Anda di halaman 1dari 108

RWTH Aachen

Institute of Ferrous Metallurgy

Study Integrated Thesis

Master Student
Alireza Saeed-Akbari
Matr. –Nr. 268696

Subject:

Investigation on Material Characterization and


Mechanical Properties of Ultra High Strength
Boron Steel at High Temperatures

Supervisors: Univ. Prof. Dr.-Ing. W. Bleck


M.Sc. Malek Naderi

December 2006
Acknowledgment

I would like to deeply appreciate the cooperations and


considerations of all of the colleagues at the Institute of Ferrous
Metallurgy (IEHK), RWTH Aachen University. I should give a
special thanks to Mr. Malek Naderi for his kind and permanent
supports and helps during the project.
The workshops, metallography and warm deformation
laboratories' technicians are those who are mostly engaged with
the related experiments of the current work, for which I am
really grateful.
Abstract

In new generation of car bodies, hot stamped boron steel parts


are widely used in safe guards, bumpers, A and B pillars. By
applying hot stamping technology, one can get a full martensitic
microstructure in the final product. Thus, investigation on the
behavior of boron steel at high temperatures would be the main
objective.
Current thesis presents a comprehensive set of laboratory works
in terms of the mechanical simulation of the industrial hot
stamping process. This was performed by means of the hot
compression tests on a quenchable boron steel used for the hot
stamping of the car bodies under the variety of
thermomechanical conditions.
The hardness and microstructural variations and dimensional
changes through martensitic transformation were studied.
Results of the present work show that the possibility for the
formation of bainite and ferrite phases during the experiments at
the temperatures far below the relevant phase regions in the
CCT diagram must be taken into account. Furthermore, the
lower isothermal compression temperatures and less amounts of
the applied strain during the continuous forming and quenching
experiments, increase the chance of having a full martensitic
microstructure in the final product.
Table of Contents
1. Introduction 1

2. Literature Survey 3
2.1. Introduction, 3 2.2. Hot Stamping Process Background, 3
2.3. Metallurgical Fundamentals, 6
2.3.1. Effects of Boron and Carbon, 6
2.3.2. Characterization of Bainitic Microstructures, 10
2.3.2.1. Isothermally Formed Bainite, 11
2.3.2.2. Continuously Cooled Bainite, 12
2.3.3. Continuously Cooled Ferritic Microstructures, 17
2.3.3.1. Bainitic or Acicular Ferrite, 19
2.3.3.2. Granular Ferrite, 19
2.3.4. Martensite and Martensitic Transformation in Steels, 21
2.3.4.1. Chemical Composition Effect, 23
2.3.4.2. Cooling Rate Effect, 24
2.3.4.3. Austenization Temperature Effect, 25
2.3.4.4. Quenching Media Effect, 26
2.3.4.5. Lath Martensite, 29
2.3.4.6. Medium Carbon Martensite, 30
2.3.5. Thermomechanical Behavior of Boron Steels, 31

3. Experimental Procedure 35
3.1. Material Characterization, 35
3.1.1. Chemical Composition, 35
3.1.2. Microstructure, 35
3.1.3. CCT Diagram Design, 36
3.2. Isothermal and Non-isothermal Compression Tests, 37
3.3. Hardness and Metallography Tests, 41

4. Results and Discussion 41


4.1. High Temperature Isothermal Compression Tests, 41
4.1.1. The Hardness Values and Microstructural Evolutions, 41
4.1.1.1. Results, 41
4.1.1.2. Discussion, 47
4.1.2. The Deformation Data Analysis, 53
4.1.2.1. Results, 53
4.1.2.2. Discussion, 57
4.1.3. The Dilatation Data Analysis, 60
4.1.3.1. Results, 60
4.1.3.2. Discussion, 65
4.2. Simultaneous Deformation and Quenching Tests, 67
4.2.1. The Strain Magnitudes Effect, 67
4.2.1.1. Results, 67
4.2.1.2. Discussion, 74
4.2.2. The Strain Rate Effect, 77
4.2.2.1. Results, 77
4.2.2.2. Discussion, 80
4.2.3. The Austenization Soaking Time Effect, 84
4.2.3.1. Results, 84
4.2.3.2. Discussion, 88
4.2.4. The Initial Deformation Temperature Effect, 91
4.2.4.1. Results, 91
4.2.4.2. Discussion, 95

5. Conclusions 98

References i
Chapter 1 - INTRODUCTION 1

CHAPTER

ONE
INTRODUCTION

Permanently increasing claims on passenger protection demand new solutions


of the automobile industry in consideration of steel light weight construction
[1]. The art or science of sheet metal stamping processes is challenged daily to
accommodate higher strength and thinner materials. Further, these materials
must be transformed into more complex shapes with fewer dies and increased
quality in the final part. High-strength and ultra high-strength steels have less
ductility, and hence less formability than lower strength steels. Thus, care must
be taken in part design and forming method selection. In addition, problems
such as springback1 are increased with yield strength and it must be accounted
for in the process design [2]. Hot stamping, i.e. simultaneous forming and
quenching, is a one-step manufacturing process for high and ultra high strength
steel profiles to be used as safety related structural components in car body
structures. In addition to the temperature-related improvement in the forming
properties of these steels, hot stamping makes it possible to further increase the
strength of materials during deformation. Within the process, thin-walled
profiles or blanks of hardenable steel are heated to the austenite region, placed
in a cooled forming tool where they are simultaneously hardened and formed to
the desired shape. Alternatively, the blanks are formed, clamped and
subsequently hardened by spray quenching and released. So the problem that is
solved by hot stamping is getting the advantages of the advanced high-strength
material's properties without the manufacturing limitations (such as springback)
in very high strength steels [3, 4].
Regarding the material selection to reduce weight and to improve the safety of
vehicles, car makers need very high resistance flat products with very good

1
Condition that occurs when a flat-rolled metal or alloy is cold-worked; upon release of the forming
force, the material has a tendency to partially return to its original shape because of the elastic recovery
of the material. This is called "springback" and influenced not only by the tensile and yield strengths,
but also by thickness, bend radius and bend angle.
Chapter 1 - INTRODUCTION 2

formability and toughness for structural and impact resistant parts.


At present, these objectives can be fulfilled by the use of HSLA steels or
aluminum. If the shape is complex, the use of aluminum is a solution for
weight saving but its strength is limited and it is an expensive solution. The use
of HSLA steels is a satisfactory solution if the shape is not very complex.
These steels exhibit good strength, weldability and impact resistance. However,
an increase in strength decreases formability. To solve this problem, the
solution is to separate the required characteristics, i.e. good formability and a
very high strength. The way to realize this objective is in the use of new
structural sheet of boron steel with very good formability which permits
complex shapes to be obtained [4]1. To partly overcome and quantify the
mentioned challenges, modeling and simulation are gaining increasing
importance in the product development of structural components whose
manufacturing is based on the thermomechanical forming. However, finite
element simulations of coupled thermomechanical processes require accurate
and efficient simulation tools as well as relevant material data based on testing
of the mechanical and thermal properties under reliably simulated processing
conditions. The final mechanical properties depend on the microstructural
evolution [5].
The current work presents a comprehensive set of laboratory works in terms of
the mechanical simulation of the industrial hot stamping process by means of
hot compression tests on a 22MnB52 steel under the variety of
thermomechanical conditions. The mechanical properties and microstructural
data are given in two independent parallel sections regarding the high-
temperature isothermal compression and simultaneous deformation and cooling
processes (Chapter 4) to find the most optimized way of providing higher
strength and hardness values besides the most achievable industrial standards.
The experimental conditions, i.e. temperature and deformation parameters, are
chosen to be comparable with different industrial applications and limitations.

1
It should be noted that the addition of boron in the range of 0.0005-0.005%wt to certain steels
increases the hardenability to great extent.
2
The experimental material is called BTR165 by the material provider.
Chapter 2 – LITERATURE SURVEY 3

CHAPTER

TWO
LITERATURE SURVEY

2.1. Introduction
Through the following pages, an overview regarding the fundamental concepts
and requirements of a successful hot stamping process and the microstructural
and mechanical characteristics of the ultra high strength boron steels are
presented. The role of boron on the thermomechanical behavior of steels and
the response of the experimental material to certain heat treating processes
followed by an appropriate high temperature deformation are described. The
current chapter is based on the definition of the hot stamping process, the
results and achievements of the previous investigations on different
optimization aspects of the hot deformation of boron steels, the role of alloying
elements, the relevant phase transformations and the final mechanical
properties of the studied materials.

2.2. Hot Stamping Process Background


The steels used in the automotive industry can be classified based on different
criteria. According to [6], the mentioned definitions are summarized as follow:

1. By metallurgical designation:
- Low-strength steels: interstitial-free (IF) and mild steels;
- Conventional high-strength steels: carbon-manganese (CMn),
bake hardenable (BH), interstitial free high-strength (IF-HS) and
high strength low-alloy steels (HSLA);
- Advanced-high-strength steels (AHSS): dual phase (DP),
transformation induced plasticity (TRIP), complex phase (CP)
and martensitic steels.
Chapter 2 – LITERATURE SURVEY 4

2. Mechanical properties – tensile strength:


- Low strength steels, LSS: tensile strength <270MPa;
- High strength steels, HSS: tensile strength 270 – 700MPa;
- Ultra-high-strength steels, UHSS: tensile strength > 700MPa.
3. Mechanical properties – total elongation versus tensile strength.

Due to continuously higher demands from different organizations and severe


legislation on passive automotive safety and effort to reduce vehicle emissions,
the use of high- and ultra high strength components in both car body and
closures have increased drastically during the last two decades. The
components in the car body that are commonly made of ultra high strength
steels are shown in figure 2-1 [7]. The different components in accordance to
this figure are:
1. Door beam;
2. Bumper beam;
3. Cross and side members;
4. A-/B- pillar reinforcement;
5. Waist rail reinforcement.

Figure 2-1- Components in car body using ultra high strength steels [7].

The hot stamping process, as was introduced in chapter one, uses boron steel
blanks which are first austenitized at a temperature of ~900°C and then formed
Chapter 2 – LITERATURE SURVEY 5

and quenched between cold tools. The forming operation at elevated


temperatures allows complex geometries to be obtained due to the high
formability of the hot material. The quenching results in a material with a very
high yield and tensile strength, which falls into the category of martensitic ultra
high strength steels. Moreover, the hardened component shows a dimensional
accuracy comparable to that of mild steel products manufactured with
conventional forming methods. It must be noted that hot stamping of
quenchable steels is a non-isothermal sheet forming process. Due to this, the
calculation of the temperature evolution in the blank is crucial, because of its
significant impact on the material deformation as well as on its final
microstructure after phase transformation. This temperature evolution is
controlled by the heat exchange between the tools and the sheets [7, 8]. Finally,
hot stamping technology can be summarized by the following steps:
1. Punching of blanks;
2. Austenization in a furnace;
3. Forming and hardening;
4. In some cases surface treatment by blasting or pickling.
Figure 2-2 shows the mentioned process as described above.

Figure 2-2- Schematic description of the hot stamping process [7].


Chapter 2 – LITERATURE SURVEY 6

2.3. Metallurgical Fundamentals


In this section as the major part of the current chapter, a brief introduction to
the effects of two important alloying elements – i.e. boron and carbon - on the
metallurgical characteristics of steels is given. Additionally, different phase
transformations in steels are evaluated based on their relation with the hot
deformation (simultaneous deformation and cooling process) of the ultra high
strength steels. The final part of current section is dedicated to an overview on
the effects of different deformation parameters on the phase transformations in
steels.

2.3.1. An Insight into the Effects of Boron and Carbon as Alloying


Elements on the Hardenability of Steels

The simplest version of analyzing the effects of alloying elements on iron-


carbon alloys would require the consideration of a large number of ternary
alloy diagrams over a wide temperature range. However, Wever [9] pointed out
that iron binary equilibrium systems fall into four main categories (Figure 2-3):
open and closed γ-field (austenite) systems, and expanded and contracted
γ-field systems. This approach indicates that alloying elements can influence
the equilibrium diagram in two ways:

• by expanding the γ-field, and encouraging the formation of austenite


over wider compositional limits. These elements are called γ-stabilizers.
• by contracting the γ-field, and encouraging the formation of ferrite over
wider compositional limits. These elements are called α-stabilizers.

The form of the diagram depends to some degree on the electronic structure of
the alloying elements which is reflected in their relative positions in the
periodic classification [10]. Among different alloying elements in steels, the
role of boron and carbon is described here (due to their relation to the topic of
the current work).
Chapter 2 – LITERATURE SURVEY 7

Figure 2-3- Classification of iron alloy phase diagrams: a. open γ-field; b. expanded
γ-field; c. closed γ-field; d. contracted γ-field [10].

Regarding the expanding γ-field elements, carbon and nitrogen are the most
important elements in this group. Although the γ-phase field is expanded, its
range of existence is cut short by compound formation (Figure 2-3b). The
Chapter 2 – LITERATURE SURVEY 8

expansion of the γ-field by carbon, and nitrogen, underlies the whole of the
heat treatment of steels, by allowing formation of a homogeneous solid solution
(austenite) containing up to 2.0 wt % of carbon or 2.8 wt % of nitrogen.
Boron (together with the carbide forming elements tantalum, niobium and
zirconium) is the most important element under the contracted γ-field category.
The γ-loop is strongly contracted, but is accompanied by compound formation
(Figure 2-3d).
Although only binary systems have been considered so far, when carbon is
included to make ternary systems the same general principles usually apply.
For a fixed carbon content, as the alloying element is added, the γ -field is
either expanded or contracted depending on the particular solute [10].
The astonishingly large effect of a minute percentage of boron on hardenability
of steel has been of intriguing interest to metallurgists since the advent of
commercial boron steels. Compared to other elements commonly added to steel
to increase hardenability, commercial boron steels exhibit the following major
unique characteristics: loss in the hardenability effect of boron on austenitizing
at relatively high temperature and marked variation in the hardenability effect
of boron with carbon content of steel [11].
A number of hardenability mechanisms have been proposed to explain these
observations, but they are capable of explaining only part of the reported
evidence. Of the proposed mechanisms, four have survived to the present. All
assume that boron influences hardenability by retarding the nucleation of ferrite
and that it does not influence the thermodynamic properties of the bulk
austenite or ferrite phases. The first assumption is based on observations that
boron does not significantly change the growth rate of ferrite or the formation
rate of pearlite and martensite. The second assumption is based on the small
amount of boron present [12].
With the possibility of boron concentrations reaching significant levels (i.e. for
the austenite grain size of more than 30µm as described in [12]), a number of
mechanisms for retarding ferritic nucleation can be considered as follows:
Chapter 2 – LITERATURE SURVEY 9

- Reduction in austenite grain boundary energy: this occurs by the


diffusion of boron to austenite grain boundaries and lowering
their energy; hence, making them less favorable sites for ferrite
nucleation;
- Reduction in diffusivity: by decreasing the self-diffusivity of iron
in austenite grain boundaries;
- Reduction in number of sites: the first mechanism that of grain
boundary reduction, assumes that the austenite grain boundary
can be treated like a continuum with nucleation possible at any
site on the boundary. However, if the crystallographic nature of
the grain boundary is considered, one finds that even in high
angle boundaries, where one would expect nucleation to be most
rapid, there are regions of relatively high and low atom density.
If regions of low atom density are favored sites for nucleation of
ferrite, it is possible that boron poisons them either through filling
up the free volume by segregation there, or through precipitating
on them as borocarbides. If boron contaminates half the sites, the
ferrite "C" curve on TTT diagram would be shifted by a factor of
two. If it contaminated all the sites, the ferrite would have to form
elsewhere at a reduced rate. One attractive feature of this
mechanism is that there is no theoretical limit to the possible "C"
curve shift, and another is that in principle there can be sufficient
boron atoms present to saturate the sites. Because of these
features, one is not pressed to explain why small amounts of
boron can have a large effect.
- Nucleation of ferrite on borocarbides: Fe23(BC)6 precipitation is
a precursor to ferrite formation. The proponents of both the
reduction in grain boundary energy and the reduction of
diffusivity theories interpret this observation on the basis that
borocarbides precipitation draws boron out of the grain boundary
and removes the inhibition effect. Those who favor a site-
Chapter 2 – LITERATURE SURVEY 10

competition mechanism can interpret the above observation, also.


