Download as pdf or txt
Download as pdf or txt
You are on page 1of 391

Wolfgang Bleck (Ed.

)
Materials Science of Steel - Textbook for
RWTH Aachen University Students
Steel Institute

RWTH Aachen University

Wolfgang Bleck (Ed.)

Materials Science of Steel


Textbook for RWTH Aachen University Students

Translated from the German Version by Susan Giffin


Edited by (1. Edition)
Wolfgang Püttgen
Edited by (2. Edition)
Piyada Suwanpinij, Yujie Feng, Florian Gerdemann
Edited by (3. Edition)
Zirong Peng, Yujie Feng, Piyada Suwanpinij
Edited by (4. Edition)
Wolfgang Bleck, Christiane Beumers, Yidu Di, Wenwen Song

At this point, all the other unnamed aiders will be thanked.


This book was created as supporting material for the lectures and labatory excersies held at the
Steel Institute at Aachen University.
It cannot be guaranteed that everything is complete and that there are no printing errors.
All rights reserved. No part of the contents of this book may be reproduced or transmitted in any
form or by any means.

4. überarbeitete Auflage, 2016


Preface
Steel is by far the most significant metallic material group. It contains over 2000
steel grades and offers a wide variety of property combinations to be applied to
different uses. Many of these steel grades were recently modified or have been
newly developed. Combined with new processing methods, they are an interesting
and challenging field for engineers. This book is an introduction to the fundamental
laws of material development and should excite students and engineers to further
study the field of steels.
This book, Materials Science of Steel, Textbook for Students at RWTH Aachen
University, introduces the reader to an engineering approach of how to develop
steels and finishing processes, of how to achieve desired properties, and of how to
appropriately use the material. This book begins by acquainting the reader with the
physical and chemical characteristics of the element iron. Important alloying
elements for steels are introduced by means of phase diagrams. The microstructure
formation is the focus of the chapter “Phase Transformation”. The technical
significance and application of these phase transformations are discussed in the
chapter “Technical Heat Treatments”. The practical relevance of the introduced
physical and chemical phenomena for steel development and application is also
explained
In many cases, the texts and figures are based on textbooks and scientific
publications, which were often modified to better fit the structure of this book and
to provide a consistent presentation of the material. A list of the material that was
used, as well as that which is recommended, is given at the end of each chapter. In
spite of the great care that was taken in compiling this book, there very well may be
mistakes, and from the point of view of a student, not all phenomena may be clearly
and comprehensively explained. The authors kindly ask for your indulgence in this
respect and would greatly appreciate any suggestions to improve this edition. Please
send an e-mail with your corrections and suggestions to bleck@iehk.rwth-
aachen.de. This third edition of this book has been revised for improvement of
English translation, for clarity of explanations as well as for an update of graphics
and tables.
A note to students: not all of the material presented in this book is relevant for the
exam. The emphasis and required knowledge will be provided in the lectures,
exercises, and laboratory assignments.

We hope this book will provide a motivating and stimulating reading experience.

Wolfgang Bleck
Contents

Contents
1 Terms, Abbreviations, Symbols ............................................................................. 1
1.1 Definitions ......................................................................................................... 1
1.2 Abbreviations and Symbols .............................................................................. 5
2 The Physical Properties of Iron and Steel .......................................................... 16
2.1 Introduction ..................................................................................................... 16
2.2 Crystal Formation ........................................................................................... 18
2.2.1 The crystal structure of pure iron ........................................................... 18
2.2.2 The influence of foreign atoms on the lattice constant of iron .............. 27
2.2.3 Real Structures ........................................................................................ 27
2.3 Thermal Properties .......................................................................................... 29
2.3.1 Changes in volume and length of iron .................................................... 29
2.3.2 Changes in volume and length of steels ................................................. 33
2.3.3 Thermal conductivity .............................................................................. 36
2.3.4 Diffusion ................................................................................................. 38
2.4 Elastic Properties ............................................................................................ 41
2.4.1 Elastic modulus (Young’s modulus) and shear modulus ....................... 41
2.4.2 Anelasticity ............................................................................................. 46
2.5 Magnetic and Electric Properties .................................................................... 50
2.5.1 Magnetic properties of pure iron ............................................................ 50
2.5.2 Magnetic properties of steels .................................................................. 56
2.5.3 Electric properties ................................................................................... 59
2.6 Further Readings ............................................................................................. 62
3 Iron Alloys.............................................................................................................. 63
3.1 Formation of Alloys ........................................................................................ 63
3.1.1 Homogeneous alloy formation by means of interstitial and
substitutional solution ............................................................................. 64
3.1.2 Heterogeneous alloy formations caused by segregation ........................ 66
3.1.3 Superlattices in substitutional solid solutions and the formation of
intermetallic phases ................................................................................ 67
3.2 Phase Diagrams of Fe alloys ........................................................................... 70
Contents

3.2.1 Phase diagrams of binary systems ........................................................... 70


3.2.2 Thermophysical basics for the expansion or contraction of the J- field. 71
3.2.3 Expansion of the J-field in iron alloys .................................................... 73
3.2.4 Restriction of the J-field in iron alloys ................................................... 74
3.2.5 The iron-carbon phase diagram............................................................... 76
3.2.6 The Phase diagrams of Fe-N and Fe-H ................................................... 80
3.2.7 Ternary iron systems ............................................................................... 80
3.2.8 The Phase diagram of Fe-C-Cr................................................................ 81
3.2.9 The ternary Fe-Cr-Ni system................................................................... 85
3.3 Segregation ...................................................................................................... 88
3.3.1 Segregation processes during solidification ........................................... 88
3.3.2 Segregation behavior during dendritic solidification ............................. 91
3.3.3 Macrosegregation .................................................................................... 93
3.3.4 Segregation in continuous casting .......................................................... 94
3.3.5 Crystal segregation .................................................................................. 97
3.3.6 Secondary segregation............................................................................. 99
3.3.7 Examination of segregation...................................................................104
3.4 Internal Cleanliness in Steel ..........................................................................110
3.4.1 Internal cleanliness ................................................................................112
3.4.2 Sulfide inclusions ..................................................................................115
3.4.3 Oxide inclusions ....................................................................................117
3.4.4 Influence of calcium on the formation of non-metallic inclusions.......119
3.5 Further Readings ...........................................................................................123
4 Phase Transformations .......................................................................................125
4.1 Ferritic-Pearlitic Transformation ..................................................................128
4.1.1 Morphology ...........................................................................................128
4.1.2 Formation of the ferritic microstructure ...............................................132
4.1.3 Formation of the pearlitic microstructure: transformation with
simultaneous precipitation ....................................................................136
4.1.4 Influence of lamellar spacing on mechanical properties ......................143
4.1.5 Influence of alloying elements on pearlite formation ...........................145
Contents

4.1.6 Special forms of pearlite ....................................................................... 148


4.1.7 Application of ferritic–pearlitic steels .................................................. 151
4.2 Martensitic (athermal) Transformation ........................................................ 152
4.2.1 Morphology........................................................................................... 152
4.2.2 Formation of the martensitic microstructure ........................................ 152
4.2.3 Influence of alloying elements on martensite formation ...................... 163
4.2.4 Influence of deformation on martensite formation .............................. 171
4.2.5 Mechanical properties of martensitic steels ......................................... 173
4.2.6 Special forms of martensite and martensite formation ......................... 177
4.2.5 Application of martensitic steels .......................................................... 179
4.3 Bainitic Transformation ................................................................................ 181
4.3.1 Morphology........................................................................................... 181
4.3.2 Formation of the bainitic microstructure .............................................. 184
4.3.3 Temperature range of bainite formation ............................................... 190
4.3.4 Influence of alloying elements on bainite formation ........................... 193
4.3.5 Widmanstätten ferrite ........................................................................... 198
4.3.6 Acicular ferrite ...................................................................................... 199
4.3.7 Application of bainitic steels ................................................................ 199
4.4 Precipitating from a Supersaturated Solid Solution ..................................... 203
4.4.1 Theoretical basics ................................................................................. 204
4.4.2 Variables influencing carbide precipitation ......................................... 216
4.4.3 Ageing ................................................................................................... 219
4.4.4 Application examples ........................................................................... 221
4.5 Further Readings ........................................................................................... 227
5 Technical Heat Treatments ................................................................................ 232
5.1 Hardening ...................................................................................................... 236
5.1.1 Definitions ............................................................................................ 236
5.1.2 Quenching and tempering ..................................................................... 236
5.1.3 Examination of hardenability ............................................................... 252
5.1.4 Calculating the hardenability ................................................................ 255
5.1.5 Case hardening ...................................................................................... 257
Contents

5.1.6 Properties of hardened materials ...........................................................262


5.2 Annealing Treatments ...................................................................................268
5.2.1 Diffusion annealing ...............................................................................268
5.2.2 Coarse grain annealing ..........................................................................271
5.2.3 Normalising ...........................................................................................273
5.2.4 Soft annealing ........................................................................................277
5.2.5 Recrystallisation annealing ...................................................................281
5.2.6 Stress-relief annealing ...........................................................................289
5.2.7 Combined annealing processes .............................................................290
5.2.8 Wire patenting .......................................................................................291
5.3 Description of Austenite Transformation for Technical Applications .........293
5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams) ...................296
5.4.1 Austenitisation with isothermal heating ...............................................299
5.4.2 Austenitisation with continuous heating ...............................................302
5.4.3 Influence of austenitisation ...................................................................307
5.5 Transformation Diagrams ..............................................................................309
5.5.1 Isothermal transformation .....................................................................309
5.5.2 Continuous transformation ....................................................................320
5.5.3 Other possibilities for the representation of transformation behavior .324
5.5.4 Influences on the transformation behavior ...........................................326
6 Thermomechanical Treatment ...........................................................................336
6.1 Terminology ..................................................................................................337
6.2 Role of microalloying elements ....................................................................339
6.2.1 Solubility behavior of microalloying elements .....................................341
6.2.2 Precipitation kinetics .............................................................................346
6.2.3 Mechanisms of microalloying elements ...............................................347
6.2.4 Influence of microalloying elements on austenite grain growth ..........350
6.2.5 Influence of microalloying elements on softening behavior ................350
6.2.6 Influence of microalloying elements on transformation behavior .......353
6.2.7 Precipitation hardening .........................................................................354
6.3 Factors influencing the TMT of microalloyed steels ....................................358
Contents

6.3.1 Austenitisation ...................................................................................... 358


6.3.2 Deformation .......................................................................................... 359
6.3.3 Cooling .................................................................................................. 363
6.3.4 Influence of deformation on the JėD transformation ......................... 364
6.3.5 Coiling temperature .............................................................................. 365
6.3.6 Summary of the influencing variables .................................................. 366
6.4 Further Readings ........................................................................................... 370
7 Index ..................................................................................................................... 374
1.1 Definitions

1 Terms, Abbreviations, Symbols


1.1 Definitions
The element iron has the symbol Fe (ferrum) and is a transition element in Group
VIII of the periodic table of elements. Iron has the atomic number 26 and an atomic
mass of 55.8. It is the fourth most common element in the Earth’s crust
(5.1 mass-%). The liquid core of the Earth consists predominantly of molten iron.
Iron is rarely found as a pure element in the Earth’s crust, but is frequently found
bound as an oxide, sulfide, carbonate, or silicate.

Elemental iron has a low chemical stability; it is easily oxidised, and directly reacts
with most non-metallic elements by forming bonds in which iron assumes an
oxidation number of +2 or +3 (and in special cases +6). For example, iron can bond
with oxygen to form FeO (+2) and Fe2O3 (+3), as well as Fe3O4 (FeO˜Fe2O3). In the
presence of oxygen, water, and electrolytes, a group of electrochemical reactions
that fall under the term ‘corrosion’ occur. At its various levels of oxidation, iron
can form organometallic bonds with carbon-containing phases. Iron can also form
non-stoichiometric bonds (alloys) with most of the metallic elements.

The name has its roots in the Illyric word “iser”, which together with the Venetain-
Illyric “eisarnon”, in which lies the Indogermanic root “eis”, developed into the
Germanic term “isarnan”. Through the movement of languages, the Gothic word
“eisarn”, the Old-nordic “isarn”, the German “eisen”, the Dutch “ijzer”, the Swedish
“järn”, and the English term “iron” evolved. The scientific name “ferrum” has
Latin-Eutruscian roots; the French term “acier” comes from the Latin “acies”, which
means sharp.

The beginning of the “iron age”, at around 1500 BC, marks the introduction of iron
as a useful metal for weapons and kitchen utensils. Before that, iron was
sporadically used in precious jewelry. For example, in Troy a piece of golden
jewelry was found which had been decorated with the then valuable iron. Egyptian
jewelry has also been known to be made of iron. Iron was probably discovered
coincidentally while melting copper – possibly because, for example, a supposed
copper ore, which in fact was iron ore, was put in the furnace. In this manner,
copper itself was most likely accidentally discovered during clay baking.

The history of iron production begins with the Hittites, around 2000 BC, in the
area known today as Asia Minor. For the first time, the targeted production of iron
and its trade products was established and passed down. Through the Phoenicians,
Etruscans, and Romans, the art of iron production made its way to the Slovenians
and Germans. Iron was mainly won in a one-step process of reducing oxide or

1
1 Terms, Abbreviations, Symbols

sulfide ores. The resulting doughy mass, which was mixed with slag, required
intensive forging to separate and reduce the slag before it was useable. In Europe,
the two-step process was first developed in the 14th century. This process includes
the creation of liquid pig iron as an intermediate product, and a subsequent refining
process to produce forgeable iron through the oxidation of tramp elements, resulting
in the purification of iron and its transformation into steel. Significant
advancements were made during the industrial revolution in the second half of the
18th century. The first examples of industrial applications were boilers, train tracks,
and bridges. Particularly symbolic are the iron bridge in Coalbrookdale, Great
Britain (built in 1779), and the Eiffel tower in Paris (built in 1889).

The biological properties of iron distinguish it as a vital mineral for humans and
animals. It is the central atom in hemoglobin and myoglobin molecules.
Hemoglobin, found in the red blood cells, transports oxygen from the lungs to the
body’s cells, and carries carbon dioxide from peripheral cells to the lungs.
Myoglobin transports oxygen within muscle fibers. An iron deficiency in humans
results in a reduced oxygen supply to the brain, one noticeable symptom of which is
tiredness. The iron content of blood is an important factor for performance. Iron
can only be absorbed into the body in its ionic form (+2), and in the presence of
vitamin C.

The term steel is used to identify a material group of which the main element is iron
(iron based alloys). Iron forms non-stoichiometric bonds with other metallic and
non-metallic elements, which are called alloys. The alloying elements can either
replace the iron atoms in the crystal lattice (substitutional), or they can dissolve in
the spaces (vacancies) within the crystal lattice (interstitial). About 2000 steels are
technically relevant, and usually have several alloying elements along with the
element iron. A typical alloying element is carbon, which can be found between
0.0002 and 2.0 mass-% in steels. Cast iron usually has a higher carbon content of
around 2.0 to 4.0 mass-%, and is further processed using casting techniques. On the
other hand, steels are usually further processed by means of hot deformation.

Important reasons why the element iron is universally, technically applicable:


x Its abundance (5.1% of the earth’s crust: deposits of >50 mass-% iron can be
found worldwide).
x Its high melting temperature TL of 1808 K (1535°C): as a result, from room
temperature up to around 450°C (= 0.4 TL), no thermally activated
microstructural changes take place. Within the technically, well controllable
temperature range of 550 to 950°C, heat treatment can be used to activate
diffusion processes and to specifically adjust microstructures.

2
1.1 Definitions

x Iron has two first order phase transformations in the solid state (A4 and A3-
transformations), which allow for the use of different crystal structures in steels.
Many properties of the solid can be calculated from the geometry of the body-
centered cubic lattice and the face-centered cubic lattice.
x A magnetic transformation (A2-transformation, second order phase
transformation) allows for materials with diverse electrical and magnetic
properties, and leads to useful anomalies, such as those which occur during
thermal expansion.
x Iron can be alloyed with about 80 other elements that involve mutual solubility
or bond formation. Therefore the solution of the foreign atoms can take place
interstitially or substitutionally.
x In alloys or under high pressure a hexagonal lattice can also emerge besides a
body-centered cubic and a face-centered cubic lattice. Apart from
paramagnetism and ferromagnetism, antiferromagnetism also occurs. The Curie
temperature indicates the transition from ferromagnetism to paramagnetism;
while the Néel temperature shows the transition from antiferromagnetism to
paramagnetism. The possible crystal structures and the respectively appearing
forms of magnetism are summarized in Figure 1.1.

Figure 1.1: Crystal structures and magnetism forms in iron and iron alloys.
The term material refers to all natural and synthetic substances. The term material
is used here to describe a solid aggregate state, from which components and
structures can be produced using specific processes.

3
1 Terms, Abbreviations, Symbols

A material is distinguished by its technically useful properties, its technical and


economical possibilities, as well as its ecological soundness. In materials science,
materials are characterised by their crystalline structure, their microstructure (type
and distribution of the lattice defects) and their macrostructure (macroscopic
appearance).

