

Assuming that there is a total of 100 mol of natural gas present, the number of moles of
O2, n(O2), required for stoichiometric combustion of methane can be calculated as such:

n(O2 ) = 2 × 92.1
= 184.2

267
Appendix A

where the 0.8% of the ‘other hydrocarbons’ was added to the CH4 total in Table A-1.

If it is assumed that air consists of 20.94% O2 and 79.06% N2 [158], the amount of N2
present in the air accompanying the 184.2 moles of O2 can be calculated as such:

n(O2 )
n( N 2 ) = × 79.06
20.94
= 695.46

Similar calculations can be repeated for the reactions listed in Eqns. (A.2) and (A.3) and
the results have been summarised in Table A-2.

Table A-2: Amount of O2 and N2 for 100% combustion for combustion of methane,
ethane and propane

Equation O2 consumed (moles) N2 required (moles)


(A.1) 184.2 695.46
(A.2) 5.25 19.82
(A.3) 3.5 13.21
Total: 192.95 733.89*

* includes N2 present in natural gas (see Table A-1)

Complete combustion in a reheat furnace is often ensured by operating at 103-112% air.


In the case of 112% air combustion, the excess O2, n(O2)excess, can be calculated as such:

n(O2 )excess = n(O2 ) × 0.12


= 192.95 × 0.12
= 23.154

Similarly, the accompanying excess N2 can be calculated:

n(O2 )excess
n( N 2 )excess = × 79.06
20.94
= 87.42

268
Appendix A

Hence, taking into account the excess amounts of O2 and N2, the composition of the
natural gas plus air mix for 112% air combustion has been summarised in Table A-3.

Table A-3: Composition of the natural gas + air mixture for 112% air combustion

Number of moles
Species Mole Fraction
(per 100 mol of natural gas)
CO2 0.3 2.65×10-4
CH4 92.1 8.14×10-2
C2H6 1.5 1.33×10-3
C3H8 0.7 6.18×10-4
N2 821.311 0.726
O2 216.104 0.191

Chemix [133] was used to determine the equilibrium partial pressure of the components
after combustion of the gas mixture in Table A-3 at a total pressure of 1 atm at 1100°C.
The results have been summarised in Table A-4.

Table A-4: Calculated partial pressures of the equilibrium products of natural gas
combustion at 112% stoichiometric combustion at 1100°C

Species Partial Pressure (atm)


CO2 0.08602
CO 3.63×10-7
N2 0.724601
O2 0.020424
H2O 0.168953
H2 3.48×10-7

269
Appendix A

Appendix A.2 Conversion relationships between rate constants


Conversion between scale thickness and weight gain rate constant
The following outlines the method used to convert between the rate constants defined in
terms of scale thickness, k', and weight gain, kp. These constants can be defined as such:

X n = k't (A.4)

n
 ∆W 
  = k pt (A.5)
 A 

where X is the total scale thickness measured, t the oxidation time, ∆W is the weight
change of the sample during oxidation, A is the surface area over which the reaction
takes place and n = 1 for linear and 2 for parabolic oxidation. Figure A-1 illustrates the
formation of a unit volume of FeO on iron. The thickness of the scale, X, is equivalent
V
to , where V is the volume and A the area of the scale. The density of the oxide, ρoxide,
A
W
can be expressed as ρ oxide = , where W is the weight of the scale. As a result:
V

W
= ρ oxide ⋅ X (A.6)
A

A
Fe FeO

Figure A-1: Schematic representing formation of a unit volume of FeO on iron.