They suggest that borocarbides block ferrite nucleation only
when they are small but they encourage ferrite nucleation on their
own interfaces when they are sufficiently large [12].
The hardenability effect of boron element and the austenitizing temperature are
related in a pattern of a single peak curve, i.e. there is an increasing trend
followed by a drop after passing through a peak in the hardenability value as
the austenizing temperature increases. The quenching operation for the
structural boron steel should be performed at a lowest possible temperature
above Ac3, and the austenitizing quenching operation under 900°C can be
applied to all kinds of the boron steel, but the austenitizing case-hardening or
quench operation above 900°C is only applicable to the low alloy boron steel
with grains not likely to develop [13].

2.3.2. Characterization of Bainitic Microstructures

The numerous terms created over the last 50 years to describe specific bainite
morphologies have led to some confusion, and it is suggested that the
commonly used terminologies do not adequately describe the full range of
bainitic microstructures which are observed [14]. Upper and lower bainite are
established terms describing microstructures which can easily be distinguished
using routine microscopy, and whose mechanisms of formation are well
understood. There are, however, a number of other descriptions of steel
microstructures which include the word 'bainite'. These additional descriptions
can be useful in communicating the form of the microstructure. But this must
be done with care, avoiding the natural tendency to imagine a particular
mechanism of transformation, simply because someone has chosen to coin the
terminology [15].
The morphological features of ferrous martensites have been rather well
characterized over the past decades. In comparison, the characterization of
bainitic microstructures and properties is much less complete. Bainite has
Chapter 2 – LITERATURE SURVEY 11

received relatively little attention, and a great deal of effort will be required to
understand the bainitic transformation more fully, particularly of bainite which
forms during continuous cooling [14].
In this part of the current chapter, the principles of the bainitic transformation
in addition to different bainite morphologies and its suggested definitions are
given. The target is to gradually bring readers' attention to the importance of
dividing the general concept of bainitic transformation into two major
categories: isothermally formed bainite and continuously cooled bainite. Later
on, the role of this second type of bainitic transformation – i.e. continuously
cooled bainite – will be discussed in chapter 4. It is then seen that having a
deep knowledge about the possible bainite morphologies can help to avoid any
misunderstandings of the final appeared phases in the microstructure of the
continuously cooled steels. As showing and describing all the possible bainite
morphologies are out of the discussions of the current report, the following
pages concentrate more on the definition and characteristics of the continuously
cooled bainite – i.e. granular bainite – in more details. For more information on
other bainitic transformation mechanisms and microstructures, please refer to
[14] and [15].

2.3.2.1. Isothermally Formed Bainite


Isothermal bainite is usually distinguished as 'upper' or 'lower' depending on
whether the carbides are distributed between individual ferrite regions or within
them, respectively. The carbides are usually cementite, although ε-carbide may
also be found in lower bainite. The difference between upper and lower bainite
is also based on whether the transformation temperature is above or below
approximately 350°C, although it has been shown that the distinction is not
universally applicable. Upper bainite comprises a lathlike morphology, and the
austenite/ferrite habit plane is thought to be near {111}γ/{110}α. Lower bainite
is generally reported to have a platelike morphology in isothermally
transformed steels, with an irrational habit plane somewhat further away from
{111}γ. The carbide in lower bainite generally consists of a single
Chapter 2 – LITERATURE SURVEY 12

crystallographic variant inclined to the apparent longitudinal axis of ferrite,


although multiple variants have also been reported (similar to those observed in
tempered martensite) [14].

2.3.2.2. Continuously Cooled Bainite


To the physical metallurgist, bainitic steels are recognized by the shape of their
CT diagram. In steels which are commercially important, a typical diagram
features the polygonal ferrite transformation (ferrite nose) shifted rightward to
regions of very slow cooling rate, thereby exposing a broad, flat bainite
transformation region. The advantage of having a CT diagram with a broad, flat
bainitic nose is that bainite with an almost constant transformation start
temperature can be produced over a wide range of cooling rates. Consequently,
bainite can be produced in heavy sections with little change in tensile
properties compared to thinner sections.
A typical CT diagram for a commercial bainitic steel is shown in figure 2-4 to
illustrate some of the important features of the bainitic transformation in
continuously cooled steels. In this figure, the bainitic transformation spans a
range of cooling rates from about 4°C/min to 600°C/min (measured between
800°C and 500°C). This range is typical of the rates experienced during
thermomechanical processing or heat treating in the commercial production of
steel components varying in thickness from 100 to 1000mm.
Chapter 2 – LITERATURE SURVEY 13

Figure 2-4- Continuous cooling-transformation diagram for a Ni-Cr-Mo steel.


Composition of steel (weight percent): 0.15C, 0.32Mn, 0.31Si, 2.72Ni, 0.41Mo [14].

Although the CT diagram in figure 2-4 indicates that bainitic microstructures


are generated over a wide range of cooling rates, the situation is complicated
because of the wide variations in microstructure which are actually observed.
For example, the light-optical microscope shows the appearance of an
'acicular' bainite microstructure (with some martensite) at a cooling rate of
461°C/min. At a much slower cooling rate of 3°C/min, a 'granular' bainitic
microstructure is produced.
In fact, the terms upper and lower bainite were originally used to describe
isothermal transformations in specific temperature regimes, but the terminology
is rather less meaningful (and even misleading) in describing the bainitic
transformation during continuous cooling were substantially different
microstructures can be obtained over a relatively constant range of
transformation temperatures.
Chapter 2 – LITERATURE SURVEY 14

Of all the unusual descriptions of bainitic microstructures, granular bainite is


probably the most useful and frequently used nomenclature. During the early
1950's, continuously cooled low-carbon steels were found to reveal
microstructures which consisted of 'coarse plates and those with an almost
entirely granular aspect', together with islands of retained austenite and
martensite (figure 2-5) [16].

Figure 2-5- Granular bainite in a Fe-0.15C-2.25Cr-0.5Mo wt% steel: Left picture,


light micrograph; Right picture, corresponding transmission electron micrograph [15].

Habraken and coworkers [17-20] called this morphology as 'granular bainite'


and the terminology became popular because many industrial heat-treatments
involve continuous cooling rather than isothermal transformation.
Granular bainite is supposed to occur only in steels which have been cooled
continuously; it cannot be produced by isothermal transformation.
Habrakan and Economopoulos [20] summarized their findings schematically in
the CT diagram which is presented here in figure 2-6.
Chapter 2 – LITERATURE SURVEY 15

Figure 2-6- Schematic representation of a CT diagram showing formation of granular


bainite (path I), upper bainite (path II), and lower bainite (path III) [14].

At relatively slow cooling rates, 'granular bainite' is formed (cooling path I).
At intermediate cooling rates (cooling path II), they reported the formation of
upper bainite. To form lower bainite, they suggested that an isothermal hold
just above the Ms temperature is required, as indicated by cooling path III in
figure 2-6.
One of the most complete studies on the nature of continuously cooled bainite
was carried out by Ohmori et al [21]. Their work examined various
microstructures which developed through both isothermal and continuous
cooling transformation in Ni-Cr-Mo steel. Using both replicas and thin foils,
they examined the fine morphological and crystallographic details in this alloy
and separated the various microstructures into three distinct classes which they
called bainite I, bainite II and bainite III (figure 2-7).
Chapter 2 – LITERATURE SURVEY 16

Figure 2-7- Schematic representation of the CT diagram of a Ni-Cr-Mo steel showing


three forms of bainite; Bainite I being a carbide-free form, bainite II being a form
similar to upper bainite, and bainite III being a form similar to lower bainite [14].

Bainite I consists of a carbide-free acicular ferrite with well-defined films of


retained austenite (and/or martensite) at the lath boundaries; bainite II is similar
to upper bainite, with cementite particles between the carbide free ferrite laths.
Bainite III is similar to the lower bainite, with cementite 'platelets' forming
within the laths. However, the acicular ferrite was found to be present in a lath
morphology, rather than in the plate morphology which is typically reported for
lower bainite.
The coarse plates referred to earlier in the current section (page 14) regarding
the granular bainite structure, do not really exist. They are in fact, sheaves of
bainitic ferrite with very thin regions of austenite between the sub-units
because of the low carbon concentration of the steels involved. Hence, on an
optical scale, they give an appearance of coarse plates (figure 2-5).
A characteristic (though not unique) feature of granular bainite is the lack of
carbides in the microstructure. The carbon that is partitioned from the bainitic
Chapter 2 – LITERATURE SURVEY 17

ferrite stabilizes the residual austenite, so that the final microstructure contains
both retained austenite and some high-carbon martensite. Consistent with
observations on conventional bainite, there is no redistribution of substitutional
solutes during the formation of granular bainite. The extent of transformation to
granular bainite is found to depend on the undercooling below the bainite-start
temperature. This is a reflection of the fact that the microstructure, like
conventional bainite, exhibits an incomplete reaction phenomenon.
The evidence therefore indicates that granular bainite is not different from
ordinary bainite in its mechanism of transformation. The peculiar morphology
is a consequence of two factors: continuous cooling transformation and a low
carbon concentration. The former permits extensive transformation to bainite
during gradual cooling to ambient temperature. The low carbon concentration
ensures that any films of austenite or regions of carbide that might exist
between sub-units are minimal, making the identification of individual platelets
within the sheaves rather difficult using light microscopy.
Finally, it is interesting that in an attempt to deduce a mechanism for the
formation of granular bainite, Habraken (1965) [18] proposed that the austenite
prior to transformation divides into regions which are rich in carbon, and those
which are relatively depleted. These depleted regions are then supposed to
transform into granular bainite [14, 15].

2.3.3. Continuously Cooled Ferritic Microstructures


In contrast to the equiaxed ferritic microstructures of conventionally hot-rolled
or cold-rolled-and-annealed steels, the ferritic microstructures formed by
decomposition of austenite and by virtue of alloying or rapid cooling, often
assume non-equiaxed morphologies. The temperature range in which the non-
equiaxed morphologies of ferrite form is intermediate to those at which
austenite transforms to equiaxed ferrite/pearlite and martensite. Therefore, this
range is the same as that in which bainitic microstructures form in medium-
carbon steels. However, the low-carbon steel ferritic microstructures formed at
intermediate temperatures differ in variety and form from classical bainitic
Chapter 2 – LITERATURE SURVEY 18

microstructures. Figure 2-8 shows continuous-cooling-transformation (CCT)


diagram for an HSLA plate steel evaluated by Thompson et al. [23].
Different parts of the diagram are labeled by the letters PF, WF, AF and GF,
which stand for polygonal ferrite, Widmanstätten ferrite, acicular ferrite and
granular ferrite, respectively.
The general reaction of austenite to ferrite implies rejection of carbon into
retained austenite, according to the dynamic solubility limits of ferrite. At very
high cooling rates, even in very-low-carbon steels or irons with sufficient
hardenability, austenite may transform to martensite.

Figure 2-8- Continuous-cooling-transformation diagram of HSLA steel containing in


mass%, 0.06C, 1.45Mn, 1.25Cu, 0.97Ni, 0.72Cr, 0.42Mo [23].

In addition to the relatively well-characterized forms of ferrite which form from


austenite at high temperatures, types of ferrite which form from austenite at
intermediate temperatures are now commonly observed in continuously cooled
low-carbon steels [22].
Chapter 2 – LITERATURE SURVEY 19

Among different ferrite morphologies described by Kraus and Thompson [22],


'bainitic (or acicular) ferrite' and 'granular (or granular bainitic) ferrite' are
those related to the intermediate formation temperatures of ferrite. As the
mentioned temperature range is also applicable for the bainitic transformation,
its concept is briefly discussed here.

2.3.3.1. Bainitic or Acicular Ferrite


With increasing cooling rates, the austenite of low-carbon and ultra low-carbon
steels transforms to much finer ferrite crystals than described above. The most
commonly used terms for the resulting ferritic microstructures are bainitic
ferrite and acicular ferrite. The transformation temperatures for the formation
of these ferritic microstructures are clearly in the intermediate temperature
range as shown in the continuous cooling transformation diagram of figure 2-8.
Although the austenite decomposition is only to ferrite, coexisting with retained
austenite or M/A constituent, the microstructural arrangement of acicular
shaped ferrite crystals in groups of parallel laths is included in the Ohmori et al.
[21], bainite classification as BI bainite and in the Bramfitt and Speer bainite
classification [14] as B2, acicular ferrite with interlath austenite. Thus, the
literature describes the fine non-equiaxed ferritic intermediate temperature
austenite transformation product as both ferrite and bainite.

2.3.3.2. Granular Ferrite


Granular bainitic ferrite or granular ferrite, GF, has many similarities to bainitic
or acicular ferrite, but there appear to be morphological differences which merit
a separate category of austenite-to-ferrite transformation. Microstructures
consisting of granular bainite also form in the intermediate austenite
transformation range, as shown in CCT diagram of figure 2-8.
Although acicular and granular ferrites form over the same transformation
temperature range, the cooling rates which form granular ferrites appear to be
somewhat slower than those which form acicular ferrite [22].
Chapter 2 – LITERATURE SURVEY 20

Similar to acicular ferrite microstructures, the microstructure of granular ferrite


coexists with dispersed retained austenite or M/A particles in a featureless
matrix which may retain the prior austenite grain boundary structure. However,
in contrast to the acicular ferrite microstructures, the dispersed particles have a
granular or equiaxed morphology. TEM images show that the ferritic matrix
consists of fine ferrite crystals, containing high densities of dislocations,
separated by low-angle grain boundaries. As for acicular ferrite
microstructures, the low-angle boundaries explain the insensitivity of the
matrix ferrite crystals to etching for light microscopy. The ferrite crystals have
granular or equiaxed shapes which cause enclosed retained austenite or M/A
regions, by default, to have the granular or equiaxed shapes resolvable in light
micrographs.
Chapter 2 – LITERATURE SURVEY 21

2.3.4. Martensite and Martensitic Transformation in Steels

Perhaps the most important allotrope of iron is 'martensite', a chemically


metastable substance with about four to five times the strength of ferrite [24].
The name martensite is after the German scientist Martens [25]. A minimum of
0.4 wt% of carbon is needed in order to form martensite [24]. It was originally
described as the hard microconstituent found in quenched steels. Martensite
remains of greatest technological importance in steels where it can confer an
outstanding combination of strength (>3500 MPa) and toughness
(>200 MPa m1/2). Martensite can form at very low temperatures, where
diffusion, even of interstitial atoms, is not conceivable over the time period of
experiment. The highest temperature at which martensite forms is known as the
martensite-start, or Ms temperature. Although it is obvious that martensite can
form at low temperatures, this is not necessary to occur. Therefore, a low
transformation temperature is not sufficient evidence for diffusionless
transformation.
Martensite plates can grow at speeds of sound in the metal. In steel this can be
as high as 1100ms-1, which compares with the fastest recorded solidification
front velocity of about 80ms-1 in pure nickel. Such high speeds are inconsistent
with diffusion during transformation [25].
When the austenite is quenched to form martensite the carbon is 'frozen' in
place when the cell structure changes from FCC to BCC. The carbon atoms are
much too large to fit in the interstitial vacancies and thus distort the cell
structure into a Body Centered Tetragonal (BCT) structure [24]. The chemical
composition of martensite can be measured and shown to be identical to that of
the parent austenite. The totality of these observations demonstrate
convincingly that martensitic transformation are diffusionless [25].
The heat treatment process for most steels to get a full martensitic
microstructure involves heating the alloy until austenite forms, then quenching
the hot metal in water or oil, cooling it so rapidly that the transformation to
ferrite or pearlite does not have time to take place. The transformation into
Chapter 2 – LITERATURE SURVEY 22

martensite, by contrast, occurs almost immediately, due to the lower activation


energy.
Martensite has a lower density than austenite, so that the transformation
between them results in a change of volume. In this case, expansion occurs.
Internal stresses from this expansion generally take the form of compression on
the crystals of martensite and tension on the remaining ferrite, with a fair
amount of shear on both constituents. If quenching is done improperly, these
internal stresses can cause a part to shatter as it cools; at the very least, they
cause internal work hardening and other microscopic imperfections. It is
common for quench cracks to form when water quenched, although they may
not always be visible.
At this point, if its carbon content is high enough to produce a significant
concentration of martensite, the resulted product is extremely hard but very
brittle. Often, steel undergoes further heat treatment at a lower temperature to
destroy some of the martensite (by allowing enough time for cementite, etc., to
form) and help settle the internal stresses and defects. This softens the steel,
producing a more ductile and fracture-resistant metal. Because time is so
critical to the end result, this process is known as 'tempering', which forms
tempered steel [24].
Ideally, as mentioned before, the martensite reaction is a diffusionless shear
transformation, highly crystallographic in character, which leads to a
characteristic lath or lenticular microstructure.
The martensite reaction in steels is the best known of a large group of
transformations in alloys in which the transformation occurs by shear without
change in chemical composition. The generic name of martensitic
transformation describes all such reactions.
It should however be mentioned that there is a large number of transformations
which possess the geometric and crystallographic features of martensitic
transformations, but which also involve diffusion. Consequently, the broader
term of shear transformation is perhaps best used to describe the whole range of
possible transformations.
Chapter 2 – LITERATURE SURVEY 23

The martensite reaction in steels normally occurs athermally, i.e. during


cooling in a temperature range which can be precisely defined for a particular
steel. The reaction begins at a martensitic start temperature Ms which can vary
over a wide temperature range from as high as 500°C to well below room
temperature, depending on the concentration of γ-stabilizing alloying elements
in the steel [26].