The traditional term ferrous metallurgy describes a scientific field of engineering


that deals with the production, processing, and application of iron and steel; ferrous
metallurgy is, therefore, a division of materials science. Metallurgy is the science
of producing and processing metallic materials; the emphasis is on processing
technology. Materials science deals with the technically relevant aspects of the
production, processing, and application of materials. The scientific expansion in
this field of study began in Germany in the second half of the 19 th century with the
appointment of professors of ferrous metallurgy at the Universities of Aachen,
Berlin, Clausthal, and Freiberg.

4
1.2 Abbreviations and Symbols

1.2 Abbreviations and Symbols


Chapter 2: Physical properties of iron and steel
Symbol Unit Definition
A °C or K transformation temperature
(K = Kelvin)
AC °C or K transformation temperature during
heating
Ar °C or K transformation temperature during
cooling
a m lattice parameter
a1 ° oscillation amplitude 1
a2 ° oscillation amplitude 2
-1
D K linear thermal expansion coefficient
B T=Vs m-2 magnetic induction
Br T remnant
Bs T saturation induction
b m atomic spacing
bcc - body-centered cubic crystal
d m material thickness
'd m thickness difference
E MPa Elastic modulus / Young's modulus
(MPa = Mega Pascal)
e As elementary charge
(A s = Ampere Second)
H - elongation
H1 - elastic elongation percentage
H2 - anelastic elongation percentage
f s-1 frequency
fcc - face-centered cubic crystal
G MPa shear modulus
-1
J K cubical thermal expansion coefficient
J - shear angle

5
1 Terms, Abbreviations, Symbols

Symbol Unit Definition


-1
H Am magnetic field strength
-1
HC Am coercive field strength
-1
k JK Boltzmann constant
(J = Joule)
L V2 K-1 Lorenz constant
(V = Volt)
l m length
/ - logarithmic decrement
-1 -1
O Wm K thermal conductivity
(W = Watt)
Ma g absolute atomic mass
-1
Mr g mol relative atomic mass
μ - Poisson's ratio
-1
μ0 Vs Am induction constant
μr - relative permeability
VR(bcc) % packing density of the bcc-lattice
VR(fcc) % packing density of the fcc-lattice
r m atomic radius
rs m spherical radius of a lattice interstitial
site
Uth g cm-3 theoretical density
-3
U g cm material density
Usp = U (T) :cm specific electrical resistance
UPh :cm temperature dependent amount of the
specific electrical resistance
U0 :cm residual resistance
V : cm -1
electrical conductivity
V MPa normal stress
T °C or K temperature
TC °C or K Curie temperature
TL °C or K melting point (Liquidus temperature)

6
1.2 Abbreviations and Symbols

Symbol Unit Definition


'T °C or K temperature difference
W MPa shear stress
3
Vat m atomic volume
3
VU m unit cell volume
3
'V m volume difference

Chapter 3: Iron alloys


Symbol Unit Definition
A M-% maximum solubility
a mol/l activity
co mol/l melt concentration
cS mol/l crystal concentration
cL mol/l liquid concentration
-1
H KJmol enthalpy of solution
-1
HD   KJmol enthalpy of solution in ferrite
-1
HJ   KJmol enthalpy of solution in austenite
k0 - solubility coefficient
1/ko - segregation coefficient
P bar partial pressure
-1 -1
R J mol K universal gas constant
'S J entropy
T °C or K temperature

Chapter 4: Phase transformations


Chapter 4.1: Ferritic-Pearlitic Transformation
Symbol Unit Definition
a, b, c m lattice parameters
D ° angle between lamella axis and
thin section surface
CD mass-% carbon concentration of ferrite

7
1 Terms, Abbreviations, Symbols

Symbol Unit Definition


c
C mass-% carbon concentration of cementite
J
C mass-% carbon concentration of austenite
D - diffusion coefficient,
D = D0 exp(-Q/RT)
'EV J total gain in enthalpy, in relation to
volume
'GB J mol-1 interfacial enthalpy
-1
'GV J mol gain in enthalpy through
transformation
'HV J m-3 volume enthalpy
J - diffusion current
O m lamellar spacing (for D =0)
OD m lamellar spacing (within angle D)
Oa m average lamellar spacing of pearlite
Omin m minimum distinguishable lamellar
spacing of pearlite
Q J activation energy per jump
-1 -1
R J mol K gas constant, R = 8.314 J/mol K
T °C or K temperature
'T °C or K supercooling
TE °C or K equilibrium temperature between
two-phases
TT °C or K transformation temperature
-2
V Jm interfacial energy
W m width of lamella
x
W m s-1 mobility rate
x m path
z n number of atomic jumps (flips)

8
1.2 Abbreviations and Symbols

Chapter 4.2 and 4.3: Martensitic and bainitic Transformation


Symbol Unit Definition
a, b, c m lattice parameters
Ad °C or K decreased austenite start temperature,
deformation induced
Af °C or K austenite finish temperature
As °C or K austenite start temperature
B1 - low carbon bainite
B2 - upper bainite
B3 - lower bainite
Bf °C or K bainite finish temperature
Bs °C or K bainite start temperature
-1
GA J mol enthalpy of volume of austenite
-1
GM J mol enthalpy of volume of martensite
-1
'G J mol nucleation enthalpy
LM - lath martensite
Md °C or K increased martensite start temperature,
deformation induced
Md 30 °C or K increased martensite start temperature,
after 30% cold deformation
Mf °C or K martensite finish temperature
Ms °C or K martensite start temperature
PM - plate martensite
RA - retained austenite
'T °C or K supercooling

Chapter 4.4: Precipitation out of the supersaturated solid solution


Symbol Unit Definition
B - constant
BH - bake-hardening
c0 mass-% initial concentration in the matrix
on the phase boundary

9
1 Terms, Abbreviations, Symbols

Symbol Unit Definition


cC/N mass-% equilibrium concentration
of carbon or nitrogen in D-iron
cc (T) mass-% equilibrium concentration
of carbon in D-iron
cE mass-% equilibrium concentration with quasi-
planar interface, equilibrium concen-
tration according to the phase diagram
cP mass-% equilibrium concentration
within the precipitated carbide
cRT mass-% carbon concentration at room
temperature
c(r) mass-% equilibrium concentration in front of
the particles with radius r

cm mass-% concentration in the matrix
2 -1
D cm s diffusion coefficient
-1
Hel kJ mol elastic strain energy
-1
'gu kJ mol change in free molar
volume enthalpy
'GO kJ mol-1 nucleation energy
-1
'G(r) kJ mol change in free enthalpy
-1
'h kJ mol difference in molar enthalpy
-2 -1
jD cm s diffusion current density
n - number of atoms
n - precipitation exponent
N nucleation rate
-1 -1
k J mol K Boltzmann constant
3
: m molar volume of particle matter
r m nucleus radius
r0 m nucleus radius at time t=0
rc m critical radius of nucleus

10
1.2 Abbreviations and Symbols

Symbol Unit Definition


-1
Q J mol activation energy in order to bring
1 mol of carbon or nitrogen atoms
in the D-matrix
R J mol-1 K-1 gas constant, R = 8.3143 J mol-1 K-1
R m attainable particle end-radius
-1 -1
's J mol K difference in molar entropy
-1
V kJ mol specific energy of the DEphase
boundary
t s time
W s time constant
T K absolute temperature
W precipitated volume fraction
x m half of the average particle spacing

Chapter 5: Technical heat treatments


Chapter 5.1: Hardening
Symbol Unit Definition
A % fracture elongation
Av J notch impact energy
b - alloying element regression coefficient
for the distance from the face x,
(hardenability calculation)
b0 - constant (hardenability calculation)
Jx HRC hardness, dependent on distance from
the surface of the sample
K - value, dependent on martensite
fraction, used to determine maximum
hardness
R - hardening degree = actual hardness /
maximum possible hardness
Rm MPa tensile strength
Rp0.2 MPa 0.2% offset yield strength

11
1 Terms, Abbreviations, Symbols

Symbol Unit Definition


Vbw MPa fatigue strength
x mm distance from face (Jominy test)
Z % reduction in area

Chapter 5.2: Annealing


Symbol Unit Definition
n - time exponent
-1
Q kJ mol activation energy for recrystallisation
-1 -1
R J mol K gas constant, R = 8.314 J mol-1 K-1
T °C or K annealing temperature
TR °C or K recrystallisation temperature
W s recrystallisation time of the JMAK
Equation
W0 s time constant of the JMAK Equation
t s annealing time
W(t) - fraction recrystallised

Chapter 5.3: Description of the austenite transformation for technical


applications
Symbol Unit Definition
Ac1 °C or K temperature at which formation of
austenite begins during heating
(temperature of three-phase
equilibrium D + J + carbide)
Ac1b °C or K for non-alloyed and alloyed steels:
temperature at which formation of
austenite begins during heating
(entering three-phase field D + J+
carbide)

12
1.2 Abbreviations and Symbols

Symbol Unit Definition


Ac1e °C or K temperature at which first
transformation ends during heating,
i.e. three phase area (D + J + carbide)
is exited and system enters two-phase
field (D + J or J + carbide)
Ac3 °C or K temperature, during heating, at which
transformation from ferrite into
austenite ends
Acc °C or K temperature at which dissolution of
carbides in austenite ends for alloyed
steels during heating
Accm °C or K temperature at which dissolution of
cementite in austenite ends for hyper-
eutectoid steels during heating
Mf °C or K temperature at which transformation
from austenite to martensite is nearly
finished during cooling
Ms °C or K temperature at which transformation
from austenite to martensite begins
during cooling

Chapter 5.4: TTA phase diagrams


Symbol Unit Definition
a - constant
C % mass content of carbon
Mi % mass content of phase i
ni - constant
TT °C or K transformation temperature

Chapter 5.5: TTT phase diagrams


Symbol Unit Definition
A - initial phase
D - material constant
2 -1
D cm s diffusion coefficient

13
1 Terms, Abbreviations, Symbols

Symbol Unit Definition


G m s-1 growth rate
-3 -1
K m t nucleus growth rate
Km K/s upper critical cooling rate; only with
regard to martensite
KP s, minute, hour cooling time which leads to a complete
pearlitic microstructure
Kf s, minute, hour quickest cooling time for steels to form
1 % proeutectoid ferrite
Kk s, minute, hour quickest cooling time for
hypereutectoid steels to form 1
% proeutectoid carbide
O - cooling parameter
Ms °C or K martensite start temperature
n - exponent
-1
Q J mol activation energy
-1 -1
R J mol K general gas constant
SV Pm , nm
-1 -1
effective grain boundary area
per unit volume
t s time
W - temperature-dependent time constant
T °C or K temperature
TE °C or K equilibrium temperature
TEV °C or K equilibrium temperature of
proeutectoid ferrite formation
TEP °C or K equilibrium temperature of
proeutectoid pearlite formation
TEB °C or K equilibrium temperature of
proeutectoid bainite formation
TEM °C or K equilibrium temperature of
proeutectoid martensite formation
U - transformation product
W Volume-% transformed phase amount

14
1.2 Abbreviations and Symbols

Chapter 6: Thermomechanical Treatment


Symbol Unit Definition
A, B - constants
Mcrit - critical degree of deformation
K - equilibrium constant
Tcr °C or K critical temperature for a complete
recrystallisation before precipitation
occurs
TNR °C or K recrystallisation stop temperature

15
2 The Physical Properties of Iron and Steel

2 The Physical Properties of Iron and Steel


2.1 Introduction

The historical significance of iron


In the periodic table of elements, iron has the symbol Fe, which is derived from the
Latin word ferrum. The significant use of iron is demonstrable from approx.
2000 BC. Seen historically, iron came to be used much later than lead, gold,
copper, and tin because of the considerable difficulty involved in processing and
handling it. In the beginning, pig iron was extracted from ore in simple, charcoal
heated ovens. However, in contrast to other metals, iron’s extreme hardness
aroused the interest of the military for its possible technical applications. Weapons
made of iron were significantly more robust than those made of bronze. Due to the
high costs, however, the Romans were the first people to commonly use iron
weapons.
Advancements in iron extraction were first achieved in the Middle Ages with the
development of the shaft furnace, the predecessor of the present blast furnace. In
the 16th century, Agricola (De Re Metallica libri XII) described the different
methods of iron production in detail. In the 18 th century, the economic importance
of iron increased as a result of improvements in furnaces and the replacement of
charcoal with coal and coke. All of the important processes of steel production,
including some still in use today, have been developed since the middle of the 19 th
century: In 1855 the Bessemer process; in 1864 the open hearth process; in 1877
the Thomas process; and in 1880 the electric process by Siemens. In 1948, Durrer
and Hellbrügge developed the surface injection method known as the LD (Linz-
Donawitz) process.

Sources of iron
The universe consists of 0.0014 mass-% iron, thus iron is the 9 th most abundant
element in the universe. It is the most important heavy metal and the fourth most
common element on Earth after oxygen, silicon, and aluminum. The inner and outer
cores (‡ 7000 km) of earth consist essentially of iron. Except for the pure form
found in meteors and basalt, iron is only found in bound form in nature. In most
cases, it can be found as an oxidised mineral, such as hematite, magnetite, or
limonite. In 2003, the largest producers of iron ore were Brazil, China, Russia,
Ukraine, Australia, Kazakhstan, and Canada. The largest deposits of iron ore in
Europe can be found in Sweden and France.
The body of an adult human, with an average weight of 70 kg, contains 4 to 5 g of
iron. 70% of the body’s iron can be found in the hemoglobin, 25% in iron
containing proteins, and 3.5% in myoglobin. The typical daily allowance for iron

16
2.1 Introduction

varies from 5 to 40 mg. Iron is non-toxic; a human would have to consume more
than 50 g of iron before it becomes life threatening.

Isotopes
Iron has four stable isotopes, of which, with a 91.7% share, Fe-56 is the most
common. Fe-54 follows with 5.8%, Fe-57 with 2.2%, and Fe-58 with 0.3%. Of the
nine radioactive isotopes, Fe-60 and Fe-59 have the longest half-lives, with 300,000
and 2.7 years, respectively. Fe-49 decomposes with a half-life of only 75 ms. The
radioactive isotope Fe-59 is used in medicine for diagnostic purposes. The
chemical properties of iron are summarised in Table 2.1. Figure 2.1 shows the
location of iron and its properties specified in the periodic table of elements.

Relative atomic mass 55.847


Ionization energy 7.870 eV
Configuration [Ar] 3d6 4s2
Melting point 1808 K
Boiling point 3023 K
Oxidation numbers 6, 5, 4, 3, 2, 1, 0, -1, -2
Theoretical density 7.87 g/cm3
Atomic radius 124.1 pm (a)
Electronegativity 1.8
Ionic radius 67 pm (+3), 82 pm (+2)
Table 2.1: Chemical properties of iron.
Atomic number Electronegativity
3
Relative atomic mass Theoretical density in g/cm
26 1.8
24 25 55.847 7.87 27 28

Cr Mn 1808
Fe 124.1
Co Ni
3134 7.870
Melting point in K Atomic radius in pm
Boiling point in K Ionization energy in eV
Figure 2.1: The position of iron in the periodic table of elements.

17
2 The Physical Properties of Iron and Steel

2.2 Crystal Formation

2.2.1 The crystal structure of pure iron


A specialty of iron is that it can transform to different crystal structures in the solid
state; this behavior is called polymorphism. The polymorphic transformation of
pure iron and its alloys makes the application of these materials possible for
technological purposes. The temperatures at which the transformations of the
crystal structure in pure iron occur can be determined using thermal analysis
(Figure 2.2). The transformation points are recognisable as steps in the temperature
curve, where the temperature does not change. At the points of transformation there
are delays in temperature change during cooling and heating. The transformation
temperature A (French, Arrêt) is indicated with an r (French, Refroidir) if it takes
place during cooling, and with a c (French, chauffer) if it takes place during heating.
The crystallographic structures of pure iron can be divided into D-iron, J-iron and,
G-iron. The body-centered cubic (bcc) lattice of D-iron is thermodynamically stable
below 911°C. When heated slowly, maintaining near-equilibrium conditions, to
temperatures above 911°C (1184 K), the Ac3-temperature, the D-crystal structure
transforms into the face-centered cubic (fcc) crystal structure of J-iron. Above
1392°C (1565 K), the Ac4-temperature, the bcc crystal structure reappears as G-iron.
This structure is thermodynamically stable until the melting point TL = 1535°C
(1808 K) is reached. When cooled at a rate near to thermodynamic equilibrium, the
same transformation temperatures apply. However, heating or cooling at a faster
rate leads to a hysteresis, which is depicted as superheating or supercooling.
A second order transformation, which cannot be directly determined by thermal
analysis, takes place at the so called Curie temperature, A2 = 769°C (1042 K). At
this temperature, instead of a change in the crystal structure, a change in the
magnetic properties takes place. Above the A2-temperature iron is paramagnetic,
and below this temperature it is ferromagnetic.
On the left-hand side of Figure 2.3 a bcc iron unit cell and on the right-hand side a
fcc iron unit cell can be seen. The black dots indicate the position of the Fe-atoms.
The distance to the next neighbor is indicated by dashed lines. The lattice constants
are given for D-iron at room temperature, and for J-iron at 911°C. The
characteristics of these crystal structures will be discussed in detail in the following
sections.