270
Appendix A

The weight of the scale can be replaced by the normalised weight change by noting that
the weight change is due to the oxygen component within the scale. As a result, the
relationship between the normalised weight change and thickness of scale produced can
be expressed as:

∆W
= f ⋅ ρ oxide ⋅ X (A.7)
A

where f is the weight fraction of oxygen within the scale. This assumes that the surface
area remains constant or changes minimally during the reaction (i.e. the scale only
grows in thickness). For FeO, f can be expressed as:

MWO
f = (A.8)
MWFe + MWO

where MWi is the molecular weight of component i. As a result, the conversion between
k' and kp can be expressed as:

k p = ( f ⋅ ρ oxide ) k '
n
(A.9)

If it is assumed that MWFe = 55.85 g/mol, MWO = 16 g/mol and ρFeO = 5.7 g/cm3 [30],
the conversion factor for the growth of FeO on iron becomes k p = 1.611 ⋅ k ` for

parabolic kinetics and k p = 1.269 ⋅ k ` for linear kinetics.

If a multilayered iron oxide scale exists (e.g. FeO, Fe3O4, Fe2O3), the treatment must be
modified. For the formation of an n-layered scale, Eqn. (A.7) can be modified:

∆W  n 
=  ∑ ri ⋅ f i ⋅ ρ i  ⋅ X (A.10)
A  i 

where ri is the thickness ratio fraction, f i the weight fraction of oxygen and ρ i the
density of scale layer i.

271
Appendix A

Taking MWFeO = 71.85 g/mol, MWFe3O4 = 231.55 g/mol, ρFe3O4 = 5.18 g/cm3, MW Fe2O3 =

159.7 g/mol, ρFe2O3 = 5.24 g/cm3 [30] and that the scale forms in the average thickness

ratio FeO:Fe3O4:Fe2O3 of 95:4:1, the following can be evaluated: fFeO = 0.223, fFe3O4 =
n
0.276, fFe2O3 = 0.301, ∑r ⋅ f
i
i i ⋅ ρ i = 1.279. Hence, the conversion factor for parabolic

kinetics becomes k p = 1.637 ⋅ k ` and for linear kinetics is k p = 1.279 ⋅ k ` .

Conversion between scale thickness and metal consumed rate constant


The total thickness of scale produced, X, and the amount of metal consumed per unit
length, l, is related by the expression (see Figure A-2):

Volume of oxide produced X


= (A.11)
Volume of metal consumed l

Gas Scale Metal

Original gas/metal interface

Figure A-2: Schematic representation of relationship between the scale thickness


and thickness of consumed metal.

Assuming that 1 mole of metal produces 1 mole of scale, the relationship becomes:

272
Appendix A

 Vm Scale 
X =  Metal ⋅l (A.12)
V 
 m 

where Vmi is the molar volume of the phase i. As a result, the appropriate relationship is:

n
 V Scale 
k ' =  m Metal  ⋅ kc (A.13)
V 
 m 

where n =1 for linear and 2 for parabolic kinetics. The molar volume can be determined
from the following equation:

i MWi
Vm = (A.14)
ρi

Taking MWFeO = 71.85 g/mol, ρFeO = 5.7 g/cm3, MWFe = 55.85 g/mol and ρFe = 7.86
g/cm3, VmFeO is calculated as 12.61 cm3/mol and VmFe as 7.11 cm3/mol. As a result,
k ' = 1.776 ⋅ k c for linear kinetics and k ' = 3.146 ⋅ k c for parabolic kinetics.

For situations in which a multilayered scale of FeO, Fe3O4 and Fe2O3 forms, Eqn.
(A.14) must be modified. For an m-phase system, the molar volume can be
approximated as:

m
V m = ∑ F iV m
i
(A.15)
i

where F i is the molar fraction of phase i. The molar fraction for each phase can be
determined by:

ρ i ri 1
Fi = ⋅
MWi m
ρ i ri
∑i MW
(A.16)
i

where ri is the thickness ratio of i.

273
Appendix A

Taking MWFe3O4 = 231.55 g/mol, ρFe3O4 = 5.18 g/cm3, MW Fe2O3 = 159.7 g/mol, ρFe2O3 =

5.24 g/cm3, Vm Fe3O4 was calculated as 44.7 cm3/mol, Vm Fe2O3 as 30.48 cm3/mol and Vm
as 13.06 cm3/mol. Assuming that the scale forms in the average thickness ratio
FeO/Fe3O4/Fe2O3 of 95:4:1, k ' = 1.84 ⋅ k c for linear kinetics and k ' = 3.39 ⋅ k c for

parabolic kinetics.