2.3.4.1. Chemical Composition Effect


The martensite start temperature, Ms, is of vital importance for engineering
steels. Hence, great efforts have been made in predicting the Ms's of steels.
Obviously, chemical composition of a steel is a main factor in affecting its Ms
although the microstructure, (dislocation, vacancies, grain, twin, interphase
boundaries, and precipitates), external stress and plastic deformation, may
sometimes play an important role, too.
The Ms temperature of engineering pure iron is estimated as 540°C. C, Mn, Mo,
Cr and Si decrease the Ms while Mo increases the Ms. The analysis indicates
that most alloying elements have similar effects upon the Ms and A3
temperature.
The interactions between substitutional alloying elements can play an important
role in changing the Ms temperature. The Si-Mn interaction strongly increases
the Ms, while Si-Mo interaction significantly decreases the Ms. So far, there is
no proper physical explanation for this though supportive evidence has been
obtained from phenomenological result. Mn and Mo have the weakest
interaction. Si and Mo themselves have weak influence but their overall effect
depends further on the concentration of other alloying elements because of the
strong interactions found with other alloying elements [27].
Once the Ms is reached, further transformation takes place during cooling until
the reaction ceases at the Mf temperature. At this temperature all the austenite
should have transformed to martensite but frequently, in practice, a small
proportion of the austenite does not transform. Larger volume fractions of
Chapter 2 – LITERATURE SURVEY 24

austenite are retained in some highly alloyed steels, where the Mf temperature
is well below room temperature.
To obtain the martensitic reaction, it is usually necessary for the steel to be
rapidly cooled, so that the metastable austenite reaches Ms. The rate of cooling
must be sufficient to suppress the higher temperature diffusion-controlled
ferrite and pearlite reactions, as well as other intermediate reactions such as the
formation of bainite. The critical rate of cooling required is very sensitive to the
alloying elements present in the steel and, in general, will be lower as the total
alloy concentration is higher [26].

2.3.4.2. Cooling Rate Effect


In general, the martensitic transformation temperature is dependent on the
cooling rate when cooling rate is not high; above a critical cooling rate,
however, the starting temperature of the transformation is constant. Although
the constant starting temperature had been established many years ago, the
issue whether the Ms is constant and independent of the cooling rate was often
raised. In iron-base alloys, it is often observed that the transformation
temperature versus cooling rate curve show two plateaus when cooling rates
exceed a critical cooling rate (figure 2-9).

Figure 2-9- Relation between the transformation temperature of iron and the cooling
rate (0.006 – 0.039%C) [28].
Chapter 2 – LITERATURE SURVEY 25

In such a case, the plateau at the lower temperature is thought to be the Ms


temperature and the one at the higher temperature to be the A3 temperature (for
iron-base alloys), corresponding to the largest supercooling [28].

2.3.4.3. Austenization Temperature Effect


It has been reported that the higher the austenization temperature, the higher the
Ms temperature. Figure 2-10 shows an example, in which the broken line
indicates that the γ grain size increases as the austenization temperature
increases. Also, the longer the heating time, the higher the Ms temperature
(figure 2-11).

Figure 2-10 – Change of Ms temperature and austenite grain size with austenitizing
temperature (Fe – 0.33%C – 3.26%Ni – 0.85%Cr – 0.09%Mo; heating time 2 min for
800°C – 1000°C, 1 min for >1000°C) [28].
Chapter 2 – LITERATURE SURVEY 26

Figure 2-11- Change of Ms temperature with heating time of austenization (same


alloy as in figure 2-10; heating temperature 800°C) [28].

2.3.4.4. Quenching Media Effect


As to the interpretation of this fact, it must be noted that a lower quenching
temperature produces more frozen-in vacancies and hence more nucleation
sites. But it is uncertain how effective this phenomenon actually is. On the
other hand, a lower quenching medium temperature must produce a larger
thermal strain during quenching; hence it is expected to raise the Ms
temperature. This effect, however cannot be very large. A more likely cause of
raising the Ms temperature is the reduction of the energy needed for the
complementary shear during transformation, which originates in the
elimination of lattice imperfections due to heating to a higher temperature [28].
Each grain of austenite transforms by the sudden formation of thin plates or
laths of martensite of striking crystallographic character. The laths have a well-
defined habit plane and they normally occur on several variants of this plane
within each grain. The habit plane is not constant, but changes as the carbon
content is increased.
Martensite is a supersaturated solid solution of carbon in iron which has a
body-centred tetragonal structure, a distorted form of bcc iron. It is interesting
to note that carbon in interstitial solid solution expands the fcc iron lattice
uniformly, but with bcc iron the expansion is nonsymmetrical giving rise to
tetragonal distortion.
Chapter 2 – LITERATURE SURVEY 27

Analysis of the distortion produced by carbon atoms in the several types of site
available in the fcc and bcc lattices, has shown that in the fcc structure the
distortion is completely symmetrical, whereas in the bcc one, interstitial atoms
in z positions will give rise to much greater expansion of iron-iron atom
distances than in the x and y positions.
Martensitic planes in steel are frequently not parallel-sided; instead they are
often perpendicular as a result of constraints in the matrix, which oppose the
shape change resulting from the transformation. This is one of the reasons why
it is difficult to identify precisely habit planes in ferrous martensite.
Perhaps the most striking advances in the structure of ferrous martensites
occurred when thin foil electron microscopy was first used on this problem.
The two modes of plastic deformation are needed for the in-homogeneous
deformation part of the transformation, i.e. slip and twinning. All ferrous
martensites show very high dislocation densities of the order of 1011 to 1012
cm-2, which are similar to those of very heavily cold-worked alloys. Thus it is
usually impossible to analyze systematically the planes on which the
dislocations occur or determine their Burgers vectors.
The lower carbon (<0.5% C) martensites on the whole exhibit only
dislocations. At higher carbon levels very fine twins (5-10 nm wide) commonly
occur. In favorable circumstances the twins can be observed in the optical
microscope, but the electron microscope allows the precise identification of
twins by the use of the selected area electron diffraction technique. Thus the
twin shears can be analyzed precisely and have provided good evidence for the
correctness of the crystallographic theories discussed above. However,
twinning is not always fully developed and even within one plate some areas
are often untwined. The phenomenon is sensitive to composition.
The evidence suggests that deformation by dislocations and by twinning are
alternative methods by which the lattice invariant deformation occurs. From
general knowledge of the two deformation processes, the critical resolved shear
stress for twinning is always much higher than that for slip on the usual slip
plane. This applies to numerous alloys of different crystal structure.
Chapter 2 – LITERATURE SURVEY 28

Thus it might be expected that those factors, which raise the yield stress of the
austenite and martensite, will increase the likelihood of twinning. The
important variables are:
- carbon concentration;
- alloying element concentration;
- temperature of transformation;
- strain rate.
The yield stress of both austenite and martensite increases with carbon content,
so it would normally be expected that twinning would, therefore, be
encouraged. Likewise, an increase in the substitutional solute concentration
raises the strength and should also increase the incidence of twinning, even in
the absence of carbon, which would account for the twins observed in
martensite in high concentration binary alloys such as Fe-32%Ni.
A decrease in transformation temperature, i.e. reduction in Ms, should also help
the formation of twins and one would particularly expect this in alloys
transformed, for example, well below room temperature.
It should also be noted that carbon concentration and alloying element
concentration should assist by lowering Ms. As martensite forms over a range
of temperatures, it might be expected in some steels that the first formed plates
would be free of twins whereas the plates formed nearer to Ms would more
likely be twinned.
However, often plates have a mid-rib along which twinning occurs, the outer
regions of the plate being twin-free. This could possibly take place when the Ms
is below room temperature leading to twinned plates which might then grow
further on resting at room temperature.
Chapter 2 – LITERATURE SURVEY 29

2.3.4.5. Lath Martensite


The lath martensite structure is one of the most important structures in steels. It
is composed of fine substructures, i.e. "packets" which are a group of laths with
almost the same habit plane, and "blocks" which contain a group of laths with
almost the same orientation (figure 2-12). A prior austenite grain is divided by
several packets which are subdivided by blocks. It was recently shown that the
blocks are further subdivided by sub-blocks in low carbon steels [29].

Figure 2-12- OM images (3% nital etched) of lath martensite structures in the Fe-
0.2C-2Mn alloy: a) prior austenite grain size is 370 µ m and b) 28 µ m, respectively
[30].

This type of martensite is found in plain carbon and low alloy steels up to about
0.5 wt% carbon. The morphology is lath-like, where the laths are very long.
These are grouped together in packets with low angle boundaries between each
lath, although a minority of laths is separated by high angle boundaries. In plain
carbon steels practically no twin-related laths have been detected [26].
Since these packet and block boundaries are high angle boundaries, the
constituents are considered to be affective grains. Thus, the strength and
toughness of lath martensitic steels are strongly related to packet and block
Chapter 2 – LITERATURE SURVEY 30

sizes. It is known that both the block width and the packet size are proportional
to the prior austenite grain size. Usually, the packet size is taken as the
effective grain size for the strength and toughness of low carbon steels [30].

2.3.4.6. Medium Carbon Martensite


It is perhaps unfortunate that the term acicular is applied to this type of
martensite because its characteristic morphology is that of perpendicular plates,
a fact easily demonstrated by examination of plates intersecting two surfaces at
right angles (figure 2-13).
These plates first start to form in steels with about 0.5% carbon, and can be
concurrent with lath martensite in the range 0.5 %-l.0 % carbon. Unlike the
laths, the lenticular plates form in isolation rather than in packets, on planes
approximating to {225} and on several variants within one small region of a
grain, with the result that the structure is very complex [26].

Figure 2-13- Plate-like martensite microstructure.


Chapter 2 – LITERATURE SURVEY 31

2.3.5. An Overview of the Previous Investigations on the


Thermomechanical Behavior of Boron Steels

Somani et al. [31] examined the effects of plastic deformation on dilatation


during the martensitic transformation in a B-bearing steel (i.e. Docol Bo 02)1.
Their results show that plastic deformation of austenite at high temperatures
enhances ferrite formation significantly and consequently, the dilatation
decreases markedly even at a cooling rate of 280°C/s.
It was found that, without plastic deformation, Ms and Mf were about 425°C
and 280°C, respectively. The change in diameter was about 0.53%
corresponding to a relative volume change of 3.2%. They mentioned that the
reason for the drastic decrease of dilatation and drop of the Ms value to 375°C
due to an increase in the prior plastic strain could be justified as a result of the
stabilization of austenite by means of plastic deformation and the presence of
retained austenite in this regard. There were, however, distinct differences in
the high temperature slopes of the dilatation curves. The slope in the deformed
specimens being smaller than that in non-deformed ones. This presumably
indicates that some ferrite formed at higher temperatures as strain-induced,
consequently, less martensite is present.
Microstructures examination also revealed that, at a cooling rate of 50°C/s, the
ferrite content was about 20~40%. Hardness measurements also confirmed that
the structure formed after severe plastic deformation was markedly softer,
about 295~375 HV10, compared to the martensite hardness of 490~500HV10.
However, martensite was still present in considerable amount, even though the
dilatation became very small. Therefore, they suggested that some other
factors, such as residual stresses due to prior plastic deformation may be an
additional reason for the decrease of dilatation.
Finally they found that, the severe plastic straining (strain 0.8~1.0) during
continuous cooling at 50°C/s results in a much lower final flow stress level
(800-950MPa at 300~200°C) than that obtained for martensitic structure in
isothermal tests (1650~1900MPa).
1
The chemical composition in wt% is: 0.22C- 0.29Si- 1.1Mn- 0.013P- 0.003S- 0.21Cr- 0.0034B- 0.05
Al- 0.0025N.
Chapter 2 – LITERATURE SURVEY 32

Another investigation by the mentioned authors [5] revealed that the Ms


temperature is lowered by about 25-70°C with increasing plastic strain from
0.16-0.39. As the reason for this, they proposed that, as a consequence of
ferrite formation, carbon becomes enriched in the remaining austenite, which
therefore transforms into martensite at a somewhat lower temperature.
It was also observed that ferrite with an ultra-fine grain size can be formed as
strain-induced by subjecting austenite to severe plastic straining at temperatures
slightly above Ar3. Hardness measurements also confirmed that the
microstructure formed after a high-temperature plastic deformation was
remarkably softer, 302-440 HV10, while the martensite had a hardness of 490-
510 HV10, which was justified due to the presence of ferrite in the
microstructure as described before.
In case of the hardness measurements, their image analysis data were in
contrast with their hardness values and made it rather impossible to determine
the martensite or bainite phases based on the optical microscopy images. They
believed that the distinction between the bainite and martensite phases might
require transmission electron microscopic examinations, which had, however
not been performed, because this matter was not very important in their
discussions.
To avoid the strain-induced phase transformation, it was suggested that the
consequences of the prior plastic deformation should be small enough or
disappear before the temperature reaches the ferritic regime level. This means
that forming should take place at a high temperature, >800°C, where the
driving force for the austenite decomposition is low, or the time should be long
enough for static recrystallisation to occur. Another, more realistic alternative
might be forming at low temperatures, such as <600°C, i.e. below the ferrite
regime. In that case, ferrite nucleation is not accelerated, although some
enhancement of bainite formation may occur. This may not be so detrimental,
however, due to the notably smaller strength difference between martensite and
bainite. Furthermore, in order to avoid straining to continue at the martensitic
stage, which would mean excessive forming loads, a major springback and high
Chapter 2 – LITERATURE SURVEY 33

residual stresses, forming should be finished above 420°C, which means that
the proper temperature range is quite narrow. Overall, it was proposed that,
minimization of the plastic strain, maximization of the cooling rate and/or
forming at 450-600°C may be suitable ways to avoid excessive ferrite
formation and to achieve the desired mechanical properties in formed and
quenched components [5].
In the last reviewed work here, Jun and coworkers [32] studied the effects of
thermomechanical processing on the microstructures and transformations of
low carbon HSLA steels with and without boron. Microstructures observed in
continuous cooled specimens were composed of pearlite, quasi-polygonal
ferrite, granular bainite, acicular ferrite, bainitic ferrite, lower bainite, and
martensite depending on cooling rate and transformation temperature. Fast
cooling rate depressed the formation of pearlite and quasi-polygonal ferrite,
which resulted in higher hardness. However, hot deformation slightly increased
transformation start temperature, and promoted the formation of pearlite and
quasi-polygonal ferrite. Hot deformation could also strongly promote the
acicular ferrite formation which was not formed in non-deformation condition.
Small boron addition effectively reduced the formation of pearlite and quasi-
polygonal ferrite and broadened the cooling rate region from bainitic ferrite and
martensite. Impurity boron segregates to grain boundaries and improves the
grain boundary cohesive strength. This causes the mentioned effective
suppression of pearlite and/or ferrite formation compared to other substitution
elements. Microhardness of granular bainite varied from 220 to 250 HV, which
resulted from high dislocation density and hard constituents. Transformation
mechanism of these bainite-like microstructures had both aspects of diffusional
and shear mechanisms. It was suggested that granular bainite forms because
carbon quickly diffuses away from the ferrite/austenite interface at relatively
slow cooling rates, preventing the formation of cementite. The increased
carbon content in the remaining austenite can stabilize austenite from further
transformation, and this entrapment of residual austenite leads to granular
Chapter 2 – LITERATURE SURVEY 34

bainite morphology. Shear mechanism for bainite-like transformation was


proposed to be more dominated as increasing cooling rates.
It was also found that the deformation causes the formation of acicular ferrite,
pearlite, and quasi-polygonal ferrite, otherwise prevents the martensite
compared to that of non-deformed condition. The corresponding transformation
curves of deformed CCT moves toward left side compared to those of non-
deformed CCT [32].
Chapter 3 – EXPERIMENTAL PROCEDURE 35

CHAPTER

THREE
EXPERIMENTAL PROCEDURE

In the current chapter, firstly the investigated material is introduced. Then the
complete set of the experimental conditions and parameters are discussed.