18
2.2 Crystal Formation

1800
Liquid
1600
TL = 1536 °C TL = 1536 °C

G-iron
1400
Ar4 = 1392 °C Ac4 = 1392 °C
Temperature in °C

1200

J-iron
1000
Ar3 = 911 °C Ac3 = 911 °C

800
Ar2 = 769 °C
D-iron Ac2 = 769 °C

600

400
Time
Figure 2.2: Cooling curve (left) and heating curve (right) of pure iron as
determined using thermal analysis; the rate of temperature change is
near to thermodynamic equilibrium.
D-iron J-iron

Lattice constant Lattice constant


a (bcc) = 0.286 nm a (fcc) = 0.364 nm

at 25 °C at 911 °C
Figure 2.3: Crystal structures of pure iron.

19
2 The Physical Properties of Iron and Steel

The body-centered cubic lattice


According to the point lattice model, the atomic nuclei of the body-centered cubic
(bcc) lattice are located at the corners and in the center of the unit cell (Figure
2.4a). According to the ball model (Figure 2.4b), which takes into consideration
the expansion of the electron shells, the atoms come into contact along the three-
dimensional diagonals. These diagonals through the center of the cell, the cell
diagonals <111>, correspond to the closest packed direction of the bcc lattice. The
atomic distance b is defined as the distance between two atomic nuclei along the
closed packed direction. Figure 2.4c shows the relationship between the atomic
distance b, the lattice constant a and the atomic radius r for the bcc lattice of D-iron
at room temperature, which can be calculated as follows:
a
r ˜ 3 1.24 ˜1010 m (2.1)
4
a
b 2r ˜ 3 2.48 ˜1010 m (2.2)
2
Since the volume of the unit cell is not completely filled by the atomic spheres,
interstitial sites exist in the lattice between the atoms. Altogether, a bcc unit cell
contains two full atoms: 1/8 of an atom at each of the eight corners of the cube, and
an entire atom in the center. The packing density VR(bcc) can be calculated as
follows:
3
4 8 §a ·
2 ˜ S ˜ r3 S¨ ˜ 3¸
3 3 ©4 ¹ S˜ 3
VR( bcc) 68% (2.3)
a3 a3 8
The name of the interstitial sites can be derived from the geometric arrangement of
their surrounding atoms: octahedral and tetrahedral interstitial sites. An octahedral
interstitial site is located in the middle of each face, and at the middle of each edge
of the unit cell (Figure 2.5). Each tetrahedral interstitial site is surrounded by two
corner atoms and two central atoms. Altogether, in one bcc unit cell there are six
octahedral and twelve tetrahedral interstitial sites. The size of the interstitial sites
can be calculated as the spherical radius rS that passes directly to the center of the
space created by the surrounding atomic spheres. The following applies:
rs
Octahedral interstitial site: 0.155 Ÿ rs 0.19 ˜1010 m (2.4)
r
rs
Tetrahedral interstitial site: 0.290 Ÿ rs 0.36 ˜1010 m (2.5)
r
where r = atomic radius of an iron atom.
The theoretical density Uth of bcc iron can be calculated using the lattice parameters
and the absolute atomic mass Ma of iron. Given the Avogadro constant Na = 6.02 ˜

20
2.2 Crystal Formation

1023 mol-1 and the relative atomic mass Mr = 55.89 g/mol, the absolute atomic mass
Ma of iron in a bcc unit cell can be calculated:
Mr
Ma 2˜ 18.55 ˜1023 g (2.6)
Na
In order to calculate the density Uth, the volume of the unit cell VU(bcc) is then
substituted into the formula:
Ma g
U th 7.88 (2.7)
VU ( bcc) cm3
The value of the theoretical density corresponds closely to the actual density
(7.87 g/cm3) at room temperature.

4r
—a

—a
—a = 4r
a Unit cell of a bcc b Unit cell according to c Geometric relationships
point lattice the ball model in the bcc lattice

Figure 2.4: Body-centered cubic structure.

21
2 The Physical Properties of Iron and Steel

a/2 a—5/4
a—3/2 a—2 a—3/2

Metal atoms Metal atoms

Interstitial atoms in Interstitial atoms in


octahedral interstitial sites tetrahedral interstitial sites
Figure 2.5: Interstitial sites in the body-centered cubic lattice.

The face-centered cubic lattice


In the face-centered cubic (fcc) lattice, the atomic nuclei are located at the corners
and in the center of each face of the unit cell (Figure 2.6a). According to the ball
model, the atoms come into contact along the plane diagonals <110> (Figure 2.6b).
With the help of Figure 2.6c, the atomic radius r and the atomic distance b can be
calculated for the fcc lattice of J-iron at 911°C as follows:
a
r ˜ 2 1.29 ˜1010 m (2.8)
4
a
b ˜ 2 2.57 ˜1010 m (2.9)
2
Altogether, an fcc unit cell contains four full atoms: 1/8 of an atom at each of the
eight corners of the cube, and a half of an atom at the center of each of the faces.
The packing density VR(fcc) for an fcc unit cell is given by the following:
3
4 §a ·
4˜ S ˜¨ ˜ 2¸
3 ©4 ¹ S˜ 2
VR( fcc ) 74% (2.10)
a3 6
As shown in Figure 2.7, an octahedral interstitial site is found at the middle of each
unit cell edge, and in the center of the cell. The tetrahedral interstitial sites are
found at 1/4 the distance along the three-dimensional diagonal from each corner of
the cell. Altogether, there are four octahedral and eight tetrahedral interstitial sites
in an fcc unit cell. The sizes of the interstitial sites in the fcc lattice are:

22
2.2 Crystal Formation

rs
Octahedral interstitial site: 0.41 Ÿ rs 0.53 ˜10 10 m (2.11)
r
rs
Tetrahedral interstitial site: 0.22 Ÿ rs 0.28 ˜10 10 m (2.12)
r

—a
4r

—a = 4r
a Unit cell of the fcc B Unit cell according to c Geometric relationships
point lattice the ball model in the fcc lattice

Figure 2.6: Face-centered cubic structures.


a/2 a—2
a—3/4

a/—2

Metal atoms Metal atoms

Interstitial atoms in Interstitial atoms in


octahedral interstitial sites tetrahedral interstitial sites
Figure 2.7: Interstitial sites in the face-centered cubic lattice.
The theoretical density Uth of iron in an fcc cell can be calculated using the absolute
mass Ma of the atoms as follows:
Mr
Ma 4˜ 37.10 ˜1023 g (2.13)
Na

23
2 The Physical Properties of Iron and Steel

Ma g
U th 7.69 (2.14)
VU ( fcc ) cm 3

The value of the theoretical density of fcc iron (7.69 g/cm3) agrees approximately
with the actual density (7.65 g/cm3) at 911°C. The slightly smaller value for the
actual compared to the theoretical density is due to vacancies in the crystal lattice
that are not taken into consideration in the calculations.
The hexagonal lattice
The hexagonal lattice consists of several layers of hexagonal lattice points. As
shown in Figure 2.8, the c-axis differs from the a-axis in length.

Figure 2.8: Hexagonal structure.


The actual unit cell that consists of two atoms is depicted in grey (Figure 2.8, left).
However, for a better illustration of a hexagonal structure, three unit cells are
usually presented. If all the atoms are the same size, then they touch each other in
the hexagonal base plane and in the neighboring planes (Figure 2.8, middle). In this
case the c/a ratio can be calculated as follows:
c 8
1.63 (2.15)
a 3
This structure is called hexagonal close-packed (hcp) and occurs, for instance, in
magnesium. This lattice type is very similar to the face-centered cubic lattice. In
the fcc lattice the plane containing the three face diagonals also has a hexagonal
structure; the fcc lattice also corresponds to a stack of closed-packed planes.
However, the two lattices have a different stacking sequence of layers: ABCABC in
fcc and ABABAB in hcp. Figure 2.9 shows the interstitial sites in the hexagonal
lattice.

24
2.2 Crystal Formation

Figure 2.9: Interstitial sites in the hexagonal lattice.


Interstitial sites are of vital importance for the formation of iron alloys. Table 2.2
shows the atomic radii of elements that dissolve interstitially. One can see that most
of the elements have a radius larger than that of the largest interstitial site.
Therefore, when an interstitial element is dissolved in an interstitial site, a lattice
distortion results. Hydrogen is the only exception; it fits into the octahedral
interstitial sites of the fcc lattice and the tetrahedral interstitial sites of the bcc lattice
without causing any distortions. Other interstitially dissolved elements are
embedded into the octahedral interstitial sites of the bcc lattice because the
distortion energy resulting from the shifting of neighboring atoms to increase the
size of the interstitial sites is lower than that of embedding the elements in
tetrahedral interstitial sites.
The characteristic properties of the bcc and fcc lattices are summarised in Table
2.3. The respective lattice constants are represented by the symbol a.
Element Atomic radius
in 10-10 m
H 0.32
N 0.75
C 0.77
O 0.72
B 0.82
He 0.93
Table 2.2: Atomic radii of interstitially dissolved elements in iron lattices.

25
2 The Physical Properties of Iron and Steel

Body-centered Face-centered Hexagonal


cube cube
Lattice constant a = 2.86˜10-10 m a = 3.64˜10-10 m
at room at 911°C
temperature
Volume of the unit cell a3 a3 2.59·a2·c
Number of atoms per 2 4 2
unit cell
Number of adjacent 8 12 12
neighboring atoms
Distance to adjacent 0.866 a 0.707 a
neighboring atoms
Number of octahedral 6 4 2
interstitial sites
Number of tetrahedral 12 8 4
interstitial sites
Size of the octahedral 0.19˜10-10 m 0.53˜10-10 m
interstitial sites
Size of the tetrahedral 0.36˜10-10 m 0.28˜10-10 m
interstitial sites
Number of the second 6 6
closest neighbors
Distance to the second a a
closest neighbors
Packing density 68 % 74 % 74%
Closed packed direction <111> <110> <11 2 0>
Closed packed plane {110} {111} (0002)
Atomic radius 1.24˜10-10 m 1.29˜10-10 m
Atomic distance 2.48˜10-10 m 2.57˜10-10 m
Theoretical density 7.88 g/cm3 7.69 g/cm3
Table 2.3: Characteristic properties of the two lattice types of iron.

26
2.2 Crystal Formation

2.2.2 The influence of foreign atoms on the lattice constant of iron


The Hume-Rothery rule (15%-rule) is a good approximation for the solubility of
foreign atoms in α-iron. It states that the difference between the space required for
the dissolved atoms and the space required for the atoms of the matrix crystal must
not be greater than 15%. The distance between neighboring atoms serves as a
measurement for the required space. The atomic volume of D-iron is small in
comparison to most other elements, nearly all alloying elements cause an increase in
the volume of the unit cell, with the exception of silicon and phosphorus. Figure
2.10 shows the influence of foreign atoms on the D-iron lattice constant.
0.2874
Zn, Al
W
0.2872 Mo
Ti V
Lattice constant a in 10 m
-9

0.2870
Cr
Cu Mn
Ni
0.2868

Co
0.2866

0.2864
Si
P
0.2862
0 2 4 6 8 10 12
Mass-content in %
Figure 2.10: Influence of alloying elements on the D-iron lattice constant at room
temperature.
Foreign atoms are of considerable importance for the crystallographic
transformational behavior of iron. They can either account for a delay or
acceleration of the DoJ transformation and therefore, for the temperature at which
the transformation occurs. The addition of large amounts of alloying elements, such
as chromium or nickel, can have the effect that a transformation does not occur over
very large temperature ranges.

2.2.3 Real Structures


Real crystals are not perfect; rather, they show certain defects. Table 2.4 and
Figure 2.11 show typical crystal structural defects of iron and steel, as well as the
dimension in which they are found in the lattice. Lattice defects regulate many
physical phenomena in metals. For example, the vacancy density is of great

27
2 The Physical Properties of Iron and Steel

importance for diffusion; dislocations are significant for plastic deformation, and
grain boundaries are important for many mechanical and physical properties.
Dimension Type of structural crystal Description
defect
0-dimensional point defect vacancy
interstitial foreign atom
substitutional foreign atom
1-dimensional linear defect dislocation
2-dimensional planar defect grain boundary
stacking fault
subgrain boundary
twin boundary
3-dimensional space defect inclusion
precipitate
intermetallic phase

Table 2.4: Dimensions, types, and descriptions of lattice defects.

Figure 2.11: Graphic representation of typical lattice defects.

28
2.3 Thermal Properties

2.3 Thermal Properties

2.3.1 Changes in volume and length of iron

The heating of a pure metal generally leads to a steady increase in the lattice
constant and, therefore, the volume of the unit cell increases, if no polymorphic
changes occur. At normal atmospheric pressure, pure iron undergoes two
transformations: the DoJ transformation (bccofcc) at 911°C (A3), and the JoG
transformation (fccobcc) at 1392°C (A4). With increasing temperature, the lattice
constant a is characterised by three curve segments, each with a different positive
slope (Figure 2.12). The difference in the slope of each curve segment is due to the
different expansion behaviors of the bcc and fcc crystal lattices.
0.295 0.373
for D- and G-iron in nm

0.293 G 0.371
Lattice constant a

Lattice constant a
for J-iron in nm
DFe JFe Fe
0.291 0.369

0.289 0.367

0.287 0.365
A3 A4
0.285 0.363
0 200 400 600 800 1000 1200 1400 1600 1800
Temperature in K
Figure 2.12: Change in the lattice constant a of pure iron with increasing
temperature.
The linear expansion coefficient D can be calculated as an average expansion
coefficient for each temperature range:

1 dl
D ˜ (2.16)
l dT

where l = length and T = temperature.

Equation 2.17 is the result of solving the previous equation for the length at a given
temperature:

l (T ) l 0 ˜ 1  D ˜ 'T (2.17)

29
2 The Physical Properties of Iron and Steel

In Table 2.5, the average thermal expansion coefficients for pure iron are listed for
different temperature ranges. The linear thermal expansion coefficient D is shown
as a function of temperature in Figure 2.13.
Lattice type Temperature range Average expansion
in °C coefficient D in 10-6K-1
D-Fe 20 – 911 15.3
J-Fe 911 – 1392 22.0
G-Fe 1392 – 1536 16.5
Table 2.5: Average thermal expansion coefficients for pure iron.

Figure 2.13: Thermal expansion of pure iron.


In the stable range for D-iron, the linear thermal expansion coefficient D rises as the
temperature increases. At the Curie temperature TC, an anomaly in form of a
decrease in the linear thermal expansion coefficient D occurs due to a change in the
magnetic properties from ferromagnetic to paramagnetic. After this, D begins to
rise again. In the stable range for J-iron, a constant expansion coefficient D of
22˜10-6 K-1 can be observed. Above the A4-point, in the stable range for G-iron, the
expansion coefficient D falls instantaneously to a value of approximately
16.5˜10-6 K-1. The dotted curves show the approximate theoretical progression of
the thermal expansion coefficient for fcc iron and paramagnetic bcc iron. At each
of the transformation points, the calculated curves join the measured curves.

30
2.3 Thermal Properties

The atomic volume changes according to the lattice constant with increasing or
decreasing temperatures. The following relationships exist between the lattice
constant a and the atomic volume:
a3
bcc lattice: Vat (2.18)
2
a3
fcc lattice: Vat (2.19)
4
Figure 2.14 shows the change in the atomic volume of pure iron according to
temperature. Analogous to the linear thermal expansion coefficient D, a cubic
thermal expansion coefficient J can be calculated as the volume-related differential
quotient of the change in volume, dV, and the difference in temperature, dT.
1 dV
J ˜ (2.20)
V dT
For isotropic materials, the cubic thermal expansion coefficient J correlates with the
linear thermal expansion coefficient D:
J 3 ˜D (2.21)
Therefore, the determination of the linear thermal expansion coefficient is sufficient
to describe the change in volume.

Figure 2.14: Change in the atomic volume of pure iron with temperature.