274
Appendix A

Appendix A.3 Mass transfer in viscous flow regime


The maximum flux of an oxidant in a flowing gas (where gas flow is in the viscous
regime) can be estimated using kinetic equations. Consider a 10 mm wide sample
(dimension of sample in direction of gas flow) oxidised in a simulated reheat
atmosphere at 1100°C (see Table A-4) at atmospheric pressure. The maximum flux of
oxidant i to the sample surface in the viscous flow regime can be approximated [139]:

k m (i ) p i
Ji = (A.17)
RT

where km(i) is the average boundary layer mass transfer coefficient for i (ms-1) and pi is
the partial pressure of i. The average boundary layer mass transfer coefficient is
estimated by [159]:

Sh ⋅ D
k m (i ) = (A.18)
L

where Sh is the Sherwood number, D the gas interdiffusion coefficient and L the sample
length in the direction of gas flow.

Assuming that the reaction gas is a binary mixture, e.g. O2 – N2, and that the component
molecules are non-polar and non-reacting, the interdiffusion coefficient can be
calculated from the Chapman-Enskog expression for an ideal gas [159]:

 1 1 
T 3  + 
 MW MW B 
(A.19)
DA − B = 1.8583 × 10 −3
A

pσ A − B Ω D
2

where MWi is the molecular weight of i, p is the pressure of the reaction gas, σ A− B is
average collision diameter of the reaction gas and ΩD is a dimensionless function of the
temperature and the interaction potential between A and B.

The average collision diameter of the reaction gas can be estimated by the arithmetic
average of the collision diameter of the components:
275
Appendix A

σA +σB
σ A−B = (A.20)
2

Taking σ O 2 and σ N 2 as 3.433 and 3.681 Å, respectively [159], then σ O 2 − N 2 = 3.557 Å.

kT
ΩD is a function of the dimensionless quantity where k is the Boltzman constant
ε A− B

and εA-B is the interaction energy for the gas A-B. The interaction energy for an O2 – N2
gas can be calculated from the geometric average of εB and εB:

( )
1
ε O2 − N2 = ε O2 × ε N2 2 (A.21)

ε O2 ε N2 kT
Taking the values [159] = 113 and = 91.5 , then = 13.46 . Interpolating
k k ε O2 − N 2

tabulated data [159] yields ΩD = 0.7158. Eqn. (A.19) then leads to DO 2 − N 2 = 2.70 × 10-4

m2s-1.

For flow past a parallel plate, the Sherwood number can be defined as [159]:

1
0.343
 η   Lvρ  2
Sh = 0.664    (A.22)
 ρD   η 

where η is the viscosity, ρ is the density of the species, D is the interdiffusion


coefficient, L is the length of the sample in the direction of gas flow and v is the velocity
of gas flow.

Taking the gas flow rate as 300 mL/min and the inner diameter of the furnace tube as 50
mm, yields v = 2.55×10-3 m/s at room temperature. The corresponding value at 1100°C
is calculated from the ideal gas equation to be 1.17×10-2 m/s.

The viscosity of a binary gas mixture can be calculated from Wilke’s formula [160]:

276
Appendix A

X Aη A X Bη B
η mix = + (A.23)
X A Φ A − A + X B Φ A − B X B Φ B− B + X A Φ B − A

where Xi is the mole fraction of i, ηi is the viscosity of pure i and

2

1
 1 1 
1  MWi  2
  η i 

2
 MW j  4

Φ i− j = 1+  
1 +  η
(A.24)
8  MW j   
  j   MWi  

The viscosity of a pure gas is given by the Chapman-Enskog expression [159]:

MWi T
η i = 2.6693 × 10 −5 (A.25)
σ i Ωi
2

kT kT
where Ωi is the collision integral of i. Taking = 12.15 and = 15.01,
ε O2 ε N2

interpolation of tabulated data [159] yields Ω O 2 = 0.8068 and Ω N 2 = 0.7837. As a result,

η O 2 = 5.88×10-4 g/cm⋅s and η N 2 = 4.93×10-4 g/cm⋅s. The viscosity of the O2 – N2

gas, ηmix , is thus 4.95× 10-5 kg/m⋅s.