3.1. Material Characterization

3.1.1. Chemical Composition


The studied material is 22MnB5 steel. This is a hot rolled boron steel in form
of the plates produces by Benteler company, Germany. The manufacturer calls
the mentioned product as BTR165 steel. The chemical analysis of the
investigated steel is given in table 3-1.

Table 3-1- Chemical composition of the experimental alloy, wt-%.


C Si Mn P S Cr Ti B
0.24 0.27 1.14 0.015 0.001 0.17 0.036 0.003

3.1.2. Microstructure
This steel contains ferrite and pearlite phases (together with carbide) in
as-received condition. Figure 3-1 shows the microstructure of the BTR165 steel
in the rolling direction.
Chapter 3 – EXPERIMENTAL PROCEDURE 36

a b
Figure 3-1- Microstructure of the as-received BTR165 sheets in the rolling direction
a) 500X, and b) 1000X.

The image analysis data shows that the microstructure contains around 78%
ferrite besides 22%, combination of pearlite and carbide. Ferrite grain size was
measured to be comparable with 11 ASTM grain size standard.

3.1.3. CCT Diagram Design


The Continuous Cooling Transformation (CCT) diagram, figure 3-2, has been
produced by means of dilatometry tests, metallographic investigations and
hardness measurements. The circled numbers indicate the final hardness values
in the HV10 scale.
Chapter 3 – EXPERIMENTAL PROCEDURE 37

Cooling rate = 30°C/s

Figure 3-2- CCT diagram of BTR165 steel.

For the heating speed of 5°C/s, the eutectoid reaction temperature, Ac1, is
722°C and the start temperature of austenite to primary ferrite transformation,
Ac3, reaches 870°C. After austenization at 900°C for five minutes followed by
quenching the microstructure becomes fully martensitic (point M in figure 3-2).
Consequently, the steel is classified in the ultra high strength steels grade. The
martensite start point, Ms, lies at 410°C and the martensite finish point, Mf, at
230°C. It can be seen that a cooling rate greater than 30°C/s results in a
martensitic microstructure. At the lower cooling rates, bainite (zone B in figure
3-2) or even ferrite (zone F in figure 1), can be formed resulting in the lower
hardness and the lower strength levels.

3.2. Isothermal and Non-isothermal Compression Tests


In the hot stamping process, the material is subjected to a temperature history
of heating and subsequent high cooling rate to ensure the formation of
martensite. Since thermo-mechanical history of the material will affect its final
microstructure and properties, appropriate deformation and temperature
histories must be applied when carrying out the material testing. A
Baehr DIL 805 deformation dilatometer (figure 3-3) was used to create the
Chapter 3 – EXPERIMENTAL PROCEDURE 38

thermo-mechanical schedules. Such conditions were produced by several


isothermal compression and quenching tests as well as simultaneous forming
and quenching tests at temperature range between 600°C – 900°C. Different
strain rates between 0.1 s-1 – 10.0 s-1 for isothermal and 0.07 s-1 – 0.4 s-1 for
non-isothermal tests were applied. Due to the technical limitations in the Baehr
deformation dilatometer, higher strain rates and lower temperature limits could
not be applied in the simultaneous forming and quenching tests. In all of the
experiments, the samples were austenitized at 900°C for five minutes and
quenched to the compression temperature by 50°C/s. The above mentioned
processes are illustrated in figure 3-4. The yield and maximum stress, Ms and
Mf, were received from the dilatation tests.
Chapter 3 – EXPERIMENTAL PROCEDURE 39

Figure 3-3- Baehr DIL 805 deformation dilatometer, and the sample set up.
Chapter 3 – EXPERIMENTAL PROCEDURE 40

1000 1000
900°C, 5'
900°C
900°C, 5'
800 50°C/s 800°C 800
850°C
750°C

Temperature (°C)
200°C/min 700°C
Temperature (°C)

650°C 200°C/min
600 600°C 600 600°C

Simultaneous forming 50°C/s


400 50°C/s 400 and quenching
Isothermal deformation at constant -1
strain rates = 0.07 - 1.0s
temperatures for different strain rates:

200 strain rates = 0.1, 1.0 amd 10.0s-1 200

0 0
0 100 200 300 400 500 600 700 800 0 100 200 300 400 500 600 700 800
Time (s) Time (s)

a b
Figure 3-4- Schematic illustration of a) isothermal deformation, and
b) simultaneous forming and quenching experiments in the current work.

The experimental procedures were as follows: inserting the cylindrical


specimen (as shown in figure 3-5) in a vacuum chamber, resistance heating to
austenization temperature and performing the subsequent compression between
SiN2 anvils prior to controlled cooling. Molybdenum foils were used to prevent
the specimens to be pasted to the anvils, and the glass powder was utilized for
lubrication. The Pt/Pt-Rh10% thermocouple was welded to the specimen in
order to measure the temperature. The atmosphere was initially protected by
vacuum and then argon and helium shower were employed for a controlled
cooling.
0,3mm±0,1mm
10,0mm±0,1mm

4,0mm±0,1mm 0,3mm±0,1mm
5,0mm±0,1mm

Figure 3-5- Schematic illustration of the cylindrical specimen used during the
dilatation experiments.
Chapter 3 – EXPERIMENTAL PROCEDURE 41

3.3. Hardness and Metallography Tests


The deformed specimens in the dilatometry machine, were then cut off and sent
for the metallography and hardness (HV10 scale) measurements.
The microstructural images besides the image analysis data were extracted and
compared with the resulted hardness and mechanical data (dilatation tests).
Chapter 4 – RESULTS AND DISCUSSION 41

CHAPTER

FOUR
RESULTS AND DISCUSSION

4.1. High Temperature Isothermal Compression Tests


In this part of the chapter, the results of the high temperature isothermal
compression tests on the 22MnB5 steel (BTR165) are presented and analyzed.
The main topic has been divided into three sections: hardness and
microstructure, deformation, and dilatation data analysis.

4.1.1. The Hardness Values and Microstructural Variations


Firstly, the variations of the final hardness data (as quenched) by taking the
strain rates as the constant quantities were examined among different
isothermal deformation temperatures from 500°C to 900°C. As the second step,
the changes of the final hardness value were investigated based on the constant
deformation temperature by increasing the strain rates from 0.1 to 10.0s-1 at two
selected isothermal deformation temperatures, i.e. 750°C and 900°C. The
relevant microstructural images are presented for comparison.

4.1.1.1. Results
Figure 4-1, 4-2 and 4-3 show the variations of the final hardness values due to
the different deformation temperatures in three different constant strain rates.
Chapter 4 – RESULTS AND DISCUSSION 42

460 452 451


442
440

420
409
400
HV~

380
368
360

340 Isothermal Deformation Temperature Increase


329
Strain Rate = 0.1s-1
Hardness Variations
320

500 600 700 800 900


Deformation Temperature (°C)

Figure 4-1– Hardness variations by increasing the deformation temperature at the


strain rate of 0.1s-1.

600
Isothermal Deformation Temperature Increase
550 Strain Rate = 1s-1 543
Hardness Variations

500
469
450 475
HV~

400 413
405
352
350
320

300 325

250
500 600 700 800 900
Deformation Temperature (°C)

Figure 4-2– Hardness variations by increasing the deformation temperature at the


strain rate of 1.0s-1.
Chapter 4 – RESULTS AND DISCUSSION 43

500
472

450

417
421
400
402
HV~

350

300 310 Isothermal Deformation Temperature Increase


Strain Rate = 10s-1
Hardness Variations
250
600 700 800 900
Deformation Temperature (°C)

Figure 4-3– Hardness variations by increasing the deformation temperature at the


strain rate of 10.0s-1.

The diagrams demonstrate a continuous increasing trend at the strain rate of


0.1s-1 and the same trend with two decreasing sections at the lowest and the
highest deformation temperatures at the strain rate of 1.0s-1. At the strain rate of
10.0s-1, the increasing trend with some deviations from a sharp rise is observed.
A rather related effect, so-called “adiabatic heating”, at the strain rate of 10.0s-1
is presented in figure 4-4. As is seen, there is a sudden increase in the
temperature value (10-20 degrees), as the amount of exerted force is increased
by continuing the deformation process.
The variations of the hardness values regarding the increase of the strain rates
at the constant deformation temperatures of 750°C and 900°C are observed in
figures 4-5 and 4-6.
Chapter 4 – RESULTS AND DISCUSSION 44

Figure 4-4 – A sample Force-time-Temperature (F-t-T) diagram of the 22MnB5


specimens regarding the isothermal deformation at 750°C and the strain rate of 10.0s-1
demonstrating the adiabatic heating effect.

450
-1
445 Strain Rate Increase (0.1~10s )
Isothermal Deformation Temperature = 750°C
442
440 Hardness Variations

435

430
HV~

425

420

415 413

410

405
402
400
0.1 1 10
-1
Strain Rate (s )

Figure 4-5- Hardness variations by increasing the strain rate at the deformation
temperature of 750°C.
Chapter 4 – RESULTS AND DISCUSSION 45

530
521
520

510

500

490
HV~

480 475

470

460 -1
Strain Rate Increase (0.1~10s )
450 Isothermal Deformation Temperature = 900°C
451
Hardness Variations
440
0.1 1 10
-1
Strain Rate (s )

Figure 4-6- Hardness variations by increasing the strain rate at the deformation
temperature of 900°C.

a b

c
Figure 4-7- The final microstructure of the isothermally deformed samples at 750°C
by different strain rates; a) 0.1s-1, b) 1.0s-1 and c) 10.0s-1.
Chapter 4 – RESULTS AND DISCUSSION 46

a b

c d

e f

Figure 4-8- The final microstructure of the isothermally deformed samples by the
strain rate of 0.1s-1 at different temperatures; a) 500°C, b) 650°C, c) 700°C, d) 750°C,
e) 800°C and f) 900°C.

Increasing the isothermal deformation temperature by the strain rate of 0.1s-1


results in the increase of martensite phase percentage in the final
microstructures which are shown in figure 4-8.
Chapter 4 – RESULTS AND DISCUSSION 47

a b

Figure 4-9- The microstructure of the isothermally deformed samples by the strain
rate of 1.0s-1 at different temperatures; a) 550°C, b) 600°C.

Figure 4-9 gives the microstructure of the deformed samples at 550°C and
600°C by the strain rate of 1.0s-1. More ferrite could be observed in 600°C
sample than 550°C.

4.1.1.2. Discussion
Due to figure 4-1, a strain rate of 0.1s-1 leads into a residence time of five
seconds for the samples during the isothermal deformation before reaching the
final strain of 0.5 at the experimental temperature. Hence, this amount should
be added to the applied time for decreasing the temperature of the samples from
the austenization temperature (900°C) to the deformation temperature by the
cooling rate of 50°C/sec. Later, in section 4.1.2, it is seen that the mentioned
cooling rate prevents the samples from entering the isothermally formed ferrite
phase region during the experiments.
At 500°C, the microstructure mostly (more than 90%) contains the bainite
phase in addition to less evident areas of ferrite and martensite (figure 4-8a). It
could be assumed that during cooling, firstly the sample has entered the
continuously cooled ferrite phase region (see section 2.3.2.3). Consequently, by
further cooling to 500°C (from ferritic transformation temperature), and by
staying at this temperature and performing the deformation for five seconds,
the continuously cooled and isothermally formed bainitic transformations have
Chapter 4 – RESULTS AND DISCUSSION 48

been started and accelerated (this complies with [15]). The final cooling below
the Ms temperature transforms the remained austenite into the martensite.
Therefore, the reported hardness value in the current sample is due to the effect
of the bainite phase instead of martensite phase in the microstructure.
Increasing the deformation temperature to 650°C results in the deformation
temperature to come out of the continuously cooled bainitic and ferritic
transformation regions; hence the sample is isothermally deformed in the
austenite phase region. By performing the isothermal deformation of austenite,
the nucleation sites of the continuously cooled bainite and ferrite phases are
increased. The jump of the ferrite phase percentage in the microstructure of
650°C sample in comparison with 500°C sample could be described based on
the probable more sensitivity of ferrite forming mechanisms to have more prior
nucleation sites produced by prior deformation (figure 4-8).
By the end of the deformation process and by further cooling, the austenite is
partially transformed to continuously cooled bainite and ferrite phases, while
more amount of it is transformed to martensite. This is because the residence
time is not enough for the isothermal bainitic transformation to occur. As a
result, the hardness value is increased by the increase of the martensite phase
percentage.
By increasing the deformation temperature, the ferritic transformation intensity
is gradually decreased; therefore besides the stability of the bainite phase
percentage, the amount of the final martensite is increased (figure 4-8). This is
because of the matter that the effect of prior deformation on generating the
nucleation sites of ferritic transformation is diminished as the deformation
temperature is increased. This fact can be described by the evaluation of the
700°C and 750°C samples. The 750°C sample mostly consists of bainite and
martensite phases and a very less amount of ferrite (figure 4-8 'c' and 'd'). The
light colored areas in the microstructure were determined as the bainite phase
due to the presence of the distributed carbides inside them.
By increasing the temperature to 800°C and 900°C, the final microstructure is
fully martensitic (figure 4-8 'e' and 'f'). This shows that the nose of the bainitic
Chapter 4 – RESULTS AND DISCUSSION 49

and ferritic transformations in the CCT diagram has no coincidence with the
cooling line during the cooling process, although there is a deformation
residence time in the middle of the continuous cooling.
In figure 4-2, it takes seven seconds for the sample to reach the experimental
temperature of 550°C by the cooling rate of 50°C/sec. from the austenization
temperature of 900°C. It should be noted that the specimen enters the bainite
phase region during cooling before the deformation to be applied. The
combination of the time used for the cooling of the sample in the bainitic phase
region (which is around three seconds) and the residence time of the sample
during the deformation, i.e. 0.5 seconds (for the deformation up to the strain of
0.5 by the strain rate of 1.0s-1), leads into the bainitic transformation to be
developed to some extents. It means that the resulted bainite is produced mostly
by staying the sample at the bainitic phase region during cooling (continuously
cooled bainite) than the acceleration of the transformation by the isothermal
deformation (isothermally formed bainite). Looking into the microstructure,
figure 4-9a, more than 90% martensite is observed. This means that despite the
presence of two bainite forming mechanisms (isothermally formed and
continuously cooled), the bainitic transformation is not developed under current
experimental conditions. This shows that the lower deformation temperatures
may lead into the formation of more martensite by hindering the bainitic and
ferritic transformations.
Increasing the deformation temperature to 600°C, and coming out of the
bainitic transformation region, continuously cooled ferrite formation occurs
(due to the observed microstructural results in figure 4-9b), and by this, the
final hardness value of the sample is decreased. At the same time, the stay of
the sample in the bainitic phase region during cooling leads into the formation
of the continuously cooled bainite phase.
By further increase of the deformation temperature to 750°C, two factors have
a competitive effect on the hardness values. Firstly, the deformation of the
sample out of the continuously cooled bainitic and ferritic phase regions which
results in the presence of more austenite in the microstructure of the
Chapter 4 – RESULTS AND DISCUSSION 50