31
2 The Physical Properties of Iron and Steel

Between room temperature and the A3-temperature, the atomic volume increases by
approximately 4% and the relative length by about 1.3%. Between room
temperature and the melting point, the atomic volume increases by about 7.5% and
the relative length by approximately 2.6%. The transformation of D-iron (VR(bcc) =
68%) into the more closed packed J-iron (VR(fcc) = 74%) at the A3-temperature is
associated with a decrease in volume and length of about 1% and 0.35%,
respectively. The increase in volume at the A4-temperature is only 0.5%, because J-
iron has a larger thermal expansion coefficient than D- and G-iron.
Table 2.6 helps to clarify the temperature dependence of the lattice constant and
density of pure iron. It also shows the results calculated for the relative change in
length of a pure iron rod at different temperatures.
Temperature Structure Symbol Lattice constant Density Relative length
°C - - nm g/cm3 (calculated)
20 bcc D 0.2866 7.88 1
911 bcc D 0.2904 7.57 1.0132
911 fcc J 0.3646 7.65 1.0097
1392 fcc J 0.3688 7.40 1.0211
1392 bcc G 0.2932 7.36 1.0229
1536 bcc G 0.2941 7.29 1.0262
Table 2.6: Lattice constants and densities of pure iron dependent on temperature.
A pure iron rod with a length l0 of exactly one meter at room temperature increases
in length to 1.0132 m after heating to 911°C (A3-temperature). At the point of
transformation, an instantaneous shrinkage of 0.35% to 1.0097 m occurs. As the
temperature increases to 1392°C (A4-temperature), the rod expands to a length of
1.0211 m. Due to the reversal of the transformation, an instantaneous expansion of
0.18% to 1.0229 m occurs. At the melting point the rod has a length of 1.0262 m, a
total increase of 26 mm from the starting length at room temperature. The formulas
to calculate the relative length within the three different temperature ranges are
given in Table 2.7.
Temperature range in °C Equation
(T in °C)
20 – 911 l (T ) l0 ˜ >1  14.8 ˜106 ˜ T  20 @
911 – 1392 l (T ) l0 ˜1.0097 ˜ >1  23.6 ˜106 ˜ T  911 @
1392 – 1536 l (T ) l0 ˜1.0229 ˜ >1  22.3 ˜106 ˜ T  1392 @
Table 2.7: Calculating the total expansion of pure iron.

32
2.3 Thermal Properties

2.3.2 Changes in volume and length of steels


The thermal expansion of pure ferritic and ferritic-pearlitic steels, as well as
quenched and tempered (QT-) steels, is similar to the thermal expansion of D-iron.
At room temperature the linear thermal expansion coefficient D of these steels is
between 11˜10-6 and 13˜10-6 K-1, similar to that of D-iron. High alloyed austenitic
steels exhibit significantly larger linear thermal expansion coefficients. Figure 2.15
shows the dependence of the linear thermal expansion coefficients of ferritic and
austenitic steels on different temperatures.
In many magnetically ordered J-iron alloys, a volume-magnetostriction appears,
whereby, through the transformation of magnetic ordering, a change in the lattice
spacing occurs, thereby influencing thermal expansion. In alloys with a sufficiently
large positive volume-magnetostriction, the magnetic coupling can compensate for
the thermally dependent change in lattice spacing. These so called Invar alloys are
characterised by a small and almost constant thermal expansion over a certain
temperature range. Classic iron based Invar alloys contain about 36 mass-% nickel
and reach thermal expansion coefficients of 1˜10-6 K-1. Superinvars additionally
contain cobalt and reach thermal expansion coefficients of approximately 1˜10-7 K-1.
Thermal expansion coefficient Din 10-6 K -1

25
Austenitic steels
20

15
Ferritic steels
10

0 100 200 300 400 500 600 700 800


Temperature in K
Figure 2.15: Thermal expansion coefficients of ferritic and austenitic steels.

33
2 The Physical Properties of Iron and Steel

Case study: Invar steel

Figure 2.16: Application of Invar steel.


Figure 2.16 shows the applications of Invar steel.
For the purpose of demonstration, Figure 2.17 shows a comparison between the
thermal expansion coefficients of an Invar alloy and a common plain carbon steel.
Easily recognisable is the comparatively larger temperature range over which the
Invar alloy containing 36 mass-% nickel shows little thermal expansion.
Figure 2.18 shows the transition temperatures of the iron-nickle system.
Paramagnetic iron-manganese-nickel alloys and iron-chromium-nickel alloys show
a thermal expansion behavior contrary to that of Invar alloys. The temperature-
dependent change in thermal expansion for each of these alloys is characterised by a
maximum. These maxima result from the transition of atoms to a higher energy
state, which increases their atomic volume. The maxima also explain the higher
expansion coefficient of austenitic chromium-nickel alloys compared to ferritic
steels.
As a result of this anomaly, face-centered cubic alloys exhibit a relatively large
thermal expansion at room temperature. Therefore, alloys with a large expansion,
for example, 75% Fe, 19% Ni, 6% Mn, can be used as active components in bi-
metallic systems. In Figure 2.19, the thermal expansion coefficient D is graphed
against the number of valence electrons per atom, e/a. A maximum of
approximately 19˜10-6 K-1 appears at an e/a ratio of 8.3. Alloys with either smaller

34
2.3 Thermal Properties

or larger numbers of valence electrons have a smaller thermal expansion at room


temperature.

Figure 2.17: Linear thermal expansion coefficients of an Invar steel (36% Ni) and a
plain carbon steel.

Figure 2.18: Transition temperatures of the nickel-iron system

35
2 The Physical Properties of Iron and Steel

Figure 2.19: Thermal expansion of various fcc alloy groups at 300 K.

2.3.3 Thermal conductivity


Temperature differences in a solid object lead to a flow of heat, which is, above all,
dependent on the thermal conductivity of the material. Generally, good electrical
conductors also serve as good thermal conductors.
Thermal conduction can function on the basis of two fundamental mechanisms; it
can occur through oscillations in the lattice, or over free electrons (conducting
electrons). The temperature dependence of thermal conductivity O is represented in
Figure 2.20.

36
2.3 Thermal Properties

Figure 2.20: Thermal conductivity of pure iron dependent on temperature.


This temperature dependence is a reflection of the impact of ferromagnetic
properties of iron. A large increase in thermal conductivity is associated with the
appearance of magnetic ordering below the Curie temperature TC.
Since conducting electrons are essentially responsible for thermal conduction at
conditions above room temperature, the correlation between thermal conductivity
and electrical conductivity can be described by the Wiedemann-Franz-Lorenz Law,
which says that the ratio of thermal conductivity O to electrical conductivity V
changes proportionally to temperature T.
O V ˜ L ˜T (2.22)
The Lorenz number L has nearly the same value for all metals and can be
determined as follows:
k2
L # 3˜ (2.23)
e2
where k = Boltzmann constant and e = (element) charge.
Above room temperature, the Lorenz number L for pure iron is 3.03 . 10-8 V2 K-1.
At lower temperatures, the Lorenz constant also becomes temperature-dependent.
In addition to electron conductance, the amount of heat transported through lattice
oscillations takes on importance with decreasing temperature. The combination of
the effects of these two mechanisms and the amount of remaining resistance results
in a maximum in thermal conductivity reached at low temperatures, as seen in the
inserted graph in the upper right-hand corner of Figure 2.20. The maximum of the
thermal conductivity increases with the purity of the crystal. Crystal lattice defects
hinder both the electrical conductivity and thermal conductivity.

37
2 The Physical Properties of Iron and Steel

2.3.4 Diffusion
Diffusion is identified by a thermally activated transport of atoms in solid and of
atoms or molecules in liquid or gas. The diffusional motion of iron atoms in an iron
lattice is called self-diffusion; while the diffusional motion of foreign atoms within
the atom matrix is called interstitial or substitutional diffusion.
The diffusion flux density I (diffusion flux over the cross area A) depends on the
number of the moving atoms n in a time period t and is proportional to the
concentration gradient dc/dx. The relation is known as First Law of Diffusion:
1 dn dc
I ˜ D ˜ (2.24)
A dt dx
The diffusion constant D has the dimension of m2/s and is described by means of an
Arrhenius equation:
Q

D D0 ˜ e RT
(2.25)
with Q = activation energy, R = gas constant, T = absolute temperature, and D 0 =
frequency factor.
The temperature dependence is a linear Arrhenius plot logD versus 1/T.
The Second Law of Diffusion describes the concentration change over time as a
function of concentration gradient:
dc d 2c
D ˜ (2.26)
dt dx 2
A specific solution of the differential equation for the one-dimensional diffusion is
usually applied to demonstrate the mean diffusion range as a function of
temperature and time:
X 2Dt (2.27)
Important influencing factors of atom diffusibility are summarized in Figure 2.21.

38
2.3 Thermal Properties

Figure 2.21: Influencing factors on the diffusion coefficient.


Diffusion parameters for the interstitially dissolved C in iron lattices and the self-
diffusion parameters for iron are listed in Table 2.8. It becomes apparent that the
interstitial atoms can diffuse faster than the substitutional atoms due to the low
activation energy. The diffusion rate in body-centered cubic lattice is significantly
higher than that in face-centered cubic lattice at the same temperature because of the
higher frequency factor. The high frequency factor can be explained by the higher
number of the octahedral interstices/interstitial sites and lower pack density in bcc
lattice.

Table 2.8: Diffusion parameters.


The Arrhenius relations of diffusion coefficients for the interstitially dissolved
atoms H, C and N and the substitutionally dissolved atoms in bcc iron lattice are
shown in Figure 2.22.
Especially revealing in this diagram is that the axis label on the right shows the
required time for a diffusion range of 20 μm. The range can be considered as a
mean distance to a grain boundary; thus it provides a measurement to evaluate if a
relevant parameter change can be counted on as a result of diffusion. This can

39
2 The Physical Properties of Iron and Steel

occur for instance in the formation of segregation on grain boundaries. It can be


inferred from the diagram that the diffusion of hydrogen in pure iron is hardly
obstructed. The diffusion of C and N over a distance of 20 μm is effected within
several days at slightly elevated temperatures of e.g. 100°C, so that the change of
microstructure and properties must be anticipated. The time-dependent phenomena
are called aging. For the substitutionally dissolved atoms the diffusion below
approx. 550°C can be neglected in a technically relevant period.

Figure 2.22: Diffusion coefficient of interstitially and substitutionally dissolved


elements in bcc and fcc iron.

40
2.4 Elastic Properties

2.4 Elastic Properties

2.4.1 Elastic modulus (Young’s modulus) and shear modulus


When a mechanical force acts on an object, a resistance force is generated in the
object, such that the internal and external forces on the object are in equilibrium.
The resistance of the object is called stress (in units of area). The kind of stress is
defined by forces perpendicular to the surface of interest (normal stress V), and
forces acting within the surface of interest (shear stress W).
Normal stresses cause a relative change in length H (elongation) of an object. In the
case of tensile normal stress, an elongation of the object occurs, while normal stress
applied as pressure results in a compression of the object. Shear stress results in a
shearing of the object at an angle J.
If the object takes on its original shape again after the stress is removed, the object
is said to be elastic. If the object does not regain its original shape after the stress
has been removed, the object is said to behave plastically. Note that an elastic
change in an object is associated with a reversible change in volume, whereas for
plastic deformation the law of constant volume applies.
The increase in volume of an object under elastic tensile stress should be made clear
by the following example. Tensile stress is applied to a round rod with an initial
volume of V0. The final volume V1 can be determined with the following
calculations. The initial volume V0 of a round rod with a diameter d0 and a length l0
is:
S 2
V0 ˜ d 0 ˜ l0 (2.28)
4
The applied elastic tensile stress results in a volume V1, giving consideration to the
increase in length and decrease in diameter:
S
˜ d 0  'd ˜ l 0  'l
2
V1 (2.29)
4
Ignoring all terms in which 'd and 'l are squared or multiplied with one other, and
substituting Vo and H results in the following equation:
§ 'd 'l ·
V1 V0 ˜ ¨¨1  2  ¸ V0 ˜ 1  2 ˜ PH  H (2.30)
© d 0 l 0 ¸¹
'd l 0
with Poisson’s number: P ˜ (2.31)
d 0 'l
'l
and the elongation: H (2.32)
l0
Therefore, the change in volume caused by tensile stress is:

41
2 The Physical Properties of Iron and Steel

'V V1  V0 V0 ˜ H ˜ 1  2 ˜ P (2.33)
With μFe = 0.235, the increase in volume of an iron object under elastic tensile stress
can be calculated.
In the elastic zone, many materials show a linear relationship between stress and
elastic deformation. Hooke’s Law describes this relationship:
V H ˜E for normal stress (2.34)
W J ˜G for shear stress (2.35)

<100>-direction
(cube edge):
y y E = 135 GPa
G = 61 GPa
x x
(100) -- Fläche
(100) plane (110)
(110)--Fläche
plane <110>-direction
(plane diagonal):
z
z
E = 217 GPa
G = 83 GPa

[101] [111]
<111>-direction
(cell diagonal):
y y
E = 290 GPa
G = 118 GPa
x [100] [110]
(111)
(111) -- Fläche
plane x
Gitterrichtungen
lattice orientation
iminkubischen
the cubic cell
Gitter

Figure 2.23: Illustration of selected directions and planes of the cubic lattice, as
well as Elastic modulus and Shear modulus dependent on the crystal
direction in an iron single crystal.
The constants in the previous formulas are the Elastic modulus (Young’s modulus)
E, and the Shear modulus G. However, these basic formulae of elastic theory are
only valid for isotropic materials. The anisotropy of crystals expresses itself in
mechanical behavior. This influence is characterised by different deformations in
different crystallographic directions, when a stress is applied. This is due to the fact
that elastic constants are dependent on crystallographic orientation. Values for the

42
2.4 Elastic Properties

Elastic modulus E and the Shear modulus G for various directions in iron single
crystals are given in Figure 2.23.

Provided that a metal is not a single crystal, but rather polycrystalline, it consists of
a multitude of randomly oriented crystals such that on the macroscopic scale the
elastic properties are independent of the orientation of the crystals. As a result of
this quasi-isotropy, the elastic behavior can be described by average Elastic-, and
Shear modulus. These average modules for polycrystalline iron are E = 210 GPa
and G = 83 GPa.
In Figure 2.24, the elastic modules of various elements are plotted against their
melting points. It is shown thereby that the Elastic modulus increases with rising
melting point. For this reason, aluminum, for example, exhibits a lower melting
point and Elastic modulus than iron.
400
Elastic modulus E in GPa

W
Mo

Fe Cr
200 Co
Ni Ta
U Pt
Mn
Si Pd V
Zn Cu
100 Ti
Sb Zr
Au
Se Al Ce Ag
Sn Cd Mg
Bi Pb Ca
0 1000 2000 3000 4000
Melting point in K
Figure 2.24: Elastic modulus for various elements dependent on melting point.

The Elastic modulus E (Figure 2.25) and the Shear modulus G of polycrystalline
iron progressively decrease with increasing temperature. The E-modulus is to a
great extent independent of the present grain size. Furthermore, it decreases,
especially around the Curie temperature as a result of the break up of magnetic
coupling. The DoJ transformation is associated with an increase in the E-modulus
(Figure 2.26). Nonetheless, the E-modulus for fcc iron is approximately 10% lower
than that for bcc iron at room temperature. This is the result of the different
temperature dependence of each phase, due to a difference in magnetic properties.

43
2 The Physical Properties of Iron and Steel

220

210

Elastic modulus E in GPa


200

190

180

170

160
0 100 200 300 400 500 600
Temperature in °C

Figure 2.25: Elastic modulus E for plain carbon, and low-alloyed steels dependent
on temperature.
220

200
Elastic modulus E in GPa

180

160

140
A3
120
TC
100
0 200 400 600 800 1000 1200 1400
Temperature in K
Figure 2.26: Temperature dependence of the Elastic modulus E for iron.

44
2.4 Elastic Properties

Another significant influence on the values of Elastic modulus E and Shear modulus
G is the concentration of alloying elements. The alloying elements are divided into
those that increase, and those that decrease the modulus. The moduli of the
individual alloying elements are of importance for this classification. Alloying
elements with larger modulus than iron raise the modulus of the iron-alloys when
they are in solid form. The effect of dissolved atoms on the Elastic modulus E and
the Shear modulus G is, to a great extent, the same.

The Elastic modulus E for various binary iron alloys is represented in Figure 2.27
as a function of chemical composition. The iron-silicon alloy curve is drawn with a
dotted line because the data was derived from measurements on a single crystal.
The Elastic modulus E changes proportionally to the concentration of dissolved
alloying elements. Carbon and cobalt are exceptions to this linear relationship. The
shear modulus G shows an analogous dependence on the chemical composition.
The influence of prior deformation on Elastic modulus E for a deep drawing steel is
graphed in Figure 2.28. This deep drawing steel exhibits a decrease in the
E-modulus with increasing elongation. However, an influence of the method used
to measure the E-modulus on the results can be recognized. The results from tensile
tests show E-modulus values that are lower and have a larger overall decrease than
the results from an ultrasonic measurement. The decrease in the E-modulus
associated with prior deformation is the result of an increase in the defect density of
the metal. This defect density can be reduced by an annealing treatment so that the
E-modulus returns to its initial value.
220
Re Co

Cr
Elastic modulus E in GPa

210
Ir
Ru
Si

Mn
200

C Ni
Rh
Pt
190
0 2 4 6 8 10 12
Alloying elements in atom-%
Figure 2.27: Influence of alloying elements dissolved in polycrystalline iron on the
Elastic modulus E at room temperature.

45
2 The Physical Properties of Iron and Steel

220
Ultrasonic measurement

Elastic modulus E in GPa


210

200

Tensile test
190

180
0 2 4 6 8 10 12 14 16
Elongation in %
Figure 2.28: Influence of prior deformation on Elastic modulus E of a deep
drawing steel.