If the density of O2 at 1100°C is taken as 0.284 kg/m3, then Sh = 4.69 × 10-1. Taking the
partial pressure of O2 in the simulated reheat atmosphere as 0.02 atm, then km (O2 ) =

1.27×10-2 m/s. The maximum flux of oxygen to the sample surface, J O 2 , is thus

2.25×10-3 mol/m2s or 7.19×10-2 g/m2s.

277
APPENDIX
B
Appendix B.1 Maximum copper-rich layer thickness
The following analysis describes the method used to determine the copper-rich layer
thickness formed in an iron-copper alloy assuming that all copper is rejected from the
scale and incorporated into the copper-rich layer, i.e. no diffusion of copper within the
metal. With reference to Figure B-1, a mass balance at the scale/copper-rich phase
interface requires that the amount of copper rejected due to scale formation equals the
amount of copper accumulated within the copper-rich phase:

o
(
l ⋅ N Cu = L ⋅ N Cu
Cu-rich
− N Cu
o
) (B.1)

where NCuo and NCuCu-rich are the concentration of copper in the original steel and
copper-rich phase, respectively; and l and L are the thickness of metal consumed and
copper-rich layer formed, respectively. The approximation is made that no copper
rejected from the oxidation process dissolved and diffuses into the underlying steel: all
is precipitated in the copper-rich layer.

278
Appendix B

Scale NCuCu-rich Fe-Cu alloy

Concentration

NCuo

l L
Distance
Original gas/metal surface

Figure B-1: Schematic of copper concentration profile in an iron-copper alloy


assuming that all copper is rejected from scale and that no diffusion of copper
occurs within the metal.

The concentration parameters in Eqn (B.1) were expressed in terms of mol. species/cm3
phase. Concentration can be expressed in this form using the following expression:

Ci
Ni = (B.2)
Vm

where Ni and Ci are the concentration of species i expressed in mol. species/cm3 and
atomic fraction, respectively; and Vm is the molar volume of the phase under
consideration. The molar volume can be determined from the following equation:

i MWi
Vm = (B.3)
ρi

where MWi and ρi are the molar weight and density, respectively of phase i.

The value of CCuCu-rich was taken as the equilibrium concentration of copper in the
copper-rich phase at the copper-rich/metal interface. This was calculated using Thermo-
Calc [126] (see Table B-1). As iron makes up a significant portion of the copper-rich

279
Appendix B

phase, ρCu-rich was calculated as a weighted fraction of its components. Taking MWFe =
55.847g/mol, MWCu = 63.546g/mol, ρFe = 7.87g/cm3 and ρCu = 8.96g/cm3, ρCu-rich was
calculated as 8.89g/cm3 and VmCu-rich as 7.1478cm3/mol at 1250°C. As a result, NCuCu-rich
= 1.31×10-1mol. Cu/cm3 copper-rich phase. The molar volume of the alloy was
approximated closely by the value for pure iron, i.e. VmFe-Cu = 7.0962cm3/mol. For a
0.47Cu-Fe alloy (≈0.41at%Cu), VmFe-Cu = 7.0962cm3/mol and NCuo = 5.82×10-4mol.
Cu/cm3 iron-copper alloy.

Table B-1: Copper concentration in copper-rich phase calculated using Thermo-


Calc [126] and an Fe-Cu-Ni-Sn database [144]

Temperature (°C) CCuCu-rich (wt%) CCuCu-rich (at%)


1250 93.6 92.8
1150 96.1 95.6
1050 96.7 96.3

The thickness of metal consumed is related to the scale thickness, X, by the following
relationships (see Appendix A.2):

For FeO-only
X = 1.78 ⋅ l (B.4)
formation:

For FeO/Fe3O4/Fe2O3
X = 1.84 ⋅ l (B.5)
formation:

As a result, all variables are now known to solve for L as a function of X in Eqn. (B.1).