experimental sample before the martensitic transformation temperature, which


means that the percentage of the martensite in the final structure and
consequently the hardness value is increased. Secondly, the formation
probability of the continuously cooled ferrite phase because of the isothermal
deformation, prior to the continuously cooled ferritic transformation
temperature, which results in the sample to enter a broad area of ferrite phase,
because of making more nucleation sites.
The dominant effect of the increase of the amount of martensite (as a harder
phase than bainite and ferrite) leads into the gradual increase of the final
hardness values of the samples.
By jump of the deformation temperature to the amounts more than 850°C to
900°C, the effect of the deformation on the stabilization of the austenite phase
prevents the sample to enter the ferritic and bainitic phase regions during the
cooling of the samples after the isothermal deformation. Therefore, in the 850
and 900°C samples, almost 100% martensite has been reported. The presence
of bainite phase in the 850°C sample could be justified due to the coincidence
of the cooling line and the bainitic transformation nose in the CCT diagram.
The decrease of the hardness value in the 900°C sample in comparison with
850°C (despite the presence of almost 100% martensite in both cases) could be
related to the difference of the martensitic structure in these two samples. This
is because in 850°C we have a 50°C decrease of the deformation temperature
from the austenization temperature (900°C), while in the case of deformation at
900°C, the deformation and the austenization temperatures are the same. It
could be imagined that the effect of the applied stress on the final martensitic
structure is diminished by a very high deformation temperature of 900°C in
comparison with 850°C. There are, however, other mechanisms such as
recovery and recrystallisation which may describe the same consequences.
In the case of isothermal deformation by the strain rate of 10.0s-1 (figure 4-3),
increasing the deformation temperature results in the increase of the hardness
value (because of an increase in the martensite content) by changing the place
where the sample enters the continuously cooled ferritic and bainitic phase
Chapter 4 – RESULTS AND DISCUSSION 51

regions. This is the same as what is seen regarding the deformation by the
strain rates of 0.1 and 1.0s-1 (figures 4-1 and 4-2).
The related fluctuations in the increasing trend of the hardness diagram at the
strain rate of 10.0s-1 could be justified due to the presence of the temperature
increase (10-20°C) by means of the adiabatic heating1 (see figure 4-4) which
could change the location of the specimen in the phase regions and change the
phase percentage at higher deformation rates.
It takes three seconds for all the specimens to come from the austenization
temperature of 900°C to the deformation temperature of 750°C by the cooling
rate of 50°C/sec. Then there are 5, 0.5 and 0.05 - second periods of time for the
0.1, 1.0 and 10.0s-1 specimens respectively to be deformed isothermally at the
deformation temperature. Therefore, due to the mentioned residence times of
the samples, there is a more probability for the 0.1s-1 specimen to enter a
broader area of continuously cooled ferritic and bainitic phase regions than 1.0
and 10.0s-1 ones. The final microstructures show that the 0.1s-1 specimen is
heavily ferritic and bainitic and the remained parts have been transformed to
martensite (figure 4-7a). The reported hardness value (figure 4-5) is due to the
presence of more bainite beside the martensite in this specimen. In the 1.0 and
10.0s-1 specimens, the residence time at the deformation temperature is almost
the same (there is a 0.45 seconds difference between them). The more
important difference is a 10°C higher temperature in the case of 10.0s-1
specimen (because of the mentioned adiabatic heating) which in addition to a
shorter residence time (by 0.45 seconds) leads into the decrease of the bainite
and ferrite phases and the increase of the final martensite in the microstructure
of the isothermally deformed specimens.
The colorful role of the bainite in the hardness value of the 22MnB5 steel
results in the decrease of the hardness value by decreasing the bainite phase
percentage in the microstructure by increasing the strain rate from 0.1 to 10.0s-1
at the deformation temperature of 750°C.
1
Heat is generated by the plastic deformation at high strain rates. The heat generated in the material is
either conducted and/or convected away to the surrounding or is used to increase the temperature of
the material. When the heat generation rate is greater than the rate of heat loss, the temperature of
the material is increased. Indeed, the heat dissipation is time dependant and therefore, in the low
speed like in quasi-static processes, the heat can be transferred to the surrounding, dissipated and
deformation occurs isothermally. During high-speed or fast enough processes, in the materials whose
flow curves are temperature dependant, the flow stress is lowered simultaneously by the continuous
rise of temperature due to adiabatic heating.
Chapter 4 – RESULTS AND DISCUSSION 52

Regarding the deformed samples at 900°C (figure 4-6), all the specimens have
a rather full martensitic structure and the reported increase in the hardness data
could be justified due to the differences in the martensitic structure under
different deformation conditions and the presence of small amounts of bainite
phase which its structure and its arrangement beside a full martensitic matrix
may affect the final hardness values in different ways.
The hardness variation diagrams of the isothermally deformed specimens at
650°C and 700°C exhibit exactly the same behavior as for the 750°C specimen.
In addition, there is a same trend for the 800°C and 850°C specimens in
comparison with the 900°C specimen. Therefore, the relevant data is not
presented here, due to preventing further complications to occur. Based on
these trends, it would be possible to divide the whole experiments into two
high-deformation-temperature and low-deformation-temperature categories and
predict the response of the experimental material to the current heating and
deformation process regarding the temperature range at which the deformation
is done. At the lower deformation temperatures, the effect of any continuously
cooled bainite and ferrite phases has to be taken into account, while after the
high temperature isothermal deformation, the specimen does not enter the
mentioned transformation regions during cooling.
Chapter 4 – RESULTS AND DISCUSSION 53

4.1.2. The Deformation Data Analysis


In this section, the whole resulted mechanical data from the dilatation
experiments, i.e. yield stresses, flow stresses, and work hardening rates are
presented and compared based on the variations of the experimental
temperatures and strain rates.

4.1.2.1. Results
Figure 4-10 shows the flow curves of the deformed samples at different
temperatures from 550°C to 900°C. It is seen that the stress level is
continuously decreased by increasing the deformation temperature.

600

500
550°C
600°C
True Stress, σ [MPa]

400 650°C
700°C
750°C
300 800°C
850°C
900°C
200

100
0.0 0.1 0.2 0.3 0.4 0.5 0.6
True Plastic Strain, ε [-]

Figure 4-10- True stress – true plastic strain curves of the deformed samples based on
different deformation temperatures at the strain rate of 1.0s-1.

The flow curves of the deformed samples regarding different strain rates are
presented in figure 4-11. Coming from the highest to the lowest strain rates,
there is a drop in the value of the flow stress at 800°C.
Chapter 4 – RESULTS AND DISCUSSION 54

350
-1
Strain Rate [s ]
10.0
300 1.0

True Stress, σ [MPa]


250 0.1

200

150

100

50

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
True Plastic Strain, ε [-]

Figure 4-11- True stress – true plastic strain curves of the deformed samples by
different strain rates (0.1, 1.0, and 10.0s-1) at 800°C.

Figures 4-12 gives the yield stress data due to the variations of the strain rate at
the constant selected deformation temperatures. The same data based on the
variations of the deformation temperature at the constant selected strain rates
are shown in figures 4-13 and 4-14. By increasing the strain rate, it is seen that
the yield stress value is increased, while by increasing the deformation
temperature, the mentioned value is decreased.
Chapter 4 – RESULTS AND DISCUSSION 55

300
280 289

260
240 Deformation Temperature = 750°C

Yield Stress (MPa)


227
220
200
Deformation Temperature = 900°C 177
180
169
160
-1
Strain Rate Increase (0.1~10s )
140 136
Isothermal Deformation Temperature = 750°C

120 120 Yield Stress Variations

0.1 1 -1 10
Strain Rate (s )

Figure 4-12– Yield stress variations by increasing the strain rate at the deformation
temperature of 750°C and 900°C.

350

332
300

250
Yield Stress (MPa)

218
194
200
169
163
150
120

100

Isothermal Deformation Temperature Increase


50 Strain Rate = 0.1s
-1

Yield Stress Variations


0
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-13–Yield stress variations by increasing the deformation temperature at the


strain rate of 0.1s-1.
Chapter 4 – RESULTS AND DISCUSSION 56

350

300 288
300

250 232

Yield Stress (MPa)


244
227
200
166

150 136
156

100

Isothermal Deformation Temperature Increase


50 Strain Rate = 1s
-1

Yield Stress Variations


0
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-14–Yield stress variations by increasing the deformation temperature at the


strain rate of 1.0s-1.
1800
Work Hardening Rate (θ) Vs. Strain
1600 Strain Rate = 1.0 s-1
Total Strain = 0.5
1400
Work Hardening Rate, θ [Mpa]

Deformation Start Temperature Variations


1200 550°C

1000 600°C

650°C
800
700°C
600 800°C
750°C 850°C
400 900°C

200

0
0.00 0.10 0.20 0.30 0.40 0.50 0.60

True Strain, ε [-]

Figure 4-15– Work hardening rate variations versus true strain based on different
deformation temperatures; strain rate = 1.0s-1; total strain = 0.5.
Chapter 4 – RESULTS AND DISCUSSION 57

The work hardening rate data based on the variations of the deformation
temperature and the variations of the strain rate are given in figures 4-15 and
4-16. The lowering effect of the deformation temperature increase on the work
hardening rate value is evident, while the mentioned value is not permanently
increased by increasing the strain rate.

1200
Work Hardening Rate (θ) Vs. Strain
Deformation Start Temperature = 800°C
1000 Total Strain = 0.5
Strain Rate = 1.0 s-1
Work Hardening Rate, θ [Mpa]

Strain Rate Variations

800

Strain Rate = 10.0 s-1


600

Strain Rate = 0.1 s-1


400

200

0
0 0.1 0.2 0.3 0.4 0.5 0.6
True Strain, ε [-]

Figure 4-16– Work hardening rate variations versus true strain based on different
strain rates; deformation temperature = 800°C; total strain = 0.5.

4.1.2.2. Discussion
In figure 4-12, the increase of the strain rate directly results in the increase of
the yield stress. This is because by increasing the strain rate from 0.1 to 10.0s-1,
a greater number of dislocations are produced at the beginning of the
deformation process. The sudden accumulation and hit of the mentioned
dislocations leads into a higher required stress level to enter the plastic state of
deformation. Based on the same note in section 4.1.1.2, the variations of the
yield stress by increasing the strain rate from 0.1 to 10.0s-1 at all the
deformation temperatures from 600°C to 900°C are the same. Therefore, here
Chapter 4 – RESULTS AND DISCUSSION 58

only the 750°C and 900°C specimens were chosen as the group leaders of the
low- and high-deformation-temperature experiments.
Figures 4-13 and 4-14 show that rising the isothermal deformation temperature
leads into the decrease of the yield stress. This is because of more softness of
the structure and the activation of the temperature-dependent dislocation
motion mechanisms such as climb and cross slip.
By the evaluation of the work hardening rate diagrams, it is found that
increasing the isothermal deformation temperature leads into a drop in the work
hardening rate curve. The general observed trend of these diagrams can also be
found through the consideration of the flow curves behavior in figure 4-10. The
flow and work hardening rate curves clarify that the increase of the
deformation temperature ruins the work hardening phenomenon by the
activation of the dynamic recovery process which can also be found through
the smooth flow curves (figure 4-10) and the work hardening rate value which
reaches the zero state in the case of strains of more than 0.2-0.3.
Comparing the flow curves of the isothermally deformed samples at different
temperatures (figure 4-10), it is concluded that the increase of the deformation
temperature also evidently results in the drop of the stress level at the
beginning, during and the end of the plastic deformation. This is because the
dislocation motion as the most important factor for the deformation of the
metals becomes much easier as the temperature reaches the higher values.
At 750 and 800°C, the work hardening rate curves do not follow the overall
decreasing trend by increasing the temperature and this is mostly because of the
data scattering than an understandable metallurgical event due to the
temperature increase.
The competitive effect of temperature and exerted strain on the work hardening
rate behavior could be observed comparing figures 4-11 and 4-16. A higher
work hardening rate is expected for a higher strain rate at higher temperatures,
while the work hardening itself faces a drop as the temperature increases.
Simultaneously, a higher strain rate produces more dislocations in a shorter
period of time and forces them to hit each other. This leads in a higher work
Chapter 4 – RESULTS AND DISCUSSION 59

hardening rate at the beginning of the deformation and the disappearance of


this behavior by further deformation1.
As the deformation is continued, more dislocations hit each other and by this,
cause the work hardening rate to be deeply decreased for the strains more than
0.2 and finally reach a zero value for the strains more than 0.4. Considering the
deformed sample by the strain rate of 1.0s-1 (see figures 4-11 and 4-16), it is
found that the mentioned specimen is in a transition condition between the
effect of the higher temperature on more deformed samples (i.e. by higher
strain rates) to make their work hardening rate to be increased, and the effect of
a higher strain rate on decreasing the work hardening rate level as the
deformation is continued (i.e. because of the dislocation accumulation which
prevents the work hardening rate to be increased). Therefore, in spite of a logic
trend in the flow curves behavior (i.e. the higher the strain rate, the higher the
flow stress level), the highest work hardening rate is observed in the 1.0s-1
specimen, although all of the work hardening rate curves reach a zero value due
to the dynamic recovery phenomenon at the strains of more than 0.4.
The point to be considered is that the increasing effect of the dynamic recovery
is completely ruined when the dynamic recovery rate overcomes the rate of
dislocation generation by further deformation. At such moment, almost no
dislocation is remained in the microstructure and all of the newly born
dislocations will be swept away by the dynamic recovery process. Therefore,
no change in the amount of work hardening is observed and finally the work
hardening rate will be zero.

1
Due to [33] and during the initial stages of deformation, there is an increase in the flow stress as
dislocations interact and multiply. However, as the dislocation density rises, so the driving force and
hence the rate of recovery increases and during this period, a microstructure of low angle boundaries
and subgrains develops. At a certain strain, the rates of work hardening and recovery reach a dynamic
equilibrium, the dislocation density remains constant and a steady-state flow stress is obtained. During
deformation at strain rates larger than ~1.0s-1 the heat generated by the work of deformation cannot all
be removed from the specimen and the temperature of the specimen rises during the deformation. This
may then cause a reduction in the flow stress as straining proceeds. In modeling the high temperature
deformation behavior, it is very important that such effects are taken into account as mentioned in
4.1.1.2.
Chapter 4 – RESULTS AND DISCUSSION 60

4.1.3. The Dilatation Data Analysis


In this section, the variations of the martensitic transformation start temperature
(Ms), the changes of the temperature range in which the martensitic
transformation occurs (i.e. Ms-Mf), and finally the dilatation curves of selected
experiments are presented and compared.

4.1.3.1. Results
The variations of the Ms and (Ms-Mf) data at the constant strain rates based on
different deformation temperatures and the variations of the microstructural
states are given in figures 4-17 to 4-22. Both Ms and (Ms-Mf) values almost
demonstrate an increasing trend due to the increase of the deformation
temperature.

450

400
378 378
378 378

378
350 378
Ms(°C)

300

250 Isothermal Deformation Temperature Increase


-1
Strain Rate = 0.1s
Ms Variations
200
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-17- Ms variations regarding the deformation temperature at the strain rate of
0.1s-1.
Chapter 4 – RESULTS AND DISCUSSION 61

450
414

400 393
384 385 408
390

367
350 357
Ms(°C)

300

250 Isothermal Deformation Temperature Increase


-1
Strain Rate = 1.0s
Ms Variations
200
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-18- Ms variations regarding the deformation temperature at the strain rate of
1.0s-1.

450

407
398
400
393
380

350 363
Ms(°C)

300

250 Isothermal Deformation Temperature Increase


-1
Strain Rate = 10.0s
Ms Variations
200
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-19- Ms variations regarding the deformation temperature at the strain rate of
10.0s-1.
Chapter 4 – RESULTS AND DISCUSSION 62

Table 4-1- The variations of the phase percentage in the microstructure of the
22MnB5 isothermally deformed specimens regarding different deformation
temperatures and strain rates (M=martensite, B=bainite, F=ferrite). The phase
percentage in the blank fields has not been reported.
Def.
Temp.
/ 500°C 650°C 700°C 750°C 800°C 900°C
Strain
Rate
60%M 66%M 76%M 96%M 100%M
0.1s-1 100%B 27%B 4%B
13%F 34%F 24%F
46%M 82%M 81%M 86%M 98%M
1.0s-1 43%B 14%B 2% B/F
11%F 18%F 19%F
30%M 96%M
10.0s-1 63%B
7%F 4%F

250
Isothermal Deformation Temperature Increase
-1
Strain Rate = 0.1s
200 Ms -Mf Variations 180

155
150
Ms-Mf(°C)

142
111

100

75 79

50

0
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-20- Ms-Mf variations regarding the deformation temperature at the strain rate
of 0.1s-1.
Chapter 4 – RESULTS AND DISCUSSION 63

250

210 212

200 186

150
Ms-Mf(°C) 128 148

125
118
100 112

Isothermal Deformation Temperature Increase


50
-1
Strain Rate = 1.0s
Ms -Mf Variations
0
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-21- Ms-Mf variations regarding the deformation temperature at the strain rate
of 1.0s-1.