2.4.2 Anelasticity
For ideal elastic behavior of an object, it is assumed that the Hooke’s Law applies,
and all external stresses V below the yield stress produce a certain elongation H.
The duration of the applied stress should not have any effect.
However, in reality, an elastic after-effect is often observed, whereby a time-
dependent elastic deformation occurs in addition to the elastic expansion according
to Hooke’s Law. If this additional, time-dependent elongation recedes over time
once the stress has been removed, so that the body returns to its original shape after
a certain time, then one speaks of anelastic behavior or anelasticity.
If an anelastic object is subjected to a stress V smaller than the yield stress over a
long period, then a pure elastic elongation H1 occurs spontaneously in the object.
This is followed by a time-dependent, anelastic elongation H2(t), which after a very
long period, can reach the limiting value H2 (Figure 2.29). The amount of anelastic
elongation H2(t) is, therefore, exponentially dependent on time. The pure elastic
elongation spontaneously returns after removing the stress. After some time, the
additional, time-dependent elongation recedes, and the object returns to its original
shape.

46
2.4 Elastic Properties

Stress Elongation

Stress, elongation
H2

H1

Time
Figure 2.29: Schematic description of the elastic after-effect.
A practical example of anelastic elongation is damping, which is associated with a
loss of energy as a result of mechanical oscillation. Damping is characterised by a
logarithmic decrement /, where a1 and a2 are the amplitudes of two consecutive
oscillations:
§a ·
/ ln ¨¨ 1 ¸¸ (2.36)
© a2 ¹
Besides the external share of energy loss through, for example, external friction, the
loss of energy due to oscillation is attributed to internal friction resulting from
material specific effects occurring additionally to the interactions with the
surroundings. The magnitude of this energy loss during oscillation is influenced by
both frequency and temperature.
For the experimental set-up, shown in Figure 2.30, a torsional pendulum is used to
excite a wire or a small sheet to torsional oscillations. In the torsional pendulum, a
periodically changing elastic deformation appears. The elastic after-effect leads to a
time-dependent phase-displacement of the stress and elongation maxima. The
maximum elongation is only reached after a certain period of relaxation, and its
value is dependent on the processes occurring within the material. From this phase
displacement, it can be concluded that damping results from the time-dependent
processes occurring within a material. One of these processes is the positional
change of the interstitially dissolved foreign atoms in the bcc lattice. The carbon
and/or nitrogen dissolved in the interstitial sites of D-iron change their positions in
the unit cell periodically according to applied stress.

47
2 The Physical Properties of Iron and Steel

Wire

Weighted rod

Wire

Furnace
Sample M

Figure 2.30: Basic design of a torsional pendulum.


In the mechanically unstressed material, all octahedral interstitial sites in the bcc
lattice (at the center of the edges and surfaces) are energetically equal so that the
interstitially dissolved carbon and/or nitrogen atoms are dispersed randomly
(Figure 2.31, left). Since the atomic radii of carbon and nitrogen atoms are larger
than the radii of the octahedral interstitial sites in D-iron, when these interstitial
sites are filled, the lattice structure is stretched in the direction of the octahedral
axes.
As a result of mechanical stress, the distances between atoms in the crystal lattice
increase in the tensile direction. Due to this extension, the interstitial sites present
in the relevant plane show energetically favorable conditions for the inclusion of
foreign atoms (Figure 2.31, right). On the other hand, compressed regions show
unfavorable conditions for inclusions. Therefore, a rearrangement of the
interstitially dissolved foreign atoms in the direction of the more energetically
favorable positions occurs. This rearrangement is dependent on time, and affects a
further expansion of the crystal lattice in the tensile direction.
If the direction of applied stress is changed during oscillation, the previously
energetically favorable interstitial sites become energetically unfavorable, effecting
the movement of dissolved foreign atoms to the energetically favorable octahedral
interstitial sites in the new tensile direction. When all external stresses are
removed, the interstitially dissolved atoms disperse themselves randomly once

48
2.4 Elastic Properties

again. The energy for these positional changes is subtracted from the oscillation
energy, and can be seen macroscopically as a dampening of the oscillation. This
process is called the Snoek effect, after its discoverer.
The relaxation time for the process of positional change is dependent on
temperature, and varies with different types of atoms. For this reason, every type of
atom is characterised by a damping maximum from a damping spectrum taken at a
constant frequency, dependent on temperature. This damping spectrum is
independent of temperature at a constant frequency. Therefore, it is possible to
quantifiably demonstrate the presence of interstitially dissolved foreign atoms such
as, carbon and nitrogen, in the bcc lattice of iron.

Figure 2.31: The Snoek effect: schematic representation of carbon-atom


distribution in the bcc lattice in an unstressed state (left) and under
tensile stress (right).

49
2 The Physical Properties of Iron and Steel

2.5 Magnetic and Electric Properties

2.5.1 Magnetic properties of pure iron


There are four main types of magnetism in solid bodies: diamagnetism,
paramagnetism, ferrimagnetism, and ferromagnetism. As a rule, an applied
magnetic field H induces an electrical current in the electron shell of a material that
results in a magnetic field B, also called magnetic induction. The size of this
magnetic field is dependent on the induction constant μ0 and the relative
permeability μr of the material. The following applies:
B P0 ˜ Pr ˜ H (2.37)
Vs
with P 0 4S ˜ 10 7
Am
in a vacuum:
B P0 ˜ H (2.38)
V ˜s A
with B given in Tesla ( T ) and H in .
m2 m
)
>> 1
(μ r

H H
tism

)
> >1
ne

μr
i sm(
ma g
Magnetic flux densitity B

et
gn
r ima
Ferr o

r
Fe
H
> 1)
etis m (μ r
Par amagn
Vacuum
μ0H
Diamagnetism (μ < 1)
r
H

Figure 2.32: Schematic representation of magnetic atomic moments.

50
2.5 Magnetic and electric properties

The orientation of the magnetic moments in the atomic shells is decisive for the type
of magnetism. This is summarised in Figure 2.32.

Diamagnetism
For diamagnetism, the resulting magnetic field, and therewith the magnetic atomic
moments, is oriented in the direction opposite to that of the applied magnetic field.
The relative permeability is μr < 1. Diamagnetism can be observed in all materials,
including the noble gases and non-metals. However, diamagnetic properties are
insignificant in paramagnetic and ferromagnetic materials.

Paramagnetism
In paramagnetic materials, the magnetic moments in the atomic shells are randomly
orientated. As a result, the average magnetic induction is zero. The relative
permeability μr is > 1. If a paramagnetic material is introduced into a magnetic
field, the magnetic moments orient themselves in the direction of the applied field.
In weak fields, the magnetic induction is proportional to the applied field. Stronger
applied fields lead, over time, to magnetic saturation, whereby all of the magnetic
moments orient themselves in the direction of the field. Increasing temperatures
reduce the induction of magnetic saturation in paramagnets. Many of the non-noble
gases, as well as all of the ferromagnetic and ferrimagnetic materials, are
paramagnetic at temperatures above their Curie temperature.

Ferrimagnetism and Ferromagnetism


Strong interactions between magnetic moments can result in an ordering of these
moments in the atomic shell. This is described as ferromagnetism if the moments
are arranged parallel to one another. If an anti-parallel ordering occurs, it is called
ferrimagnetism or antiferromagnetism. In ferrimagnetic materials, the magnetic
moments compensate for one another at an atomic scale, resulting in an average
magnetic induction of zero. In contrast, in the atomic shells of ferromagnetic
materials, a magnetic moment exists without the presence of an external magnetic
field. The relative permeability of ferrimagnetic and ferromagnetic materials is
μr >>1 and ranges between 103 and 106. Other ferromagnetic elements and materials
besides iron, are the elements cobalt and nickel, the rare earth elements Gadolinium
(Gd) and Dysprosium (Dy), and the Heusler alloys (e.g.: Cu2AlMn).
Ferromagnetic materials are spontaneously magnetic up to the point of saturation
without any external forces, because they have internal magnetic fields. When all
of the magnetic moments are oriented in the same direction, a saturation induction
BS exists. BS decreases with increasing temperature because the magnetic ordering
decreases due to lattice oscillations. The temperature dependence of magnetic
induction is graphed in Figure 2.33. Spontaneous magnetism becomes zero at the
Curie temperature TC. Ferromagnetic and antiferromagnetic coupling breaks up at

51
2 The Physical Properties of Iron and Steel

the Curie temperature due to the thermal energy of the crystal lattice. When T > T C,
a previously ferromagnetic material is solely paramagnetic, since spontaneous
magnetism is no longer present. The Curie temperature of pure iron is 769°C
(1042 K).
1.0

0.8

0.6
8
BS/B

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1.0 1.2
T/TC

Figure 2.33: Temperature dependence of spontaneous magnetisation in D-iron with


TC = Curie temperature = 1041 K, BS = saturation induction, and
Bf = induction at infinitely large field strengths.
A ferromagnetic material can appear non-magnetic because the material is divisible
into many microscopically small areas, known as white zones, or magnetic domains,
each of which is independently spontaneously magnetic. These spontaneous
magnetic orientations are randomly distributed, so that they cancel one another.
The resulting non-magnetic state is shown in state 1 of Figure 2.34.
If an external magnetic field is introduced to a material, and its field strength H is
steadily increased, the magnetic induction B or the magnetisation M in the material
begins to follow the initial magnetisation curve (Figure 2.34 and Figure 2.35) until
it reaches its saturation point BS for the external field strength HS. With very small
external field strengths, a reversible displacement of the borders of neighboring
magnetic domains (Bloch walls) occurs. This leads to a growth of the favorably
oriented magnetic areas (1) in comparison to the external magnetic field.
Furthermore, the atomic magnetic moments within the white zones change their
direction to match the main direction of magnetisation. Finally, growth in the
magnetic domains is expressed by irreversible Bloch wall displacements and sudden
jumps in the magnetisation as the magnetic moments of small groups of atoms
become aligned, which is called the Barkhausen effect (2). When the polarisation
of saturation is reached, the preferred magnetic direction equals to that of the

52
2.5 Magnetic and electric properties

external magnetic field (3). When the external magnetic field is removed, the
magnetic induction does not drop back to zero. Instead, a residual induction, the
remnant polarisation BR, remains (4). This is because only reversible processes can
return their conditions to the original state after the external magnetic field has been
removed; irreversible processes create a remnant. In order to return the magnetic
induction to zero, an opposing field with a strength equal to the coercive field
strength, HC, must be applied (5).
The larger HC is, the larger the area of hysteresis will be. In an alternating magnetic
field, irreversible processes lead to hysteresis. As the area of hysteresis increases,
the magnetic hysteresis loss per cycle also increases, and the material becomes
“magnetically harder.” There are additional losses due to eddy currents at very high
frequencies. The eddy currents are dependent on electrical conductivity.
It is shown in Figure 2.36 that the magnetisation process of pure iron is strongly
dependent on the crystal orientation; it is strongly anisotropic. A magnetisation in
the <100>-direction can occur very quickly with a steep rise at the beginning, and
rapidly reaches saturation induction. This occurs because iron is spontaneously
magnetized in the <100>-direction. A more gradual rise in magnetic induction is
observed if the applied magnetic field lies in the <110>- or <111>-direction. The
saturation magnetisation is only reached at very large field strengths. This is due to
the fact that spontaneously magnetized regions can orient themselves with the
nearest edge of the cube without great resistance, while a considerably higher
amount of energy is necessary to orient these regions in the direction of the applied
magnetic field. When the external magnetic field is removed, the magnetic
moments reorient themselves in the direction of the nearest edge, but do not resume
their original orientations. The anisotropy of the magnetisation of iron is utilized in
industry, as an example, for electrical sheets by developing suitable textures.

53
2 The Physical Properties of Iron and Steel

4 B 3
BS 3
4 Initial
magnetization
2 curve

5 2

5 1
H
-HC HC HS
6
1

-Br
6

Figure 2.34: Magnetisation curve of a ferromagnetic material with H = Field


strength of the applied magnetic field, B = Magnetic flux density or
magnetic induction, Bs = Saturation induction, Br = Remnant, HC =
Coercivity, and HS = Field strength at the saturation induction.

Figure 2.35: Explanation of the course of the initial magnetisation curve during
magnetisation.

54
2.5 Magnetic and electric properties

2.5

[100] Bs = 2.15 T

Magnetic induction B in T
2.0 [110]

[111]
1.5

1.0
0 100 200 300 400 500 600 700
-1
Field strength H in Am
Figure 2.36: Anisotropy of the magnetisation of iron.
Case study: Electrical sheet

Figure 2.37: Stacking of cold-rolled electrical sheets as “E-I” construction.


The electrical sheet is the most important soft magnetic material on a value and
quantitative basis with an annual production of approx. 10 million tons worldwide.
According to their properties electrical sheets can be divided into grain-oriented and
non-grain-oriented material. More specifically, the cold-welded strips made from
iron-silicon-alloys are called electrical sheets, from which electrical circuits are
produced to manufacture electrical machines like transformers, relays and inductors
etc.
The most well-known structure is the “E-I” steel as shown in Figure 2.37, which is
used in transformers.

55
2 The Physical Properties of Iron and Steel

2.5.2 Magnetic properties of steels


The magnetic properties of D-iron solid solutions are influenced in varying degrees
by different alloying elements. For example, the Curie temperature rises with
increasing amounts of vanadium and the ferromagnetic element cobalt. However,
with increasing amounts of Ti, Mn, Si, Al, and Ni, the Curie temperature decreases,
whereby the largest decrease is caused by Ti. Up to a concentration of
approximately 5 mass-%, chromium raises the TC, while concentrations above this
level decrease the TC (Figure 2.38).
1150
Co
Curie temperature TC in K

1100
V

1050
Ni
Cr
Al
1000
Mn
Si

950 Ti
0 2 4 6 8 10 12
Alloying content in mass-%
Figure 2.38: Influence of alloying elements on Curie temperature.
The ferritic iron-silicon alloys are of great importance for technical use. The main
areas of application are magnetic iron cores in electrical machines, transformers,
and electric sheets, as well as use in heavy current engineering. The influence of
silicon on the magnetic properties of iron is shown in Figure 2.39.
Iron-silicon alloys are magnetically soft materials. They possess a small coercive
field strength during changes in magnetisation, a small saturation induction, and a
steep slope in the initial magnetisation curve (see Figure 2.35). This means they
have a high relative permeability and, therefore, a small range of hysteresis.
Furthermore, silicon reduces the crystalline anisotropy. At Si concentrations of
>2.2 mass-%, no J-D transformation occurs. This allows for the production of a
coarse grained, more failure-free, single-phase microstructure. As a result,
displacing the Bloch walls is easier, resulting in low magnetic hysteresis losses in
Fe-Si alloys. The saturation magnetostriction approaches zero at a mass content of

56
2.5 Magnetic and electric properties

about 6% silicon, so that the otherwise present periodic changes in length of the
sample in an alternating field are minimised.
Through deformation processes, such as rolling, a favorable texture, or crystal
orientation, can be set in the microstructure of a material so that the direction of
magnetisation most easily attained is that of the applied field. The optimal texture
for electric sheets in cubic materials is in the {100}-plane, and <001>-direction.
However, in practice, this is only possible with a coarse grained and coarse Bloch
wall structure. Furthermore, the extreme anisotropy of this crystal orientation is
undesirable. The Goss texture {110}<001> is technologically possible.
Conventionally, this texture is used for the production of grain oriented electrical
sheets. Table 2.9 shows the important characteristic hysteresis values for different
Fe-Si alloys in comparison to other materials.

Figure 2.39: Influence of silicon on the magnetic properties of iron.

57
2 The Physical Properties of Iron and Steel

Material Relative Saturation flux Coercive field


permeability μr density Bs in T strength Hc in Am-1
D-Fe 5000 2.14 72
Fe- 3% Si 8000 2.01 56
Fe- 3% Si
50000 2.01 7.2
(grain oriented)
Supermalloy
800000 0.8 0.5
(Ni79Fe16Mo5)
J-Fe (800 °C) 1.15
Table 2.9: Characteristic hysteresis data for various materials.
The addition of silicon greatly increases the electrical resistance (Figure 2.40),
which can also cause a reduction in the eddy current losses (ecl). This is especially
desirable for high-frequency applications. The following formula applies to the
relationship between eddy current losses, specific electric resistance, frequency, and
sheet thickness:
1 d2
ecl v ˜f2˜ (2.39)
U sp U
where Usp = specific electrical resistance, f = frequency, d = sheet thickness, and U =
material density.
Relationship 2.39 shows that increasing the specific electric resistance, and
decreasing sheet thickness can reduce eddy current losses. The specific electric
resistance of a Fe-3% Si alloy is 30 μ:cm at room temperature, while the resistance
of pure D-iron is only 10 μ:cm. Very thin sheets would be desirable; however,
technological limitations restrict the reduction of sheet thickness. In practice,
electric sheets should not be thinner than 0.22 mm; otherwise, rolling produces
unfavorable textures.
While increasing the mass content of silicon in iron alloys has many advantages, it
also results in a great increase in handling difficulties during forming processes.
Hot rolled FeSi sheets cannot have more than 4.5 mass-% silicon, and cold rolled
FeSi sheets no more than 3.5 mass-% silicon. Otherwise, non-deformable
superstructures form in the crystal lattice.
Another application for magnetic soft iron alloys is as magnetic cores in induction
coils. A high initial permeability, and therefore, a steep slope in the initial
magnetisation curve is required; since the currents during the switching operations
are very small, the Bloch walls have to be very mobile. Some well-suited materials
are, for example, the subordinate solid solutions FeNi3 (Permalloy) or Supermalloy.
Normally, these are magnetized in the <111>-diagonal of the fcc lattice.