280
Appendix B

Appendix B.2 Time taken for dissolution of a copper-rich layer


of known thickness
The aim of this calculation is to determine how long it would take for a copper-rich
phase of known thickness to dissolve into the metal if consumption of the metal (and
therefore enrichment of copper) was halted and the system held at the reaction
temperature. The equation describing diffusion from a fixed boundary where the
concentration at the boundary is constant is given by:

C Cu = C Cu
Metal / Cu − rich
(
− C Cu
Metal / Cu − rich
)
 x
− C Cu ⋅ erf 
o 
 (B.6)
 2 Dt 

where CCu is the copper concentration at distance x, CCuMetal/Cu-rich is the solubility limit
for copper in the metal substrate, CCuo is the bulk concentration of copper within the
metal substrate, D is the diffusivity for copper in the metal substrate and t is the time
available for diffusion. From Fick’s first law, the flux J of copper atoms across x = 0 is
given by:

∂C Cu
J =−D (B.7)
x =0
∂x x =0

∂C Cu
The quantity is given by:
∂x

Metal / Cu − rich
∂C Cu C Cu − C Cu
o
 x2 
= exp −  (B.8)
∂x πDt  4 Dt 

As a result, Eqn. (B.7) becomes:

J x =0
(
= C Cu
Metal / Cu − rich
− C Cu
o
) D
πt
(B.9)

The amount of copper entering the metal at x = 0 after time t, MMetal, is given by:

281
Appendix B

t
M Metal = ∫ J dt
0
(B.10)
(
= 2 C Cu
Metal / Cu − rich
− C Cu
o
) Dt
π

To simplify the calculation, the copper concentration within the copper-rich phase was
assumed uniform. Its value was approximated by the equilibrium concentration of
copper in the copper-rich phase at the copper-rich/metal interface of the iron-copper
system. Given a known thickness of copper-rich phase L, the amount of copper within
the copper-rich phase MCu-rich can be calculated as:

M Cu − rich = L.C Cu − rich (B.11)

where CCu-rich is the concentration within the copper-rich phase. The time taken for
dissolution of the copper-rich phase is given when MCu-rich = MMetal. Rearrangement of
Eqns. (B.10) and (B.11) yields:

2
π  L.C Cu −rich 
t= 
( o 
)
Metal / Cu − rich
(B.12)
D  2 C Cu − C Cu 

282
APPENDIX
C
Appendix C.1 Derivation of an equation describing diffusion
within the copper-rich phase
Require an equation that describes copper diffusion within the copper-rich layer, i.e.
diffusion between two moving interfaces where the concentration at each interface is
fixed. Non-steady state diffusion is described by Fick’s second law:

∂C i ∂ 2 Ci
= Di (C.1)
∂t ∂x 2

where Ci and Di are the concentration and diffusion coefficient within phase i,
respectively; t is the time and x is the distance from the boundary defined at t = 0. Eqn.
(C.1) assumes that D is not a function of concentration.

A variable is now required which is an appropriate combination of x and t that will


convert the partial differential equation to an ordinary differential equation. The
function is:

x
η= (C.2)
2 Dt

which is seen to be dimensionless. As a result:

∂η x
=−
∂t 4 Dt 3
and
∂η 1
=
∂x 2 Dt

283
Appendix C

Thus,

dC dC dη
= ⋅
dt dη dt
x  dC 
=−  
4 Dt 3  dη 
and
∂ 2C ∂  dC  ∂η 
=  
∂x 2
∂x  dη  ∂x 
1  d 2C 
=  
4 Dt  dη 2 

Substitution in Eqn. (C.1) yields:

x dC 1  d 2C 
− ⋅ =  
4 Dt 3 dη 4 Dt  dη 2 

which simplifies to:

dC 1  d 2C 
= −  2  (C.3)
dη 2η  dη 

Since both x and t have been eliminated, Eqn. (C.3) is an ordinary differential equation.