250

208

200
193
157

150
Ms-Mf(°C)

143

121
100

50 Isothermal Deformation Temperature Increase


-1
Strain Rate = 10.0s
Ms -Mf Variations
0
400 500 600 700 800 900
Deformation Temperature (°C)

Figure 4-22- Ms-Mf variations regarding the deformation temperature at the strain rate
of 10.0s-1.

The dilatation curves of the experimental samples regarding three different


deformation temperatures are given in figures 4-23, 4-24 and 4-25 due to the
increase of the strain rates. As the strain rate is increased, the martensitic
transformation valleys in the dilatation curves are shifted to the right hand side
Chapter 4 – RESULTS AND DISCUSSION 64

of the diagram. This behavior is seen in all three different experimental


temperatures.

0.0
-1
Strain Rate = 0.05s
Change in Length (%)

-0.1

-0.2 -1
Strain Rate = 10.0s

-1
-0.3 Strain Rate = 1.0s

-1
Strain Rate = 0.1s

-0.4

0 100 200 300 400 500 600 700


Temperature (°C)

Figure 4-23- Dilatation curves of the 22MnB5 steel by increasing the strain rate from
0.05 to 10.0s-1; austenization temperature = 900°C;
deformation temperature = 650°C.

0.0

-0.1
-1
Strain Rate = 1.0s
Change in Length (%)

-1
-0.2 Strain Rate = 0.1s

-0.3

-0.4
-1
Strain Rate = 10.0s

-0.5
0 100 200 300 400 500 600 700
Temperature (°C)

Figure 4-24– Dilatation curves of the 22MnB5 steel by increasing the strain rate from
0.1 to 10.0s-1; austenization temperature = 900°C;
deformation temperature = 800°C.
Chapter 4 – RESULTS AND DISCUSSION 65

0.0

-0.1 -1
Strain Rate = 10.0s

Change in Length (%)


-1
Strain Rate = 1.0s
-0.2

-0.3

-0.4
-1
Strain Rate = 0.1s

-0.5
0 100 200 300 400 500 600 700
Temperature (°C)

Figure 4-25– Dilatation curves of the 22MnB5 steel by increasing the strain rate from
0.1 to 10.0s-1; austenization temperature = 900°C;
deformation temperature = 900°C.

4.1.3.2. Discussion
By increasing the deformation temperature, Ms is decreased at all the strain
rates. This phenomenon is in contrast with the increasing effect of the
austenization time and temperature on the Ms value (see figures 2-10 and
2-11 in chapter 2). This could be justified due to the fact that, the austenite
matrix uses the higher temperature as an accelerating factor besides the exerted
strain (driving force for the new grains to be born) to make a finer austenite
grain structure after the isothermal deformation and before quenching. As
higher austenization time and temperature means higher austenite grain size (in
comparison with smaller grains due to the isothermal deformation at higher
temperatures), decreasing the grain size leads into the decrease of the Ms value
as described before.
The amount of martensite in the final microstructures is also increased at the
higher deformation temperatures (table 4-1). It could be justified regarding the
fact that, continuously cooled bainitic and ferritic transformations are hindered
with higher deformation temperatures as described before. As this is the only
Chapter 4 – RESULTS AND DISCUSSION 66

way of consuming the austenitic matrix, more austenite remains till the
beginning of the martensitic transformation. Therefore, the final martensite
content is increased by isothermal deformation at higher temperatures.
Considering table 4-1 and figures 4-20 and 4-21, it is found that, despite the
value of Ms-Mf somehow gives an overview of the martensite phase percentage
in the final microstructure, but the occurrence of the bainite and ferrite phases
and their amounts may ruin the correlation between these two quantities. In
most cases (but not all cases), it is seen that the presence of bainite and ferrite
phases in the microstructure can decrease and increase the value of Ms-Mf
respectively.
It is seen through comparing figures 4-20, 4-21 and 4-22 that the overall trend
of three diagrams is increasing and the strain rate of 10.0s-1 shows a rather
higher level than two others. It means that by increasing the isothermal
deformation temperature, the amount of martensite and the value of Ms-Mf are
increased. Also the higher level of figure 4-22 (strain rate of 10.0s-1) shows that
the increase of the strain rate can increase the time required to produce a certain
amount of martensite.
In figures 4-23, 4-24 and 4-25, it can be found that by increasing the strain rate
at three deformation temperatures (650, 800 and 900°C), the martensitic
transformation regions of the dilatometry curves are shifted to the higher
temperatures. This fact beside what is seen in figures 4-20 to 4-22, is because
the residual stresses as a result of the deformation of the sample in the
austenitic phase region (the long range elastic stresses) leads into the increase
of the martensitic transformation start temperature (Ms) and the effect is
increased as the amount of deformation is increased [28].
Chapter 4 – RESULTS AND DISCUSSION 67

4.2. Simultaneous Deformation and Quenching Tests


In the second part of the current chapter, the results of the simultaneous
deformation and quenching tests will be discussed in detail. This part has been
divided into four sections including: deformation duration effect, strain rate
effect, austenization soaking time effect, and deformation start temperature
effect.

4.2.1. The Strain Magnitudes Effect (Deformation Duration Effect)


The strain rate and the start temperature of the experiments were fixed at 0.1s-1
and 800°C respectively and the durations of the deformation tests were set to
be 1, 2, 3, 4 and 5 seconds (with respect to the total strain values of 0.1, 0.2,
0.3, 0.4 and 0.5). As the deformation is started at 800°C, the sample is cooled
down by the cooling rate of 50°C/sec. The cooling process is continued till the
sample reaches the room temperature. Therefore, the compression tests are
performed simultaneously during the cooling process.

4.2.1.1. Results
Figure 4-26 shows the various dilatation curves of the different deformation
periods from one to five seconds at the deformation start temperature of 800°C
and the strain rate of 0.1s-1.
Chapter 4 – RESULTS AND DISCUSSION 68

0.0
Deformation Time = 5 sec.
-0.1 Deformation Time = 4 sec.

-0.2

Change in Length (%)


-0.3

-0.4

-0.5
Deformation Time = 3 sec.

-0.6 Deformation Time = 2 sec.


Deformation Time = 1 sec.
-0.7

-0.8
100 200 300 400 500 600 700
Temperature (°C)

Figure 4-26- Dilatation curves of the 22MnB5 steel; deformation start


temperature = 800°C; strain rate = 0.1s-1; deformation duration = 1-5 seconds.

As can be seen in this figure, increasing the amount of the deformation


(basically shown by increasing the duration of the deformation process from
one to five seconds), slightly retards the picks and shifts the transformation
sections of the curves to the lower temperatures. Furthermore, the depths of the
valleys in these curves are substantially decreased by increasing the
deformation time from one to five seconds.
The Ms and Mf values were found, and the extracted Ms and (Ms-Mf) data were
compared in figures 4-27 and 4-28 respectively.
Chapter 4 – RESULTS AND DISCUSSION 69

400

390
382
380
374
378
370
364
361
Ms(°C)

360

350

Strain Value Increase


340 Strain Rate = 0.1s
-1

Deformation Start Temperature = 800°C


330 Ms Variations

320
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Final Strain, ε

Figure 4-27- Ms variations regarding the final strain value.

180

160
159 136
140

120
105
100
Ms-Mf (°C)

100
80 87

60
Strain Value Increase
-1
40 Strain Rate = 0.1s
Deformation Start Temperature = 800°C
Ms-Mf Variations
20

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Final Strain, ε

Figure 4-28- Ms-Mf variations regarding the final strain value.

It can be easily seen that the (Ms-Mf) values are decreased by increasing the
amount of the strain from 0.1 to 0.3 (figure 4-28) followed by an increase for
the strains above 0.3, while initially there is a decreasing behavior for the Ms
values (figure 4-27) followed by a rising trend due to the strains higher than
0.3.
Chapter 4 – RESULTS AND DISCUSSION 70

Figure 4-29 shows how the amount of the martensite in the as deformed sample
varies with respect to the exerted strain. The graph shows a drop in the amount
of the martensite in the matrix which is followed by an increase for the strains
larger than 0.3.

Strain Value Increase


100 Strain Rate = 0.1s
-1

98 Deformation Start Temperature = 800°C


Amount of Martensite (%)

80
Amount of Martensite (%)

60
58

40 45

20
15

0 ~0

0.1 0.2 0.3 0.4 0.5


Final Strain, ε

Figure 4-29- The variations in the amount of martensite in percentage regarding the
final strain value.

The effect of an increase in the amount of the applied strain on the flow curve
behavior of the 22MnB5 samples are shown in figure 4-30.
Chapter 4 – RESULTS AND DISCUSSION 71

500
450 Final Strain = 0.5
Deformation Time = 5 sec.
400 Temperature at the End of
Deformation = 550°C
350

True Stress, σ (MPa)


Final Strain = 0.4
Deformation Time = 4 sec.
300
Temperature at the End of

250 Deformation = 600°C

200
Final Strain = 0.3

150 Final Strain = 0.2 Deformation Time = 3 sec.

Final Strain = 0.1 Deformation Time = 2 sec. Temperature at the End of


100 Deformation Time = 2 sec. Deformation = 650°C
Temperature at the End of

50 Temperature at the End of Deformation = 700°C


Deformation = 750°C
0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
True Plastic Strain, ε [−]

Figure 4-30– Flow curves variations by increasing the applied strain from 0.1 to 0.5;
strain rate = 0.1s-1; deformation start temperature = 800°C.

Going from the shortest to the longest deformation duration, there is a


continuous increasing trend for the flow curves behavior.
A sample time-temperature-Force (t-T-F) diagram of a specimen which was
deformed by the strain rate of 0.1s-1 and the start temperature of 800°C for two
seconds is shown in figure 4-31. As is seen, the temperature curve starts from
800°C and ends at 700°C (at the cooling rate of 50°C/sec. for two seconds),
while the specimen simultaneously starts to receive the load at 800°C and is
unloaded at 700°C.
Chapter 4 – RESULTS AND DISCUSSION 72

8000 820

7000 Force
800

6000
780
Temperature

5000

Temperature (°C)
760
Force (N)

4000

740
3000

720
2000

700
1000

0 680
571.5 572.0 572.5 573.0 573.5 574.0 574.5 575.0
Time (sec.)

Figure 4-31- A sample t-T-F diagram of the experimental material;


Strain rate = 0.1 s-1; Start temperature = 800°C,
Duration of the test = 2 seconds.

Figure 4-32 shows the microstructures of the experimental specimens after the
deformation process regarding their different deformation times from one to
five seconds.
Chapter 4 – RESULTS AND DISCUSSION 73

c
Figure 4-32- Microstructure of the 22MnB5 samples after the simultaneous
deformation and quenching process; deformation start temperature = 800°C,
strain rate = 0.1 s-1; a) deformation time = 1 sec., b) deformation time = 3 sec.,
c) deformation time = 5 sec.
Chapter 4 – RESULTS AND DISCUSSION 74

4.2.1.2. Discussion
When austenite is plastically deformed, residual stresses and lattice defects are
introduced. The residual stresses are principally long-range elastic stresses that
raise the start temperature Ms and lower As temperature. This effect increases
with increasing the magnitude of deformation and reaches a saturation value.
On the other hand lattice defects (i.e. short-range stresses) lower the martensite
finish temperature Mf and raise the Af temperature. In contrast, at temperatures
as high as 525°C, internal stresses are relieved; hence, the factors raising the Ms
will become lessened. However, there will still remain lattice defects that are
not annihilated by heating to 525°C. Some such defects can accelerate the
transformation below the Ms. This is why the amount of the martensite is
increased at a low degree of prior deformation. At higher amounts of
deformation, lattice defects which stabilize austenite are formed (carbon atoms
migrate to them) and lower the Ms and decrease the amount of martensite; that
is, the austenite is considerably stabilized. Finally, for more than 30%
deformation, only a further fluctuation of the carbon concentration occurs.
Consequently Ms is raised, the austenite becomes unstable and the amount of
martensite increases [28]. These facts are confirmed by the evaluation of
figures 4-27, 4-28 and 4-29.
Comparing the resulted data from figures 4-25 and 4-29, it is concluded that the
deep valley in the case of sample which was deformed for one second by the
strain rate of 0.1s-1 and the decrease of this depth by going to higher
deformation times (from one to three seconds), are due to the formation of the
higher amounts of martensite (98%) in the case of deformation for one second
and drop of this amount to almost 0% for the deformation up to three seconds.
The presence of the considerable martensite in the microstructure in spite of
less dilatation (i.e. dilatation curves of the deformed samples for four and five
seconds) has been discussed by Somani et al [31]. With respect to this work
and also to the microstructural evaluation of the current steel (figure 4-32), the
shallower depth of the valleys in the case of samples which were deformed for
four and five seconds can be justified regarding the presence of the other
Chapter 4 – RESULTS AND DISCUSSION 75

effective parameters such as residual stresses due to prior plastic straining and
prior probable strain-induced ferrite formation which are able to decrease the
amount of dilatation as the deformation time is increased from three to five
seconds.
Considering the microstructural images, and later by the analysis of the flow
curves of the deformed specimens, also the possibility of the strain induced
and/or continuously cooled ferrite formation (instead of normal static austenite
to ferrite transformation at high temperatures), it is seen that besides the
observed granular bainite microstructure (which is evident by the presence of
carbide free regions in figure 4-32), almost significant amount of ferrite is
determined regarding the deformed samples for five seconds. This confirms the
latter justification for the decrease of dilatation as the strain induced
transformation or other disturbing mechanisms occur.
The extracted (Ms-Mf) values which were presented in figure 4-28 indirectly
show the amount of the remained austenite which can be transformed to
martensite by cooling down the sample below Ms. Also this confirms the data
which were discussed in figure 4-29. The less opportunity to produce
martensite, e.g. by the presence of the bainitic and ferritic transformations
(which leads into the less amounts of Ms-Mf in figure 4-28), the less martensite
is formed (figure 4-29).
The severe deformation of austenite prior to its transformation hinders the
growth of martensite, causing a reduction in the fraction of the transformation
in spite of an increased number of nucleation sites' density. In this regard, it has
been shown that, deformation during the thermomechanical processing of steels
also accelerates the rate of bainite reaction [15]. It was also demonstrated by
Samoni et al that the plastic deformation above Ms leads to strain-induced
ferrite formation [31]. These mean by doing the mechanical work on the
samples, we are changing the shape of the CCT diagram of the experimental
alloy in such a way that the areas of the bainitic and ferritic transformations are
shifted to the left hand side of the diagram. This leads into the fact that by
increasing the deformation time, the only way to reach a full martensitic
Chapter 4 – RESULTS AND DISCUSSION 76

structure is by increasing the cooling rate to the amounts more than what were
examined here (50 °C/sec.) and probably out of the industrial possibility range
of cooling rate.
As is evident in figure 4-32, increasing the deformation time (or by other words
the amount of applied strain) results in an increase in the amounts of bainite
and strain induced ferrite phases in the microstructure, while a rather full
martensitic microstructure is achieved after the deformation process with the
lowest duration (the least amount of the exerted strain).
Figure 4-30 demonstrates that no phase transformation occurs during the
deformation for one to five seconds. This is revealed by a simple evaluation of
the slope of the flow curves that is not changed by further deformation. As the
final reported structures consist of bainite, ferrite and martensite, one can
conclude that whole transformations were taken place after the end of the
deformation process. However, the probability of ferrite formation prior to the
deformation must be taken into account.
Moreover, due to the simultaneous deformation and quenching nature of the
experiments, the most possible type of the produced bainite is 'granular'. This
fact was deeply introduced in chapter two. In each case, the observed ferrite
regions could be part of a carbide free granular bainite or an independent
continuously cooled or strain induced ferrite.
The difference between these types of ferritic microstructures is not clear at
least by means of optical microscopy during our experiments.
In figure 4-30, it is clear that, the bigger applied strains, the bigger work
hardening and more dislocation creation. Therefore, the flow curves were
monotonically increased.
Chapter 4 – RESULTS AND DISCUSSION 77

4.2.2. The Strain Rate Effect


Taking the deformation start temperature as a constant value (800°C), the
variations of the strain rate and its consequent effects on the properties of the
experimental material are studied in this section.