58
2.5 Magnetic and electric properties

Specific electrical resistance Usp in μ:cm


100

Si
80
Al
60
Mn
40 C
Cr
20
Ni

0
0 2 4 6 8 10 12
Alloying element content in mass-%
Figure 2.40: Increase in resistance of D-iron caused by alloying elements.

2.5.3 Electric properties


Pure iron
The electrical conductivity of metals is based on the mobility of the charge-carriers
(electrons) in the electric field. The mobility of the electrons is hindered by lattice
oscillations (phonons) and lattice defects (e.g. foreign atoms). The greater the
increase in lattice oscillations, and therefore, the greater the number of defects, the
more the mobility of the electrons is reduced. The measure of the restriction of
conductance is the specific electrical resistance Usp. The electrical resistance is
divided into a temperature-dependent component UPh, and a constant component U0,
the remaining resistance. UPh is explained by phonons that are excited through
lattice oscillations and has a nearly linear temperature dependence. At high
temperatures lattice oscillations almost exclusively hinder the electrons, while the
remaining resistance becomes negligible. At low temperatures, the lattices are
essentially frozen, and UPh | 0. The remaining resistance becomes dominant at low
temperatures and increases more strongly as the number of defects in a crystal
increases. In a superconductor, the electrical resistance decreases to zero when
falling below the transition temperature (Figure 2.41). The Matthiesen’s rule
describes specific resistance as the temperature-dependent quantity Usp.
U sp U 0  U Ph (T ) (2.40)

59
2 The Physical Properties of Iron and Steel

where U0 = remaining resistance and UPh = temperature-dependent amount.


Normal conductor
ρ=f(T) Superconductor

ρ Linear trend
Linear in the liquid phase
zone
Residual
resistance
Non-linear rising
until melting point

0K Transition T[K]
Melting point
temperature

Figure 2.41: Schematic representation of the electrical resistance of a normal and


superconductor.
The specific resistance of iron at room temperature is 0.098˜10-6 :m. The electrical
conductivity (V) is the reciprocal of the temperature-dependent specific resistance.
As was mentioned previously in Chapter 2.3.3, a linear correlation exists between
thermal conductivity O and electrical conductivity V according to Wiedemann-
Franz’s Law:
O V ˜ L ˜T (2.41)
As a first approximation, the temperature dependence of electrical resistance can
also be described in the following form:
U sp U 0 ˜ (1  D ˜ T ) (2.42)

where D = linear thermal expansion coefficient.


Steels
As indicated in Chapter 2.3.3, the magnetic transformation from ferromagnetism to
paramagnetism results in a higher electrical resistance because of the subordinate
state of the electron spin. In fcc crystals, the increase in resistance is clearly
reduced with increasing temperature.
Alloying elements interfere with the crystallographic ordering of iron, and lead, in
principal, to an increase in the (remaining) electrical resistance at room temperature
(Figure 2.40). Polyvalent metals, those with an incomplete outer shell, and carbon

60
2.5 Magnetic and electric properties

cause a larger increase in resistance than the transitional metals nickel and
chromium.
Increasing the temperature also raises the electrical resistance of steels (Figure
2.42). The increase in electric resistance with temperature is similar for an
equivalent crystal structure which is shown for pure iron and ferritic steels. The
change in specific electric resistance for austenitic steels, on the other hand,
deviates greatly from these curves until the transformation temperature is reached.
Austenitic steels already have a relatively high electrical resistance at low
temperatures, which only moderately increases with rising temperature. Bcc alloys,
however, have a smaller resistance at low temperatures, which then exhibits a
greater increase with rising temperature. As expected, the shape of the resistance-
temperature curve after the transformation from the bcc to the fcc structure is very
similar for all steels.
Specific electrical resistance Usp in μ: cm

160

rising alloy content


120 4

3
80
2
1
40

0 200 400 600 800 1000 1200


Temperature in °C
Figure 2.42: Electrical resistance of ferritic and austenitic steels with 1 = pure iron,
2 = plain carbon ferritic steel, 3 = alloyed ferritic steel, 4 = austenitic
steel.

61
2 The Physical Properties of Iron and Steel

2.6 Further Readings


Béranger, G.; Henry, G.; Sanz, G.:
The Book of Steel
1st ed., Intercept, Andover, 1996

Harvey, P.D.:
Engineering Properties of Steel
ASM International, 1982

Leslie, W. C.:
Materials Science and Engineering Series
The Physical Metallurgy of Steels
McGraw-Hill Book Company, New York, 1981

Verein Deutscher Eisenhüttenleute (Ed.):


Steel
Volume 1: Fundamentals
Volume 2: Applications
Verlag Stahleisen, Düsseldorf
Springer Verlag, Berlin, Heidelberg, New York, Tokyo, 1992/1993

62
3.1 Formation of Alloys

3 Iron Alloys
3.1 Formation of Alloys
Metals are rarely used in their pure form. More often, one or more metallic or non-
metallic elements are added to change the end usage properties and processing
properties, such as strength, toughness, ductility, and resistance to corrosion.
The selective addition of elements is known as alloying. Alloys are produced by
mixing at least two metals, or a metal with a non-metal that shows metallic
properties in its liquid phase. Alloys can also be formed from sintering metallic
powders, as is common practice in powder metallurgy. The elements that form an
alloy are known as components. In steel production, elements other than the
selectively added components are often present in the material. These unwanted
elements, in order to distinguish them from the components of an alloy, are called
tramp elements. Some elements, such as copper or sulfur, can appear as either
alloying or tramp elements. The totality of all possible mass-% combinations of an
alloy is referred to as an alloy system; the phases present in the system can be
graphed over a temperature range as a phase diagram.
The following parameters are decisive in the formation of an alloy:
x The lattice type of the base and alloying elements,
x The size of the foreign atoms in comparison to the unit cell atoms,
x The number of valence electrons of the alloying element,
x The temperature, and
x The pressure.
Figure 3.1 gives an overview of the possibilities for alloy formation. The main
distinguishing factor in the classification of alloy formations is the nature of the
alloy, i.e. homogeneous or heterogeneous. Homogeneous alloy formations can be
further divided into the categories ‘interstitial solid solutions’ and ‘substitutional
solid solutions’. The category of interstitial solid solutions is not further broken
down. Substitutional solid solutions can be categorised according to the formation
of their microstructure through either an inordinate distribution in the lattice, cluster
formation, or the creation of a superlattice. Heterogeneous alloy formations are
separated into precipitates, intermetallic and intermediate compounds, and
segregates.
A phase is defined as all areas, in which an alloy exhibits the same physical and
chemical properties. A mixture consists of two or more phases which can be
physically separated by, for example, sifting or decantation.

63
3 Iron Alloys

Figure 3.1: Classification of alloys.

3.1.1 Homogeneous alloy formation by means of interstitial and substitutional


solution
To a certain extent, alloying and tramp elements can be included in the lattice of the
base metal. The resulting crystal, which contains foreign and base metal atoms, is
called a solid solution, or more commonly, a solid solution. An alloy is
homogeneous when it cannot be microscopically distinguished from a pure metal.
Foreign atoms can be dissolved into the lattice of the base metal in two ways. First,
foreign atoms can replace base metal atoms, resulting in a substitutional solid
solution. This is the case with Fe-Ni and Fe-Mn. Second, foreign atoms can be
dissolved into the interstitial sites of the lattice. This is known as an interstitial
solid solution. Examples are Fe-C and Fe-N.
The possibilities for interstitial or substitutional inclusions of foreign atoms are
shown in Figure 3.2. Interstitial inclusions, as well as the substitution of a lattice
atom (A) by a foreign atom (B; C; or D) cause distortions in the crystal.
A distortion in the ordered structure of the crystal lattice occurs when atoms are
substitutionally included. This is because the atomic radius is slightly smaller or
larger than that of the lattice atoms (Figure 3.3). Furthermore, the inclusion of
foreign atoms in an ideal unit cell can only take place at the largest interstitial sites.
These sites are the octahedral interstitial sites in the fcc lattice and the tetrahedral
interstitial sites in the bcc lattice.
Unrestricted interstitial integration is only possible for small foreign atoms
possessing an atomic radius smaller (or insignificantly larger) than the diameter of
the interstitial sites. The inclusion of carbon (rC/rFe=0.61) or nitrogen (rN/rFe=0.55)
atoms is only possible with a lattice distortion.

64
3.1 Formation of Alloys

Figure 3.2: Configuration of a solid solution: interstitial solid solution (left),


substitutional solid solution (center), combined interstitial and
substitutional solid solution (right).
A combination of the atoms of several components, to form a complex solid
solution with both substitutional and interstitial inclusions, is possible. Some
elements, such as oxygen, can be included both substitutionally and interstitially.
The number of foreign atoms that can be integrated into a homogeneous solid
solution is limited, because the lattice distortion increases with increasing
concentration of foreign atoms.

Figure 3.3: Lattice distortions in a substitutional solid solution.


Solubility is defined as the temperature-dependent maximum dissolved
concentration. In general, the solubility of foreign atoms within a lattice increases
with rising temperature, since the required energy for lattice deformation decreases
due to an enlargement in the unit cell and thermally activated lattice oscillations
(Figure 3.4). However, the solubility is dependent on the existing equilibrium
phase.

65
3 Iron Alloys

Temperature in °C

-1 100 200 300 400 500 600 8001000 1400


0.100

0.050
Log mass-% N or log mass-% C

0.025

Mass-% N or mass-% C
-2 Fe16N2, D 0.010
Fe3C
Fe4N,J 0.005

0.0025

-3 N2 gas 1atm. 0.001

0.0005
Meta-stable
carbide, H 0.00025

-4 0.0001
32 30 28 26 24 22 20 18 16 14 12 10 8 6 4
Inverse temperature 1/T in 104/K

Figure 3.4: Carbon and nitrogen solubility of D-iron in equilibrium with various
phases.
According to the Hume-Rothery rules, alloying and tramp elements will only show
a complete substitutional solid solution formation, when:
x The components in the liquid state are completely miscible with one another
(single phase),
x The components are of the same lattice type (e.g.: fcc-fcc),
x The difference between the atomic radii of the alloying element and the base
metal is not greater than 15 % (e.g.: rFe(fcc)=0.124 nm, rCr=0.125 nm), and
x The components have a similar electronegativity (chemical affinity should be
small).

3.1.2 Heterogeneous alloy formations caused by segregation


The segregation of two components can occur in the melt, for example in the system
Fe-Pb, or during eutectic or peritectic solidification, for example in the system Fe-
C. In systems with an eutectic transformation, a homogeneous melt with a defined,
eutectic composition breaks down isothermally into a heterogeneous structure with
two types of crystals of different chemical composition, D1 and D2 (Figure 3.5).
The prerequisite for the decomposition of a homogeneous solid solution into two or

66
3.1 Formation of Alloys

more phases is a decreasing solubility (increasing supersaturation) of alloying


elements with decreasing temperature.

L L
L+D L+D L+D L+D L

Temperature

Temperature
Temperature

L+D L+E
D D
D
D D E
D DD
DD

A % B A % B A % B
Figure 3.5: Transition of a system with a miscibility gap in the solid state (left) into
a system with eutectic decomposition (right).

3.1.3 Superlattices in substitutional solid solutions and the formation of


intermetallic phases
A homogeneous solid solution with a random atomic ordering of the components
can develop an ordered superlattice during cooling (Figure 3.6, left and center). In
this case, the atoms of alloying elements are not evenly spread throughout the base
metal lattice. Instead, they occupy locally preferred lattice positions. This single-
phase segregation cannot be seen macroscopically. A two-phase segregation occurs
when the coherence of this zone with the matrix is lost, and a precipitate is formed
(Figure 3.6, right).

Figure 3.6: Diagram of precipitate formation: homogeneous solid solution (left),


localised superstructure (center) and precipitation (right).
Segregation in a homogeneous solid solution can lead to the formation of
intermetallic compounds between the base metal Am and the alloying element Bn.
In such compounds, the relation of a metal to another metal or non-metal occurs in a
fixed ratio, m:n, which is determined by the lattice type of the resulting phase
(Figure 3.7). Due to the modified crystal structure, the physical and chemical
properties of the intermetallic phase vary significantly from those of the solid
solution from which they were formed. The compound can be stoichiometrically

67
3 Iron Alloys

composed over a large temperature range (intermetallic), as is the case with CaMg 2,
or the compound can show a range of solubility for its components (intermediate),
whereby the stoichiometric arrangement is reached at the highest temperature within
the range of the phase.
Examples of intermediate compounds are the ‘Laves phases’ of the type AB 2 (e.g.:
WFe2, ZrFe2), the Sigma phase (FeCr), the complex Chi phase (e.g.: in Fe-Cr-Ni-Ti
alloys), and the Hume-Rothery phases of the type AB (e.g.: CuZn). A melting of a
compound without a previous decomposition is called ‘congruent melting’.
In the solid state, a homogeneous solid solution can decompose into two crystals of
differing composition (eutectoid transformation), with decreasing temperature due
to a miscibility gap, or isothermally due to decreased solubility. Furthermore, a
homogeneous solid solution can combine with another solid solution and transform
into a new solid solution, known as a peritectoid transformation.
The abbreviations are defined as follows:
x D03 structure: Fe3Si (largest possible number of Fe-Si partners as closest
and second closest neighbors),
x B2 structure: Degree of order decreases; smaller number of Fe-Si
partners as the second closest neighbor.

68
3.1 Formation of Alloys

Figure 3.7: Superlattices in the Fe-Si lattice.

69
3 Iron Alloys

3.2 Phase Diagrams of Fe alloys

3.2.1 Phase diagrams of binary systems


In phase diagrams, the near-equilibrium state of each phase is given for every
temperature at constant pressure. These phases consist of two (binary), three
(ternary), four (quaternary), five (quinery), or more components of the possible
chemical compositions. The term ‘phase’ is to be understood first of all as the
aggregate state, and subsequently as the allotropic modification, i.e. various
crystallographic structures and their lattice formation. Even complicated binary
phase diagrams can be created by combining a small number of basic phase
diagrams (Figure 3.8).

a b c

Liquid Liquid
Liquid
Temperature

Temperature

Temperature
L+A SS+L
L+B
L+D L+E
A+B crystals Solid Solution D E
(SS) DE

A B A B A B
B in % B in % B in %

d e

Liquid Liquid
Temperature

Temperature

L+E L+D
C+L
E D C+L
L+D L+B

D DE C+D C+B

A B A B
B in % B in %

Figure 3.8: Binary phase diagrams with: a) eutectic, b) complete solubility,


c) eutectic with limited solid solubility, d) peritectic, and
e) intermetallic compound.
The transformation temperature of pure iron is changed through the addition of
alloying elements. Alloying also decreases the liquidus temperature of pure iron,
with the exception of the addition of Ir, Os, Re, and Ru. Especially noticeable is the

70
3.2 Phase diagrams of Fe-alloys

influence of alloying elements on the J-phase. At high temperatures this phase is


limited by the bcc solid solution (G) and at low temperatures it is limited by the
allotropic transformation of the J-solid solution to the bcc solid solution (D).

3.2.2 Thermophysical basics for the expansion or contraction of the J- field


Alloying elements can either expand or restrict the J-phase region (austenite). With
reference to the temperature interval in which the fcc structure of pure iron is stable,
alloying elements are classified into those that expand the J-field by decreasing the
A3-temperature and raising the A4-temperature (austenite formation), and those that
constrict the J-field by decreasing the A4-temperature and raising the A3-
temperature (ferrite formation). Figure 3.9 shows, in terms of thermodynamics, the
influence of various types of alloying elements on the austenite phase. 'H is the
difference between the enthalpy, or heat content, of an alloying element in austenite
and ferrite:
'H = HJ+D
Elements that lead to a negative value of the latent heat of transformation 'H for the
allotropic JoDtransformation cause an increase in the A4-temperature and a
decrease in the A3-temperature and are responsible for an expansion of the J-phase
field. Elements that lead to a positive value of the latent heat of transformation 'H
for the allotropic JoDtransformation cause a decrease in the A4-temperature and an
increase in the A3-temperature and are responsible for a reduction of the J-phase
field. The value of 'H varies greatly amongst alloying elements. This value is an
indicator of the relative strength of an element as an austenite-former or ferrite-
former (Figure 3.10). 