Let
dC
y=

and thus Eqn. (C.3) becomes:

1 dy
y=−
2η dη
or

284
Appendix C

dy
= −2ηdη
y

Integration yields:

ln y = −η 2 + ln A'

where ln A' is the integration constant. As a result:

y = A' exp − η 2( )

Making use of the definition of y yields:

∫ dC = A' ∫ exp(− η )dη


2
(C.4)

The boundary condition at the scale/metal interface is:

 x1 
CCu-rich = CCu-rich/Scale at x = x1 η =  and t > 0
 2 DCu − rich t 
 

Applying this to Eqn. (C.4) gives:

∫C
CCu − rich
Cu − rich / Scale
η
dC Cu − rich = A' ∫ x2 ( )
exp − η 2 dη
2 DCu − rich t

A' π   x  
 − erf  x1 

C Cu − rich − C Cu −rich / Scale = erf
2   2 DCu − rich t 

2 D
 Cu − rich t



or

285
Appendix C

  x   x1 
C Cu − rich − C Cu −rich / Scale = Aerf   − erf   (C.5)
  2 DCu −rich t 

2 D
 Cu − rich t



At x = x2, CCu-rich = CCu-rich/Metal. Hence, Eqn. (C.5) becomes:

  x2   x1 
C Cu − rich / Metal − C Cu − rich / Scale = Aerf   − erf  
  2 DCu − rich t  2 D 
Cu − rich t
(C.6)
  

Eliminating A from Eqns. (C.5) and (C.6) yields:

 x   x1 
erf   - erf  
C Cu − rich − C Cu − rich / Scale 2 D t  2 D t 
 Cu − rich   Cu − rich 
= (C.7)
C Cu − rich / Metal − C Cu − rich / Scale  x2   x1 
erf   − erf  
2 D t  2 D t 
 Cu − rich   Cu − rich 

which describes diffusion of copper within the copper-rich layer (x1 ≤ x ≤ x2).

286
Appendix C

Appendix C.2 Derivation of an equation describing diffusion


within the metal
Require an equation that describes copper diffusion within the metal when a growing
copper-rich layer exists at the metal surface, i.e. diffusion at a moving interface where
the concentration at the interface is fixed. For the metal phase, the initial condition is:

CMetal = Co at t = 0 (η = ∞) and x ≥ 0

and when applied to Eqn. (C.4) yields:


Co

C Metal

( )
dC Metal = B' ∫ exp − η 2 dη
η

 x 
C o − C Metal = Berfc  (C.8)
2 D t 
 Metal 

At x = x2, CMetal = CMetal/Cu-rich. Hence, Eqn. (C.8) becomes:

 x2 
C o − C Metal / Cu − rich = Berfc  (C.9)
2 D t 
 Metal 

Eliminating B from Eqns. (C.8) and (C.9) yields:

 x 
erfc 
C o − C Metal 2 D t 
=  Metal 
(C.10)
C o − C Metal / Cu − rich  x2 
erfc 
2 D t 
 Metal 

which describes copper diffusion within the metal (x ≥ x2).

287
Appendix C

Appendix C.3 Estimation of rate constant using an Arrhenius


relationship
The temperature dependence of the oxidation rate can be described by an Arrhenius
relationship:

 Q 
k p ' = A exp −  (C.11)
 RT 

where A is a constant, Q is the activation energy for oxide growth on the alloy, R the
universal gas constant and T the temperature. Lee [67] reported Q = 124kJ/mol for the
oxidation of A1006 in a simulated reheat furnace atmosphere (112% Air). The parabolic
rate constant for the 0.47Cu steel at 1050°C was determined to be 1.3×10-6cm2/s. As a
result, A = 1.05×10-1cm2/s.

At 1150°C, k p ' was calculated to be 3.0×10-6cm2/s. This rate constant (in terms of scale

thickness) can be expressed in terms of the thickness of metal consumed using the
relationship developed in Appendix A.2. As a result, k cp = 9.4×10-7cm2/s.

288

You might also like