4.2.2.1. Results
Figure 4-33 shows the flow curves of the deformed samples at 800°C regarding
their different strain rates.

600

Strain Rate = 0.07 s -1


Temp. = 622°C
-1 Final Temp. = 500°C
(Strain Rate = 0.07 s )
500
Temp.= 675°C
(Strain Rate = 0.1s -1 ) Strain Rate = 0.1 s -1
Final Temp. = 600°C
400
True Stress,σ (MPa)

Strain Rate = 0.2 s -1


Final Temp. = 700°C

300 Strain Rate = 0.4 s -1


Final Temp. = 750°C

Temp. = 693°C
200 (Strain Rate = 0.07 s -1 )

Temp. = 763°C
(Strain Rate = 0.2 s -1 )
100
Temp. = 781°C
(Strain Rate = 0.4 s -1 )

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45
True Plastic Strain, ε [-]

Figure 4-33- True stress – true plastic strain curves of the deformed samples by
different strain rates at 800°C; 0.07, 0.1, 0.2 and 0.4s-1.

As is seen in this figure, the value of the maximum stress increases by


decreasing the strain rate from 0.4 to 0.07s-1. In addition, the shape of the flow
curves in the samples which were deformed by higher strain rates, i.e. 0.2 and
0.4s-1 show not such a rising behavior as for the lower strain rates (i.e. 0.07 and
0.1s-1).
Chapter 4 – RESULTS AND DISCUSSION 78

The variations of the maximum stress (the stress regarding the strain of 0.4)
and the hardness values have been demonstrated in figures 4-34 and 4-35.
There is a sharp drop in the amount of the maximum stress, while the as
quenched hardness value grows meaningfully when the strain rate increases.

Final Temperature = 500°C


500 Strain Rate Increase (0.07~0.4s-1)
500 Deformation Start Temperature = 800°C
Final Strain = 0.4
σ0.4 Variations
Maximum Stress, σ0.4 (MPa)

450

400 410 Final Temperature = 600°C

Final Temperature = 700°C


350
Final Temperature = 750°C

319
300 308

0.050.07 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45


-1
Strain Rate (s )

Figure 4-34- Maximum stress, σ0.4, variations by increasing the strain rate.

335 Final Temperature = 750°C


Strain Rate Increase (0.07~0.4s-1)
Deformation Start Temperature = 800°C 334
330
Final Strain = 0.4
Hardness Variations
325

320 Final Temperature = 700°C


HV~

318
315
311
310
Final Temperature = 600°C
305

300
297
295 Final Temperature = 500°C

0.07 0.1 0.2 0.4

-1
Strain Rate (s )

Figure 4-35- Hardness variations by increasing the strain rate.


Chapter 4 – RESULTS AND DISCUSSION 79

Figure 4-36, illustrates the microstructures of the deformed samples by the


strain rates of 0.07, 0.1, 0.2 and 0.4s-1. In all cases, considerable bainite and
martensite phases are observed.

a b

c d

Figure 4-36- Microstructures of the deformed samples by different strain rates;


a) 0.07s-1, b) 0.1s-1, c) 0.2s-1 and d) 0.4s-1. Start temperature = 800 °C,
final strain = 0.4.

Figure 4-37 shows the work hardening rate values of the samples deformed by
the strain rates of 0.07, 0.1, 0.2 and 0.4s-1 during the deformation process. After
the initial decreasing trend, all the graphs show a rather stable stage followed
by an increase at the end. The sample which was deformed by the strain rate of
0.07s-1 has the highest level among the others, while the 0.4 s-1 sample shows
the lowest level.
Chapter 4 – RESULTS AND DISCUSSION 80

3000
Work Hardening Rate (θ) Vs. Strain (ε)
Deformation Start Temperature = 800°C
2500 Total Strain = 0.4
Work Hardening Rate, θ [MPa]

2000 ~ 586°C

650°C
-1
~ 753°C Strain rate (s )
1500 ~ 733°C 725°C
0.07
~763°C

1000
0.1

500 0.4
777°C
~788°C 0.2
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
True Plastic Strain, ε [-]

Figure 4-37– Work hardening rate variations against true strain;


Start temperature = 800°C; Total strain = 0.4.

4.2.2.2. Discussion
The competitive role of the applied strain rate and the deformation temperature
affects the variations of the flow stress curves and the resulted work hardening
rate data presented in figures 4-33 and 4-37.
In the very initial stages of the flow curves in figure 4-33, the deformed
samples with the highest strain rates show the highest stress level while the
deformed samples with the lowest strain rates have the lowest levels. By
increasing the strain, this trend will be reversed. The mentioned behavior till
point A is normal, because a higher stress level is usually achieved with respect
to the higher strain rate.
The effect of the experimental temperature becomes more colorful as the
deformation continues. This is because by increasing the extent of the
deformation, the temperature distance among the deformed specimens by
different strain rates is increased. The higher the strain rate, the deformation is
finalized at a higher temperature. Due to this fact, the temperature level of the
Chapter 4 – RESULTS AND DISCUSSION 81

deformed specimens by higher strain rates is higher than the lower strain rate
ones at a same strain. Therefore, from point A onward, the effect of the
temperature on the flow stress is going to be dominant. This is evident by
examining the flow curve behavior of the 0.07s-1 specimen which is
substantially increased as coming to the higher strains due to the much lower
temperature among the other tests at a same strain. As coming to point B, the
effect of the strain rate on the flow stress level is totally disappeared and a pure
temperature dependent order of the flow curves is observed. The lower the
temperature (i.e. at lower strain rates), the maximum stress is increased (figure
4-34).
Figure 4-36 shows that by increasing the strain rate from 0.07 to 0.4s-1 more
bainite and less martensite are observed in the final microstructures of the
deformed specimens. Almost no ferrite is evident in these pictures. This means
that, a simultaneous deformation and quenching has changed the shape of the
CCT diagrams of the specimens in a way that the nose of the bainitic
transformation is sharply shifted to the left hand side of the diagram. By this,
the higher strain rates normally mean that the higher amounts of bainite and
less martensite are formed.
An important point to be considered by the evaluation of the flow curves and
the microstructural images (figures 4-33 and 4-36) is that no transformation
occurs during the deformation by the strain rates of 0.1 to 0.4s-1, because there
is no change in the slope of the flow curves among these specimens. In
contrast, it is seen that at point B, the slope of the flow curve of the 0.07s-1
specimen is increased and this means that a new phase with a higher strength
than the austenite phase, i.e. bainite, is formed. A simple comparison of the
work hardening rate and the flow stress diagrams of this specimen around point
B results in defining the bainitic transformation start temperature to be around
600°C for it.
It could be imagined that the bainitic transformation is started somewhere
between point B and the strain of 0.3 (see and compare figure 4-33 and 4-37).
In the meanwhile, a pure temperature dependent stress level is transformed to
Chapter 4 – RESULTS AND DISCUSSION 82

the combination of the lower temperature and the bainite phase effects on
increasing the flow stress level.
It is concluded that despite the role of temperature in determining the order of
the flow curves level at the strains of more than 0.2, the higher amounts of the
flow stress for the 0.07s-1 specimen can be emphasized by means of the bainitic
transformation effect.
Figure 4-35 compares the hardness values of the samples after quenching
regarding their different strain rates.
The criteria of the hardness values after quenching are as follow:
- The extent of the ferritic and bainitic transformations which consume
the austenite in the matrix and decreases the amount of the final
martensite in the structure;
- The extent of ferrite (soft phase) after quenching;
- The extent of bainite (rather hard phase) after quenching.
The increasing trend of the hardness values by increasing the strain rates in
spite of the decrease of the martensite content shows that the bainite phase has
a more powerful effect on changing the hardness values of the 22MnB5 steel in
these experiments than the martensite.
In figure 4-37, basically because of the constant cooling rate for all the
experiments (50°C/sec.) and the fact that the parameter which shows the time
effect is the strain rate, the deformed samples by the strain rates of 0.07 and
0.4s-1 show the lowest and the highest temperatures regarding an equal strain,
respectively.
Basically, it must be mentioned that the increase of the applied strain (i.e. by
continuing the deformation to the higher strains, or by a higher strain rate at a
same strain) increases the work hardening by the production of more
dislocations in the specimen. In contrast, the work hardening rate may be
decreased by means of further exerted strain. This is because as the material is
more deformed and work hardened, the changes of the work hardening value
are decreased. This is because the material is much stronger than before and in
Chapter 4 – RESULTS AND DISCUSSION 83

spite of the increase of the work hardening, the work hardening rate (i.e. the
change in the work hardening value) is decreased.
Furthermore, the increase of the deformation temperature makes the work
hardening rate to be decreased by ruining the effect of the exerted strain due to
the softening effects of the higher temperatures.
The increasing trend of the work hardening rate from the strain of 0.1 to the
strain of almost 0.2 (in the case of 0.07s-1 sample) is due to the increase of
strain in the sample which is completely in the austenite phase. Moreover,
because of a low strain rate despite a high temperature, the increase of the
exerted strain has no decreasing effect on the work hardening rate; therefore by
increasing the strain in this area, the work hardening rate is increased.
Comparing the work hardening rates of the specimens till the strain of 0.3 in
figure 4-37, shows that the deformed samples by higher strain rates have a
lower work hardening rate due to a higher instant temperature at a same strain.
Also, the jump of the slope of the work hardening rate in the case of 0.07s-1
specimen after the strain of 0.3 is directly related to the occurrence of bainitic
transformation.
Chapter 4 – RESULTS AND DISCUSSION 84

4.2.3. The Austenization Soaking Time Effect


In this section, the effects of three different austenization soaking times (i.e. 5,
10 and 15 minutes), on the consequent mechanical properties and
microstructural variations of the experimentally deformed 22MnB5 steel are
investigated.

4.2.3.1. Results
Figure 4-38 shows the flow curves of the deformed samples regarding their
different austenization soaking times. No meaningful deviation among the flow
curves is observed by increasing the soaking time from 5 to 15 minutes.

350
Austenization Time = 15 min.
300

250
True Stress, σ [MPa]

Austenization Time = 10 min.


200
Austenization Time = 5 min.

150

100 Austenization Temperature = 900°C


Deformation Start Temperature = 800°C
50 Strain Rate = 0.1 s
-1

Final Strain = 0.2


0
0.00 0.05 0.10 0.15 0.20 0.25
True Plastic Strain, ε [−]

Figure 4-38– Flow curves of the deformed samples regarding their different
austenization soaking times.

The dilatation results of the mentioned samples are demonstrated in figure


4-39. As can be easily seen, the locations of the valleys in terms of the
temperature have not been changed sharply in the case of austenization for 10
and 15 minutes, while it has been shifted to lower temperatures regarding the
shortest austenization soaking time.
Chapter 4 – RESULTS AND DISCUSSION 85

0.0

Austenization Time = 5 min.

-0.1

Austenization Time = 15 min.

Change in Length (%)


-0.2

-0.3

Austenization Time = 10 min.


-0.4

-0.5
100 200 300 400 500 600 700
Temperature (°C)

Figure 4-39– Dilatation curves of the 22MnB5 steel by increasing the austenization
soaking time from 5 to 15 minutes; austenization temperature = 900°C; deformation
start temperature = 800°C; strain rate = 0.1s-1.

The amount of the phases in the as quenched microstructure of the samples and
the variations of the hardness values are shown in figure 4-40 which includes
the effect of the austenization soaking time. The amount of martensite is
increased while the amount of bainite and ferrite are decreased by increasing
the austenization soaking time from 5 to 15 minutes. The hardness values show
a drop followed by a rise during the same variations of the soaking time.
Chapter 4 – RESULTS AND DISCUSSION 86

200
500 500
432
180
450 450
385.7
HV~

160
400 363.3 400
350
140 350
300
120 300
100 250
Phase Percent (%)

80 75 200
65 martensite
60
60 150
40 25 29 100
22
20 bainite 50
15 6 3 ferrite
0 0
0 5 10 15 20
Austenization Soaking Time (min.)

Figure 4-40– Distribution of the microstructural phases and the hardness values
in terms of the austenization soaking time.

Figure 4-41 demonstrates the variations of the work hardening rate for the
above mentioned samples. As can be seen in this figure there is no evident
difference in the work hardening rate by changing the austenization soaking
time from 5 to 15 minutes.
Chapter 4 – RESULTS AND DISCUSSION 87

3500
Work Hardening Rate (θ) Vs. Strain (ε)
Deformation Start Temperature = 800°C
3000 -1
Strain Rate = 0.1 (s ),Total Strain = 0.2
Temperature at the end of deformation = 700°C
Work Hardening Rate, θ [MPa]

2500

2000 Austenization
Time = 10 min. Austenization
Time = 15 min.
1500
Austenization
Time = 5 min.
1000

500

0
0 0.05 0.1 0.15 0.2 0.25
True Plastic Strain, ε [-]

Figure 4-41– Work hardening rate variations of the deformed samples regarding
different austenization soaking times.

Figure 4-42 confirms the presence of the mentioned phases in figure 4-40 by
the demonstration of the microstructural images of the deformed samples
regarding the austenization soaking time.
Chapter 4 – RESULTS AND DISCUSSION 88

a b

c
Figure 4-42– Microstructures of the deformed samples regarding different
austenization soaking times: a) 5 minutes, b) 10 minutes and c) 15 minutes.

4.2.3.2. Discussion
Basically, it can be imagined that increasing the austenization soaking time
may lead into the shift of the transformation curves to the right hand side of the
CCT diagram and retarding the nucleation and growth based transformations
such as bainitic and ferritic transformations. This means that further
austenization hinders the ferrite and bainite phases to occur. In addition,
considering the stability of the flow curves despite an increase in the
austenization soaking time (figure 4-38), and the fact that the deformation
process in all three cases has been finalized at 700°C, it can be concluded that
the mentioned temperature is higher than the ferritic transformation start
temperature. Therefore, all the specimens have been deformed based on the
same conditions in the austenitic region of the CCT diagram and no
transformation occurs during the deformation process. Moreover, it is found
that the probable change of the austenite grain size by increasing the
Chapter 4 – RESULTS AND DISCUSSION 89

austenization soaking time has no colorful effect on the flow curve behavior of
the specimens.
Regarding the dilatation curves (figure 4-39), normally it is expected that
increasing the amount of produced martensite, results in increasing the depth of
the transformation valley in the dilatation diagram. As is seen in figure 4-39,
this trend is true due to the increase of the austenization soaking time from 5 to
10 minutes. Nevertheless, the decrease of the dilatation despite the increase of
the austenization soaking time (from 10 to 15 minutes) could be justified by
means of the other effective parameters – e.g. internal stresses – than the
occurrence of a more martensitic structure (in 15 minutes samples than 10
minutes ones) which normally increases the depth of the valleys in the
dilatation curves.
In addition, it is observed that decreasing the austenization soaking time to
5 minutes gradually shifts the martensitic transformation sections of the
dilatation diagrams to the left. It means that, shorter austenization time retards
the martensitic transformation, and a higher driving force (i.e. lower Ms) is
required to start the transformation.
This is because of the matter that, at higher austenization times, there is more
opportunity for the growth of the austenite grains. Therefore, the austenite to
martensite transformation becomes much easier. In other words, ferritic and
bainitic transformation are less probable to occur based on the same cooling
rate in comparison with the shorter austenization times. Hence, higher cooling
rate and/or lower Ms are required for the martensitic transformation to happen
after the austenization for a shorter period of time.
By the evaluation of the hardness data and the distribution of the different
phases in the microstructure (figures 4-40 and 4-42), it is found that one of the
most effective parameters for the hardness values to be defined is the amount of
bainite phase in the microstructure, although the amount of martensite has its
own natural effect on the mechanical properties. Due to this fact, despite an
increase in the amount of martensite (from 15 to 65%) and decrease of ferrite
(from 25 to 6%) by increasing the austenization soaking time from 5 to 10
Chapter 4 – RESULTS AND DISCUSSION 90

minutes, the hardness value is decreased because of a decrease in the amount of


bainite phase (from 60 to 29%) in the final microstructure. As the amounts of
ferrite and bainite phases in the microstructures are almost stable (by increasing
the soaking time from 10 to 15 minutes), increasing the amount of martensite
from 65 to 75% leads into an increase in the hardness value.
There is no meaningful difference among the work hardening rate values of the
deformed samples regarding their different austenization soaking times as
shown in figure 4-41. This is because of the same reason as discussed in figure
4-38.
Chapter 4 – RESULTS AND DISCUSSION 91

4.2.4. The Deformation Start Temperature Effect


The variations of the deformation start temperature and its effect on the
mechanical properties of the experimental material are studied here.
Furthermore, the microstructural changes and dilatation data are evaluated
based on their relation to the deformation temperature and conditions.