71
3 Iron Alloys

Austenite favored Ferrite favored


T D G J T
D
['H=0] A4
A4 ['H=0]

J
['H negative] J D
 D ['H positive]
J

A3 ['H=0] ['H=0]
A3
D
DJ
- + - +
Fe Fe

Alloying element in mass % Alloying element in mass %


Figure 3.9: Effect of substitutional alloying elements on the Jfield. The range is
extended for a negative 'H and is restricted for a positive 'H.

Figure 3.10: Relative strength of various alloying elements acting as ferrite- or


austenite-formers.

72
3.2 Phase diagrams of Fe-alloys

3.2.3 Expansion of the J-field in iron alloys


As the alloying content of Ni, Mn, Co (Ru, Rh, Pa, Os, Ir and Pt) increases, which
form substitutional solid solutions with J-iron, an expansion of the J-field takes
place (Figure 3.11). Elements whose atoms can be interstitially included in the fcc
iron lattice such as, C, N, H, B, as well as Cu, Zn, Re and Au also cause an
expansion of the J-field. However, when the content of alloying elements is high,
the J-field is restricted by two-phase fields (heterogeneous solid solutions).
1800
G
1600
Temperature in °C

1400
1200
J
1000
800
600

D DJ
400
200

0 10 20 30 40 50 60 70 80 90 100
Nickel content in mass-%
Figure 3.11: Iron-nickel phase diagram.
The Fe-Mn system is another two-component system, which expands the J- field.
This system is graphed in the following phase diagrams (Figure 3.12). The left-
hand graph shows a near-equilibrium system (theoretical graph), while the right-
hand graph shows what is observed experimentally (reality).

73
3 Iron Alloys

1600 1500 Liquid


1536 °C G
1500 1493 °C
1400
1400 G Fe
1392 °C J
1300 900
1200

Temperautre in °C
Temperature in °C

800
1100
J-range 700
1000
911 °C
900 600 D’ J
800 500 Ms
700
400 JD’
600 D

500 D Fe 300

400 200 HJ


300 JH
0 10 20 30 40 50 100
0 10 20 30
Mn in mass-% Mn in atom-%
Figure 3.12: Iron-manganese phase diagrams: theoretical equilibrium diagram
(left); actual phase diagram, taking into consideration difficulties in
maintaining equilibrium conditions at high alloying contents (right).

3.2.4 Restriction of the J-field in iron alloys


The alloying elements Cr, Si, Al, P, Ti, V, Mo, W and Be can cause anything from a
reduction to a complete removal of the J-field. When large amounts of these
alloying elements are present, a transformation from a bcc to an fcc crystal does not
occur during heating or cooling, i.e. during a heat treatment procedure.
The alloying elements Ta, Nb, Zr, and Ce do not cause a complete removal of the J-
phase field, but rather cause its enclosure by heterogeneous phase fields. The Fe-Cr
system is shown in Figure 3.13 as an example of this kind of phase diagram.

74
3.2 Phase diagrams of Fe-alloys

Figure 3.13: Fe-Cr phase diagram.


Another example is the Fe-Si system which is shown in Figure 3.14. The right-
hand diagram is an enlarged section of the left-hand phase diagram showing the
ranges for various superlattices. The meanings of the abbreviations D0 3 and B2 are
the same as mentioned previously with regard to Figure 3.7. The superlattices
cannot be avoided through quenching (freezing the microstructure). Because of
their crystallographic structure they are not suitable for plastic deformation.
Thereby, for example, the technically relevant upper limits can be determined for
hot rolled (<4.5 mass-% Si) and cold rolled (<3 mass-% Si) silicon alloyed steels for
the production of electric sheets.

75
3 Iron Alloys

Figure 3.14: Iron-silicon phase diagram (left); selected segment from the left-hand
graph showing superlattices (right).

3.2.5 The iron-carbon phase diagram


The element carbon is of particular significance during the production of steel.
Depending on the process of reducing ore to iron, carbon appears as a tramp
element. Furthermore, even at low concentrations, carbon is effective as an alloying
element in expanding the J-phase field.
The iron-carbon system is generally graphed as a partial diagram in the form of a
double graph, up to a carbon content of 7 mass-%. In this range, carbon can be in
equilibrium either as graphite, or in a metastable compound with iron, as an iron
carbide (Fe3C called ‘cementite’), as shown in Figure 3.15. Fe-C alloys with less
than 2% carbon are called ‘steel’, and represent a metastable system in which
carbon is in equilibrium as cementite. Above 2% carbon, Fe-C alloys are typically
referred to as cast iron, a stable system in which carbon takes the form of graphite at
equilibrium.
The metastable iron-cementite (Fe-Fe3C) phase diagram will be described in the
following sections (Figure 3.15, Figure 3.16). The phase boundaries are shown as
characteristic points and dashed (Fe-C) or solid lines (Fe-Fe3C) in the diagram.

76
3.2 Phase diagrams of Fe-alloys

Carbon content in atom-%


0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0 22.5 25.0
1600
D’
1536 A
1500 B
H 149 Fe3C
3 °C
I (cementite)
1400 N
1392
D
1300
Temperature in °C

1200 C’
E’ 1153°C F’
E 1147°C C F
1100

1000

911 G
900

800
M 769 O
S’ K’
P’ 738°C
K
700 P S 723°C

600

500
0 1 2 3 4 5 6 L 7
Q
Carbon content in mass-%
0 10 20 30 40 50 60 70 80 90 100
Cementite content in mass-%
Figure 3.15: Phase diagram (double graph) of Fe-C (dashed line) and Fe- Fe3C
(solid line).
An eutectic crystal mixture is not found if the primary melt contains less than
2.06 mass-% carbon. During cooling, when low concentrations of carbon are
present, primary crystals of Gferrite precipitate out of the melt in first order. For
this reaction, AB is the liquidus line, and AH the solidus line. At the point H
(0.10 mass-% C; 1493°C), the Gphase reaches its saturation limit. A peritectic
transformation can be observed here: the Gphase of composition H, reacts with the
melt B (0.51 mass-% C) and forms the Jphase I (0.16 mass-% C). If the carbon
concentration is to the right of I, the precipitated primary Gphase disappears
entirely at the AH line. Further solidification takes place through the precipitation
of the Jphase, whereby BC represents the liquidus line, and IE the solidus line. If
the carbon concentration is to the left of I, then the melt is exhausted before all of

77
3 Iron Alloys

the G-phase is converted, and the further transformation takes place in the solid state
whereby the Gphase, which corresponds to the HN curve, is in equilibrium with the
Jphase, which corresponds to the IN curve. The reaction ends at the temperature
indicated by the point at which the line IN is cut by the perpendicular corresponding
to the carbon concentration. At 1392°C, the A 4-temperature, the HN line meets the
IN line at the point N, which is the GoJtransformation point of pure iron. The
GoJtransformation point is raised by increasing carbon content, and is then spread
over an interval. Alloys between H and I, i.e. between 0.1 and 0.16 mass-% carbon,
pass directly into this GJ phase field after solidification ends at 1493°C. Alloys to
the left of H form a Gphase of uniform composition within a small, limited
temperature range below their solidus temperature.
1600 H

H G-phase and liquid


1536°C A
1500 B 1493°C liquid
G-phase I
1400 N
D
1300 G-phase +
Temperature in °C

austenite
austenite + liquid
1200 cementite + liquid
1147°C F
austenite E C
1100

1000
911°C G austenite + cementite
900 austenite +
ferrite
800
ferrite 723°C K
700 P S

600

500
0 1 2 3 4 5 6 7
Fe Carbon content in mass-% Fe3C
Figure 3.16: Phase diagram of Fe-Fe3C (metastable equilibrium).
The J-phase (austenite) is of more importance due to its much larger solubility for
carbon compared to the G-phase. The J-phase is formed in alloys with less than
0.51 mass-% carbon through a peritectic reaction of G-solid solutions with the
residual melt B. In alloys with 0.51 to 4.3 mass-% carbon, the J-phase precipitates
directly from the melt. At 1147°C austenite reaches its highest carbon content of
2.06 mass-%.
If the carbon content is to the left of S (0.80 mass-% C), the D-phase begins to
precipitate out from austenite at temperatures decreasing from the A3-temperature,
911°C. The D-phase has a very low carbon content with a composition that follows

78
3.2 Phase diagrams of Fe-alloys

the GP curve. The remaining austenite becomes enriched with carbon as its
composition follows the curve GS. If the carbon content of an alloy is greater than
0.80 mass-% carbon, then cementite precipitates out during cooling, while the
solubility of carbon in the remaining austenite decreases along the line ES. When
the A1-temperature of 723°C is reached, the remaining austenite, of composition S,
breaks down into an eutectoid consisting of D-phase and cementite. Cementite also
precipitates from the ferrite by further cooling below 723°C according to the
solubility curve PQ. At 769°C, the A2-temperature, a magnetic transformation
occurs in D iron, as it changes from a paramagnetic to a ferromagnetic state. The
transformation occurs at a constant temperature for compositions up to a carbon
content of about 0.5 mass-%, as shown by the line MO. Ferromagnetic D-iron
precipitates directly out of the non-magnetic J-phase along the curve OSK, when the
carbon content is higher than 0.5 mass-%. Iron carbide also experiences a magnetic
transformation at 210°C, which is called the A0-point. This point of transformation
is independent of the carbon content of the alloy.
At room temperature, the structure of alloys with a carbon content less than 0.02
mass-% carbon consists of primary D-phase with tertiary cementite at the grain
boundaries. Hypoeutectoid alloys with less than 0.80 mass-% carbon contain
primary D-phase, and eutectoid (pearlite) consisting of ferrite and cementite.
Hypereutectoid alloys (0.80-2.06 mass-% C) consist of pearlite and cementite that
precipitates out on the former austenite grain boundaries. As the temperature is
reduced to room temperature, the solubility of carbon in ferrite decreases to
0.0002% carbon. The structural symbols of the near-equilibrium phases of the solid
solutions are:
x G-phase = G-ferrite,
x J-phase = austenite,
x D-phase = D-ferrite,
x Fe3C = cementite.
Furthermore, the cementite is distinguished according to the phase out of which it
precipitated:

x Primary cementite, if it crystallises as the first phase out of the melt


(hypereutectic) along the line CD,

x Secondary cementite, if it precipitates after solidification from the primary


austenite along the line SE,
x Tertiary cementite, if it precipitates from the D-ferrite solid solution along the
line OQ.

79
3 Iron Alloys

If an eutectoid alloy has a fine lamellar structure consisting of D-ferrite and


cementite (Fe3C), then it is called ‘pearlite’. If an eutectic alloy has a structure
consisting of pearlite and cementite, it is called ‘ledeburite’. For most carbon steels
the phase diagram up to 2.06 mass-% carbon is adequate. If the influence of other
alloying elements and cooling rates are neglected, the phase diagram for up to 0.8
mass-% carbon is sufficient to describe the near-equilibrium transformations for
over 90% of the steels produced in Germany.

3.2.6 The Phase diagrams of Fe-N and Fe-H


The gases hydrogen and nitrogen can also be dissolved in iron. The maximum
solubility is graphed according to temperature in Figure 3.17 for ambient pressure.
The solubility of these gases is dependent upon their partial pressure and
temperature, as described by Sievert’s Law. Equation 3.1 describes the following
relationship:
'S 'H

aN = p N · e R R ·T
2
(3.1)
Conversion of the previous equation results in:
1
log aN = const.· (3.2)
T
where aN represents the activity of the given gas, P is the partial pressure of the
given gas, 'S is the entropy, 'H is the enthalpy, R is the gas constant, and T is the
absolute temperature.

3.2.7 Ternary iron systems


Two-component systems are often not sufficient to describe the phases of a complex
alloy consisting of many components. The phase fields can be graphed in a 3D
ternary phase diagram. To simplify the interpretation of the graph, isothermal cuts
of the 3D diagram or 2D diagrams at a given (constant) concentration for one
component are used. Due to the complexity of phase diagrams for systems with four
or more components, a pseudo-binary or ternary diagram is used instead.
New, complex material systems can be extrapolated, and the phases and associated
ranges for specific compositions can be calculated from thermodynamic data. To
calculate thermodynamic equilibria, or to assemble a complete phase diagram based
on experimental data, the “CALPHAD” (CALculation of PHAse Diagrams) method
can be used. According to this method, thermodynamic properties of individual
phases are critically assessed, and the resulting phase equilibria can be calculated
according to the criterion of free-energy minimisation. The application of this
method also allows for the extrapolation of existing data, in order to determine the
thermodynamic equilibrium of a range for which experimental results are not

80
3.2 Phase diagrams of Fe-alloys

available. Examples of software packages that are based on this method of


calculation are Chemsage and Thermo-Calc.
1800 Temperature in °C Temperature in °C
1600
1400
1200
1000

1800
1600
1400
1200
1000
800

800
600

400

600

400
-1.2 1.6
Liquid Fe Liquid Fe

J-Fe
-1.6 1.2

J-Fe
G-Fe
-2.0 0.8
G-Fe
Log [%N]

Log [%H]

-2.4 0.4
D-Fe D-Fe

-2.8 0

-3.2 -0.4
4 8 12 16 4 8 12 16
4 -1 4 -1
Inverse temperature in 10 K Inverse temperature in 10 K

Figure 3.17: Solubility of nitrogen in iron (left) and solubility of hydrogen in iron
(right).

3.2.8 The Phase diagram of Fe-C-Cr


A classic chromium steel is a three component alloy with iron as base metal and
between 12 and 17% chromium and up to approx. 2% carbon. As an example of the
arrangement of the phase fields, Figure 3.18 shows the phase fields for a constant
carbon content. Similarly, Figure 3.19 shows the phase fields for a constant
chromium content. The position and expansion of the phase fields, especially in
relation to the J-phase, are of extreme importance for the development of steels, and
their heat treatment. Results from dilatometric and metallographic analyses of a
three component alloy (Fe-C-Cr) at carbon contents of 0.05% and 0.4% have been
used to demonstrate the influence of carbon on the borders of the homogeneous
austenite phase field, and the positions of the heterogeneous phase fields of
austenite (Figure 3.18).

81
3 Iron Alloys

At higher carbon concentrations, carbides are characteristically formed as a result of


the strong affinity of chromium for carbon, resulting in an increased range for stable
carbides. In comparison to the two-component iron-carbon system, the solubility of
carbon in the fcc crystal at 13% and 17% Cr is clearly reduced (Figure 3.19). Only
partial austenitisation is possible at low carbon contents (ferritic-martensitic
chrome-steels). When the carbon content is extremely low, austenitisation is not
possible at all (ferritic chrome-steels).

82
3.2 Phase diagrams of Fe-alloys

1600
L
1500
L+D
1400

1300
D
1200
J DJ
1100

1000
D+K1
900
DJ+K1
DJ+K2
Temperature in °C

800
DJ+KC D+K2
700 +KC
D+KC
0 10 20 30

1600
L
1500
L+J L+D
1400
L+D+J
1300
DJ
J L+D+K1
1200
DJ+K1
1100
J+K1
1000
J+K2 D+K1
900
2
K
1+
K
J+

800
D+J+K2
D+J+KC +K2
700 D+K2
D+K2+K C
0 10 20 30
Chromium content in mass-%
Figure 3.18: Phase diagram Fe-C-Cr with 0.05 mass-% C (top) and 0.4 mass-% C
(bottom).

83
3 Iron Alloys

1600

1500 L+D
D L
L+ D+ J
1400 D
+ L+J
J
1300
J L+K2
1200
L+ J+K2

1100
J+K2 J+K2+KC
1000
J+K1 J
+
900 D K1
+ +
D J+
Temperature in °C

K1 K2 D+ J+K2
800 D
D+K1 +K1 D+K2
+K2 D+K2 +KC
700
0 1 2 3 4
1600

1500
L+D L
D L+ D+ J
1400
D L+J
+
1300 J
J L+K2
1200 L+ J+K2

1100
J+K2
1000
D J+K1 J
+ +
J K1
900 +
K1 +K2
800 DJ+K2
D+K1 D+K1
+K2 D+K2 D+K2 +KC
700

0 1 2 3 4
Carbon content in mass-%
Figure 3.19: Phase diagram Fe-C-Cr with 13 mass-% Cr (top) and 17 mass-% Cr
(bottom).