4.2.4.1. Results
Figure 4-43 gives the flow curves of the continuously cooled and deformed
specimens due to their different deformation start temperatures. As is evident,
increasing the deformation start temperature – i.e. from 700°C to 850°C -
continuously lowers the flow stress level during the deformation.

500
Deformation Start Temperature = 700°C

Deformation Start Temperature = 750°C


400
True Stress, σ [MPa]

300

200 Deformation Start Temperature = 800°C

Deformation Start Temperature = 850°C


100

0
0 0.05 0.1 0.15 0.2 0.25 0.3
True Plastic Strain, ε [-]

Figure 4-43– Flow curves of the continuously deformed samples regarding different
deformation start temperatures.
Chapter 4 – RESULTS AND DISCUSSION 92

5000
Work Hardening Rate (θ) Vs. Strain (ε)
Strain Rate
Strain Rate==0.1
0.1(s(s-1)
-1
)
Total Strain = 0.3
4000
Work Hardening Rate, θ (MPa)

3000 Deformation Start Temperature = 700°C


Deformation Start Temperature = 750°C
Deformation Start Temperature = 800°C

Deformation Start Temperature =850°C


2000

1000

0
0 0.05 0.1 0.15 0.2 0.25 0.3
True Plastic Strain, ε(-)

Figure 4-44– Work hardening rate variations due to different deformation start
temperatures.

The resulted work hardening rate data from the flow curves of figure 4-43 are
presented in figure 4-44 against the increasing exerted strain. It is seen that at
lower amounts of deformation, the specimens which started to be deformed
from the lower temperatures have relatively higher work hardening rate level,
while the order is substantially decreased while coming to higher strains.
Increasing the deformation start temperature from 700°C to 750°C, the
hardness value is decreased (figure 4-45). Further rise of the deformation
temperature increases the hardness value as is shown in figure 4-45. The
mentioned increasing trend of the hardness data is almost linier against the
deformation start temperature.
Chapter 4 – RESULTS AND DISCUSSION 93

345

340 342

335
332.3

330
HV~

329.7

325

320

318.3
315
650 700 750 800 850 900
Deformation Start Temperature (°C)

Figure 4-45– Hardness value variations at different deformation start temperatures.

Figure 4-46 demonstrates the variations of the dilatation curves of the deformed
specimens at different temperatures. There is a rather continuous shift for the
martensitic transformation region to the left hand side of the diagram by
decreasing the deformation temperature, while the depth of the transformation
valleys shows some fluctuations by coming to higher deformation start
temperatures.
Chapter 4 – RESULTS AND DISCUSSION 94

0.0
Deformation Start Temperature = 700°C

Deformation Start Temperature = 750°C


-0.1

Change in Length (%)


-0.2

-0.3
Deformation Start Temperature = 850°C

-0.4 Deformation Start Temperature = 800°C

-0.5
100 200 300 400 500 600 700
Temperature (°C)

Figure 4-46– Dilatation curves of the continuously deformed specimens at different


deformation start temperatures.

a b

c d

Figure 4-47– Microstructure of the deformed samples at different deformation start


temperatures; a) 700°C, b) 750°C, c) 800°C and d) 850°C.
Chapter 4 – RESULTS AND DISCUSSION 95

Microstructural variations of the specimens based on the deformation start


temperatures can be seen by the evaluation of figure 4-47. The amount of
bainite is decreased by increasing the deformation temperature to 750°C.
Coming to the higher temperatures, the fraction of bainite phase in the
microstructures is increased to great extent.

4.2.4.2. Discussion
The flow stress level in figure 4-43 has been totally affected by the deformation
temperature. The lower the deformation start temperature, the higher flow
stress levels are achieved. This is simply because the dislocation motion as the
reason for the formability of metals is intensified by increasing the
experimental temperature. Therefore, the materials can be deformed by lower
stresses at higher temperatures.
As mentioned in previous sections, no phase transformation occurs during the
current simultaneous deformation and quenching tests. This is found because
no change in the slope of the flow stress or work hardening rate curves is
observed. It can be concluded that, as the final structures in all the deformation
temperatures consist of ferrite, bainite and martensite, the start temperature of
the ferritic and bainitic transformations must be less than 550°C in the case of
sample which its deformation is started at 700°C. This is because it takes three
seconds for the strain of 0.3 to be achieved by the strain rate of 0.1s-1. As the
cooling rate is 50°C/sec., the temperature comes at 550°C after three seconds.
Therefore, for the bainitic and ferritic transformations to be appeared without
any effect on the flow curves behavior, the transformations start temperatures
must be less than 550°C. The mentioned calculated low temperature for a static
ferritic transformation shows that the reported ferrite must be categorized as the
strain induced ferrite or ferrite regions in the granular-carbide-free-bainite, and
not the result of a normal static transformation at high temperatures.
Considering figure 4-44, rather higher work hardening rate level in the case of
deformed specimens at lower temperatures in comparison with higher
temperature experiments can be distinguished at the strains of less than 0.2. At
Chapter 4 – RESULTS AND DISCUSSION 96

more than this strain, the mentioned difference is vanished. This is because at
lower temperatures, the applied strain leads into the formation of the
dislocations which are not able to be ordered as fast as they are produced. At
lower strains, this may result in the decrease of the work hardening rate, but the
value of work hardening rate is higher for a lower deformation temperature due
to the higher efficiency of locking more dislocations at lower temperatures. At
higher strains, a kind of saturation in the work hardening rate value is reached.
It means that regardless of the deformation temperature, exerting more strains
on the sample does not change the work hardening rate in any way and a steady
state occurs. This is because for all the samples, there is a compromise between
the rate of the dislocation formation and the amount of ordered dislocations at
the deformation temperature. It means that at each temperature, the rate of
dislocation rearrangement is increased parallel – but not equal- to the rate of the
dislocation formation. Therefore, the specimen is work hardened continuously
by a constant work hardening rate.
Considering the hardness values in figure 4-45 and the microstructural images
in figure 4-47, it is found that the variations of the hardness value mostly
follow the amount of bainite in the final microstructure than ferrite or
martensite content. As can bee seen, the amount of bainite is decreased by
increasing the deformation temperature from 700 to 750°C, and then there is a
continuous increase in the amount of bainite as the deformation start
temperature increases from 750 to 850°C. The amount of bainite at 700 and
800°C specimens is almost the same. The same trend for the variation of the
hardness value is observed. Meanwhile the amount of martensite in the
microstructure is increased by increasing the deformation temperature from 700
to 750°C and after a decrease from 750 to 800°C, it is increased when coming
to 850°C. Except for the 750°C specimen, no meaningful variation for the
amount of ferrite in the final microstructure is observed. In the case of 750°C,
the amount of ferrite phase is increased.
Except for the deformation at 750°C which gives a different behavior, the
evaluation of the hardness and microstructural data shows that by increasing
Chapter 4 – RESULTS AND DISCUSSION 97

the deformation start temperature, the bainitic transformation region is shifted


to the left hand side of the CCT diagram and that is more bainite is formed at a
same cooling rate by increasing the deformation temperature.
Figure 4-46 demonstrates that the martensitic transformation start temperature
is transferred to the lower temperatures as the simultaneous deformation start
temperature is decreased. It is found that the formation of martensite becomes
easier as the deformation takes place at higher temperatures.
This is justified due to the presence of less austenite in the microstructure of the
deformed samples at higher temperatures after finalizing the bainitic
transformation. Because of this, as more amounts of austenite in the
microstructure need a more powerful quenching medium (i.e. less quenching
temperature) to be transformed to martensite, the Ms temperature is decreased.
Although, due to the presence of different parameters (as mentioned in
previous sections), the depth of the valley does not directly show the amount of
martensite formed, but the decreasing behavior of Ms value by decreasing the
deformation temperature is also evident between 800 and 850°C specimens.
For this to be observed, one should consider the broad valley of the martensitic
transformation and its relevant Ms of more than 400°C in comparison with the
Ms value of less than 400°C for 800°C specimen.
Regarding the above mentioned justification and by going to the higher
temperatures in the case of dilatation curves of 800°C and 850°C specimens,
the bainitic transformation valleys can also be distinguished.
Chapter 5 - CONCLUSIONS 98

CHAPTER

Five
CONCLUSIONS

- The attention must be paid to the continuously cooled bainite and ferrite
formation which is observed during both isothermal compression and
simultaneous forming and quenching experiments. This is because even
in case of the isothermal compression, the specimens are under the
continuous cooling prior to and after the deformation.
- During the experiments, almost no phase transformation was observed
during the deformation. This was resulted from the consideration of the
slopes of the flow curves which do not show a sudden change due to a
phase transformation.
- The effect of the prior deformation on generating the nucleation sites of
ferritic transformation is diminished as the deformation temperature is
increased in isothermal compression tests.
- The lower isothermal compression temperature may lead into the
formation of more martensite by hindering the bainitic and ferritic
transformations.
- The colorful increasing effect of bainite phase percentage on the
hardness values beside the martensite content must be taken into account
when studying the 22MnB5 steel.
- During the isothermal compression tests, there is more opportunity for
the continuously cooled bainitic and ferritic transformations to occur at
lower temperatures than higher temperatures.
- The mutual effect of strain rate and deformation temperature on the
work hardening rate values must be considered.
- During the isothermal compression tests, in most cases, the presence of
bainite and ferrite phases can decrease and increase the value of Ms-Mf
respectively.
Chapter 5 - CONCLUSIONS 99

- When evaluating the dilatation curves, the effect of the internal stresses
and prior phase transformations must be taken into account.
- The severe plastic deformation of austenite prior to its transformation
hinders the growth of martensite causing a reduction in the fraction of
the transformation in spite of an increased number of nucleation sites'
density.
- Despite the change of the martensite content, increasing the
austenization time has no meaningful effect on the flow curves and the
resulted work hardening rate data regarding the simultaneous forming
and quenching experiments.
REFRENCES i

REFRENCES

[1] ThyssenKrupp Tailored Blanks Company Information Website,


http://www.tailored-blanks.com.
[2] Steel Bumper Systems for Passenger Cars and Light Trucks, American
Iron and Steel Institute, Revision Number Two, February 15, 2003.
[3] Gehard Schießl et al.: Manufacturing a Roof Frame from Ultrahigh
Strength Steel Materials by Hot Stamping, Technical Report by Audi AG
and ThyssenKrupp Stahl AG, Duisburg, Germany, pp. 158-166, 2004.
[4] X. Bano and JP. Laurent: Heat Treated Boron Steels in the Automotive
Industry, 39th MWSP Conf. Proc., ISS, Vol. XXXV, pp. 673-677, 1998.
[5] Mahesh C. Somani, L. Pentti Karjalainen, Magnus Eriksson, and Mats
Oldenburg: Dimensional Changes and Microstructural Evolution
in a B-bearing Steel in the Simulated Forming and Quenching
Process, ISIJ International, Vol. 41, No. 4, pp. 361-367, 2001.
[6] International Iron and Steel Institute: Advanced High Strength Steel
(AHSS) Application Guidelines, http://www.worldautosteel.org, 2005.
[7] Paul Åkerström: Modeling and Simulation of Hot Stamping, Doctoral
Thesis, Luleå University of Technology, Luleå, Sweden, 2006.
[8] L. Garcia Aranda, P. Ravier and Y. Chastel: Hot Stamping of Quenchable
Steels: Material Data and Process Simulations, Proceedings of the IDDRG
Conference, Bled, Slovenia, pp. 155-164, 2003.
[9] Wever, Archiv, Eisenhüttenwesen, 1928-9, 2, 193.
[10] The Effects of Alloying Elements on Iron-Carbon Alloys, Knowledge
Article from www.Key-to-Steel.com.
[11] R. A. Grange and J. B. Mitchell: On the Hardenability Effect of Boron in
Steel, Transactions of the ASM, Vol. 53, pp. 157-185, 1959.
[12] J. Morral and T. Cameron: Boron in Steel, ed. by S. Banerji and J.
Morral, AIME, Warrendale, PA, 235, 1979.
REFRENCES ii

[13] Shi Chongzhe: The Influences of Austenizing Temperature on the


Hardenability of Boron Steel, Paper for the 5th ICHTM, pp. 441-445.
[14] B. L. Bramfitt and J. G. Speer: A Perspective on the Morphology of
Bainite, Metallurgical Transactions A, Vol. 21A, pp. 817-829, April 1990.
[15] H. K. D. H. Bhadeshia: Bainite in Steels, Second Edition, The University
Press, Cambridge, 2001.
[16] K. A. Ridal and J. McCann: Physical Properties of Martensite and
Bainite, Special Report 93, Iron and Steel Institute, London, 1965,
147-148.
[17] L. J. Habraken, Rev. Met., 53, 1956, 930.
[18] L. J. Habraken, Compt. Rend., 19, 1957, 126.
[19] L. J. Habraken: Physical Properties of Martensite and Bainite, Special
Report 93, Iron and Steel Institute, London, 1965, 147.
[20] L. J. Habraken and M. Economopolous: Transformation and
Hardenability in Steels, Climax Molybdenum, Ann Arbor, Michigan,
USA, 1967, 69-107.
[21] Y. Ohmori, H. Ohtani and T. Kunitake: Bainite in Low-Carbon Low-Alloy
High-Strength Steels, Trans. ISIJ, 11, 1971, 250.
[22] George Krauss and Steven W. Thompson: Ferritic Microstructures in
Continuously Cooled Low and Ultra Low Carbon Steels, ISIJ Journal,
Vol. 35, No. 8, pp. 937-945, 1995.
[23] S. W. Thompson, D. J. Colvin and G. Krauss: Continuous
Cooling Transformations and Microstructures in a Low-Carbon
High- Strength Low-Alloy Plate Steel ,Metall. Trans. A, Vol.
21A, p. 1493, 1990.
[24] Wikipedia, The Free Encyclopedia, http://www.wikipedia.org/.
[25] H. K. D. H. Bhadeshia: Martensite in Steels, University of Cambridge
Online Sources, http://www.msm.cam.ac.uk/phase-trans/index.html.
[26] Key to Steel Website: The Formation of Martensite,
http://www.keytosteel.com/cn/Articles/Art155.htm.
REFRENCES iii

[27] Jiajun Wang, Pieter J. van der Wolk and Sybrand van der Zwaag:
Determination of Martensite Start Temperature in Engineering Steels,
Part I. Empirical Relations Describing the Effect of Steel Chemistry,
Materials Transactions, JIM, Vol. 41, No. 7, pp. 761-768, 2000.
[28] Zenji Nishiyama: Martensitic Transformation, Academic Press Inc., 1978.
[29] S. Morito, H. Tanaka, R. Konishi, T. Furuhara and T. Maki:
The morphology and crystallography of lath martensite in Fe-C alloys,
Acta Mater., 51, 2003, 1789.
[30] S. Morito, H. Saito, T. Ogawa, T. Furuhara and T. Maki: Effect of
Austenite Grain Size on the Morphology and Crystallography of Lath
Martensite in Low Carbon Steels, ISIJ International, Vol. 45, No. 1,
pp. 91-94, 2005.
[31] M. C. Somani, L. P. Karjalainen, M. Oldenburg, M. Eriksson: Effects of
Plastic Deformation and Stresses on Dilatation During the
Martensitic Transformation in a B-bearing Steel, J. Mater. Sci. Technol.,
Vol. 17, No. 2, pp. 203-206, 2001.
[32] H. J. Jun, J. S. Kang, D. H. Seo, W. Y. Choo, and C. G. Park: Effects of
TMP and Accelerated Cooling on Continuous Cooling Transformation
And Microstructure in Low Carbon HSLA Steels With/Without Boron.
[33] F. J. Humphreys: Recrystallization and Related Annealing Phenomena,
Elsevier Science Ltd., 1995.

Anda mungkin juga menyukai