84
3.2 Phase diagrams of Fe-alloys

3.2.9 The ternary Fe-Cr-Ni system


The most common Fe-Cr-Ni alloys solidify by primary precipitation of G-ferrite,
which, at the end of solidification and during subsequent cooling, partially
transforms into austenite. The amount and distribution of G-ferrite significantly
influences the mechanical properties of an alloy and the formation of hot cracks.
For technical applications, a multi-component system can be reduced to a pseudo-
ternary Fe-Cr-Ni system by calculating the chromium equivalent and the nickel
equivalent:
x chromium equivalent = % Cr + % Mo + 1.5 (% Si) + 0.5 (% Nb) + 2 (% Ti),
x nickel equivalent = % Ni + 30 (% C) + 0.5 (% Mn).
Both the chromium equivalent and the nickel equivalent take into account the effect
of different ferrite and austenite forming elements on the austenite phase field. In
relating these two equivalencies to one another, the Schaeffler diagram (Figure
3.20) can be derived, in which the nickel equivalent (austenite formers) is graphed
against the chromium equivalent (ferrite formers). By means of this diagram, the
presence of the microstructural constituents ferrite, austenite and martensite can be
easily determined.
30

0%
e
5%
t
rri
Fe

25
%
10
%
A 20
20
Nickel equivalent

%
40
15 A+M
%
A+F 80

10
M 10 0%
A+M+F

5
F+M M+F F

0 5 10 15 20 25 30 35 40
Chromium equivalent

Figure 3.20: Schaeffler diagram for the determination of microstructural


components dependent on the chemical composition.

85
3 Iron Alloys

When applying the system Fe-Cr-Ni to practical steel alloys, it must be remembered
that each steel is a multi-component system, and in many cases, the influence of
other alloying elements on the sequence of phases cannot be ignored. This
influence can be seen in the effect of the alloying elements on the free enthalpy on
the phase boundaries, and in the appearance of new phases.
Case study: Stainless steels – Spoon

Figure 3.21: Spoon for toddlers made of X5CrNi18-10.


Classically the cutlery is made of stainless steel. The spoon is known as the oldest
tool for eating. Its shpae imitates a hand that is cupped. Usually it is marked
‘18/10’, ‘18’ denoting a chrome content of 18% and ‘10’ denoting a nickel content
of 10%. Figure 3.21 shows a spoon for toddlers as an example.
There have been many investigations on the three-component system Fe-Cr-Ni.
However, these only cover certain concentration ranges. The ternary system Fe-Cr-
Ni is limited by the border systems Fe-Cr, Fe-Ni, and Cr-Ni. The Fe-Cr system is
characterised by a complete miscibility in the solid state. The Cr-Ni system is an
eutectic system whereas the Fe-Ni system is peritectic. Figure 3.22 shows an
isothermal section through the three-component system Fe-Cr-Ni, with the
displacement of the solidus-, and liquidus- lines between 1500 and 1470°C, and the
three-phase range (liquid + G + J  7he three lines that extend from the iron-rich
corner of the Fe-Cr-Ni system to the chromium-rich area represent the three-phase
range (liquid + G + J . The arrows marked by G and J represent the direction of the
primary solidification of the liquid that takes place when a change in the chemical
composition occurs.

86
3.2 Phase diagrams of Fe-alloys

Cr
40
TL: Liquidus temperature
TS: Solidus temperature
L: Liquid

70 30

Ch
-%

rom
ss
ma

ium
G
t in

80 20

co
ten

n
J

te n
n
co

t in
n

ma
Iro

ss
TL

-%
90 10

TL TS

TS
TL (L+G+J)
TS 1480°C 1470°C
1500°C
Fe 0 10 20 30 40 Ni
Nickel content in mass-%
Figure 3.22: Isothermal section through the ternary Fe-Cr-Ni system, with the
displacement of the solidus- and liquidus-lines between
1500 and 1470°C, and the lines of the three-phase area (liquid +G
+J .

87
3 Iron Alloys

3.3 Segregation
The term segregation refers to the separation of a melt, leading to an uneven
distribution of the components. The enrichment of elements in a certain area is
called positive segregation, while the depletion of elements in a certain area is
called negative segregation.
An inhomogeneous distribution of alloying and tramp elements is undesirable
because the homogeneity of properties is most often necessary to ensure the quality
of a workpiece. With reference to grain size, segregation can appear as either
micro-segregation, or, if present throughout the semi-finished product, then as
macro- segregation. Primary segregation is the segregation of a homogeneous melt,
which takes place during solidification, in the phase transition from liquid to solid.
It is a result of the different solubilities of the melt (liquid) and the crystals (solid)
for alloying and tramp elements. The concentration difference, which is present
after the solidification is complete, influences the subsequent transformation
behavior (secondary segregation). A difference is made between block segregation
and crystal segregation. Block segregations form during the solidification of an
ingot, because tramp elements and inclusions are pushed ahead of the solidification
front and enrich the remnant melt. In contrast to block segregations, crystal
segregations are localised segregations in the form of microscopic concentration
differences, which occur because crystals containing a low concentration of tramp
elements (P, S, C) precipitate out first at the beginning of solidification. As
solidification advances, the solidified part of the melt becomes enriched with these
elements. The crystals demonstrate a concentration gradient: the amount of alloying
elements increases from the inner to the outer regions. Areas with residual melt and
a corresponding lower melting point show a higher concentration of tramp
elements. The extent of segregation is determined by the size of the solidification
range, the ability of the tramp elements to diffuse in both phases involved, and the
solidification conditions.
In contrast to the segregations resulting from the liquid-solid phase transition, grain-
boundary segregations develop after solidification and are controlled by diffusion.
The grain boundaries, as planar lattice defects are a favored area for the inclusion of
foreign atoms. Grain boundary segregations, which are also simply referred to as
segregations, can be removed through high temperature annealing.

3.3.1 Segregation processes during solidification


The concentration ratios during the solidification of a melt can be determined from
the phase diagram of a system. The assumed linear progression of the liquidus and
solidus lines is satisfactory for a simplified description of primary segregation
(Figure 3.23, left). During the solidification of a homogeneous melt with the
concentration co, a crystal with the concentration cSo is formed. The concentration

88
3.3 Segregation

of the solid phase cSo is in equilibrium with the concentration of the liquid phase c Lo
= co. At the solidification front, which is the phase boundary between the melt and
the crystalline (solid) state, a jump from the solid concentration to the liquid
concentration occurs (Figure 3.23, right). The concentrations of the liquid and the
crystal at equilibrium are connected by a tie-line, the horizontal line shown in the
two-phase region of the phase diagram. The concentration ratio is described by the
equilibrium-distribution coefficient ko, which, as the ratio of the solidus and
liquidus concentrations cS/cL, is a measurement of the difference in solubility
between the solid and the liquid state: k o = cS/cL. The coefficient of delta-ferritic
solidification kG/L, and that of austenitic solidification k JL, are to be distinguished;
both are shown in Table 3.1. The reciprocal of the equilibrium-distribution
coefficient, which is given as a whole number, is the segregation coefficient 1/ko,
which indicates the tendency towards segregation.

Figure 3.23: Phase diagram of the two-component system A-B with complete
solubility: liquidus and solidus lines (left); concentrations at the
solid/liquid phase boundary (right).
Element 1/k0(GG/L) k0(GG/L) k0(JJ/L)
Fe 1 1 1
Cr 1 0.95 0.83
Ni 1 0.79 0.90
Si 1 0.77 0.52
Mn 1 0.75 0.78
Mo 1 0.81 0.63
C 5 0.19 0.34
P 6 0.18 0.09
S 20 0.05 0.035
Table 3.1: Different segregation and solubility coefficients for various elements in
iron.

89
3 Iron Alloys

Models to describe the formation of (block) segregation


The basis for the description of the solidification process is represented most simply
using the phase diagram of a two-component system with complete solubility in the
melt (liquid state) and in solid solution (solid state). Four different models are used
to explain the occurrence of segregation during solidification. These base on
various combinations of the mechanisms of concentration change, diffusion in the
crystal, as well as diffusion and convection in the melt (Table 3.2).
Liquid Crystal
Model Diffusion Convection Diffusion
1 X X X
2 X X -
3 X - -
4 X (X) -
X: possible, (X): limited possibility, -: impossible
Table 3.2: Combination possibilities for diffusion and convection during solidifi-
cation – to derive segregation models.
In model 1, it is assumed that both diffusion and convection occur in the melt, and
that diffusion in the crystals is possible. In this ideal case, complete concentration
equilibrium is reached, and segregation does not occur. After complete
solidification, the whole block of the material has a homogeneous composition,
whose concentration is the same as the starting concentration of the melt.
In technological solidification processes, the possibility for diffusion in the solid
state is not sufficient to reach a complete balance in concentration. Due to the lack
of diffusion in the crystal (model 2), the melt becomes enriched with the elements
that tend to segregate as solidification proceeds. The result is a distinct center-
segregation. The concentration of these elements in the outer layer of crystals also
increases, following the solidus line in the phase diagram.
In the third model, it is assumed that only diffusion in the melt occurs. This means
that the mixing of the melt is realized only through diffusion; convection plays no
role. As a result, a concentration peak forms in front of the solidification front
(Helmholtz layer).
The fourth model assumes no diffusion in the crystals, a limited amount of
convection and an unlimited diffusion in the liquid. This model is a good
approximation of the actual solidification process during the production of billets
and slabs. Figure 3.24 shows the change in the segregated element concentration
from the surface to the center of a block for all four solidification models.

90
3.3 Segregation

Block surface Block core

1
co

Concentration
3
4

2
koco

0 Amount of solid 1
Figure 3.24: Change in the concentration of a segregated element from the surface
to the core of a block (solidification models 1 to 4).

3.3.2 Segregation behavior during dendritic solidification


Dendritic solidification is of special importance for most cast materials. The
cooling conditions on the solidification front can lead to a random growth of
crystals in the melt. The growth of a protuberant, which grows parallel to the
direction of the heat removal, is associated with an enrichment of segregating
elements in the surrounding melt, so that the liquidus temperature TLiq decreases
around the dendrites in accordance with the phase diagram. The crystals, which are
shaped like dendrite arms, also grow perpendicular to the direction of greatest heat
loss if the enrichment of elements and the heat of crystallisation in front of the
dendrite tips are higher.
The enriched residual melt is entrapped between the dendrite roots and arms. The
concentration of alloying elements increases from the cores of the dendrite roots
and arms to their surfaces, in the direction of the remaining melt. After
solidification, areas of lower (dendrites) and higher (residual melt) alloying and
tramp element concentration are found directly next to one another. The
concentration trend corresponds to that of the so called “zone of solid solutions”
(Figure 3.25).

91
3 Iron Alloys

Figure 3.25: Solidification and concentration gradient of a steel with 1% carbon.


The slower the rate of solidification, the coarser the dendrite structures become.
Dendrites that are already formed can change during the solidification process. For
example, thin dendrite arms can begin to dissolve at the sides or tips. The driving
force here is the difference in activity, dependent on the dendrite size. If this
dissolution occurs at the point of connection of several arms, then free-swimming
crystals arise and dissolve in the melt. These crystals can act as nuclei for further
solidification and can eventually sink or rise. The energy required to melt the
crystals comes from the melt, and therefore, superheating is minimised.
For primary austenitic solidifying steels, austenite grain boundaries form around the
dendritic skeleton (Figure 3.26, left). In contrast, during primary D-ferritic
solidification, a transformation to an austenitic structure preferably occurs parallel
to the secondary dendrite arms, which means that it forms transverse to the dendrite
core (Figure 3.26, right).

92
3.3 Segregation

Figure 3.26: Sketch of the formation of grain boundaries during primary austenitic
(left) and primary ferritic solidification (right).

3.3.3 Macrosegregation

Gas bubble segregation


In unkilled cast steel blocks, segregation of gaseous tramp elements can occur.
During solidification gases are set free at the solidification front due the reduced
ability of solid steel to absorb gases such as oxygen, nitrogen, or hydrogen. Oxygen
reacts with some of the carbon present in the steel to form CO. These carbon-
monoxide gas bubbles can become trapped in the steel as it solidifies and form outer
and inner gas bubble defects, as is shown in Figure 3.27.

Figure 3.27: Sketch of an ingot section with gas bubble segregations.

Gravity segregation
This type of segregation is a result of the separation of material due to density
differences within the melt. Free floating crystals with a density higher than that of
the melt sink to the bottom of the melt and form an “angle of repose”. In the case of
continuous casting, the crystals sink to the bottom of the shell.
V- and A-segregations
A-segregations, also called ‘ghost lines’, are found in the dendritic area of the ingot
near the boundary of the angle of repose. They show an enrichment of elements,
particularly sulfur and phosphorus. Contrarily, the V-segregations are aligned near
the axis in the core zone (Figure 3.28). As the solidifying block shrinks, the

93
3 Iron Alloys

formation of the A-segregations results in a formation of cracks. These cracks can


partly be filled with some of the enriched residual melt (crack segregation).
Longitudinal cut

A-segregation Transverse cut

V-segregation

Center of block

Figure 3.28: Sketch of V- and A- segregations.

Inverse ingot segregation


In some extraordinary cases, a higher concentration of alloying elements is detected
in the periphery. Such a situation is given the name inverse ingot segregation, and
results from the following process. As the ingot cools, the outer area of the block
solidifies before the core solidifies, which is still in the liquid state. The solidified
outer area shrinks, and due to shrinkage, cracking occurs. The liquid core, which is
enriched with alloying elements, is sucked into the cracks, supported by capillary
action so that these areas become enriched with alloying elements.

3.3.4 Segregation in continuous casting


The processes of microstructural development in continuous casting and ingot
casting are essentially the same. Each measure taken to influence the structure also
has an effect on segregation, because the process of segregation is closely
connected to structure formation. The microstructure, segregation and internal
cleanliness are dependent on many factors, such as superheating, casting rate,
chemical composition, and cooling conditions. The construction of the process line,
and process-dependent casting conditions, such as superheating, the length of
solidification, and the cooling rate, all lead to different solidification and
segregation profiles (Figure 3.29).
Due to the process conditions, the cooling rate during the continuous casting
process is greater than during ingot casting. The microstructure, both along the
length and in the cross-section, of a slab is more regular than that of the
conventional ingot. Macrosegregation (center segregation) develops in the middle
of the slab during continuous casting of steel. The carbon profile of a slab cross-
section reveals a maximum concentration in the center of the slab, and a few

94
3.3 Segregation

millimeters apart, a minimum concentration of carbon. There are several possible


explanations for the development of this type of segregation. For example, such
macrosegregation can be caused by suction processes, ‘zone-melting segregation’,
or external shape changes, such as the bulging of the slab between the rolls of the
casting machine. This can lead to a high concentration of sulfur, which in turn
leads to an increased concentration of inclusions (manganese sulfides).

1 1
2 2 1
3 3 2 1
5 5 3
4 2
6 3
6 7 5 6 7 5 6 7
7

Vertical Bending- Bow type Caster with


caster straightening caster progressive bend
caster

1 Ladle 5 Roller guide with secondary cooling


2 Tundish 6 Withdrawal and straightening
3 Mold with primary cooling 7 Flame cutting
4 Bending-zone with secondary cooling

Figure 3.29: Design types for continuous casting machines.


Figure 3.30 shows the spacing of secondary dendrite arms as a function of the
cooling rate. The secondary arm spacing varies, as approximately the third root of
the local solidification rate, over a wide range of solidification conditions.
Consequently, the localiszed solidification rate is a function of the growth rate, the
temperature gradient, and the composition of the alloy. The secondary dendrite arm
spacing is important because it determines the spacing of the precipitates or pores
and in this way has an observable influence on the mechanical properties of cast
alloys.

95
3 Iron Alloys

Figure 3.30: Secondary dendrite arm spacing as a function of cooling rate.


Troubles may arise when segregations and other specific defects occur
simultaneously, for example center segregations together with internal cracking.
Segregation cannot be avoided in continuous cast steels with small dimensions and
alloy contents of more than 1% manganese, 1% chromium and 1% carbon. The
center segregation of an unstirred bloom slab can only be removed through long
time homogeneous annealing.
The solidification of killed (deoxidised with Al or Si) steels is necessary for the
continuous casting process as it reduces convection in the melt. Segregation of
killed steels during continuous casting can be improved through stirring (agitating)
of the liquid core of the strand. The electromagnetic processes of inductive and
conductive stirring and continuous casting with a rotating mould are technologically
well advanced. As a result of improved convection, the concentration balance is
improved, and superheating of the melt is reduced. The desired equiaxed core zone
becomes noticeably larger through the application of stirring processes, because
dendrite tips break off and stay in the melt. The substantial amount of center
segregation, which is detectable after the initiation of the continuous casting
process, can be reduced through agitation, while the excellent surface quality is
preserved (Figure 3.31).

96
3.3 Segregation

Figure 3.31: Segregation profiles of ingot casting, as well as unstirred and stirred
continuous casting.
The effect of agitation is first noticeable through the appearance of negative
segregations (depletion), called “white bands”. The localised depletion is the result
of the reduction of the concentration peak, due to the temporary transition from the
convective to the non-convective solidification mode (from Model 2 to Model 3).
Further steps make it possible to improve the segregation profile in the strand. The
formation of macrosegregation and shrink holes can generally be avoided through
small deformations (soft-reduction) in the area of the crater tip. This is especially
true for the minimisation of mini-block formation, in which the afterflow of melt is
hindered when two dendrite fronts grow together during solidification (bridge
formation).

3.3.5 Crystal segregation


The following criteria can be used to describe the extent of crystal segregation:
x size of the solidification interval and slope of the solidus and liquidus lines,
i.e. the segregation tendency as given in the phase diagram (equilibrium
distribution coefficient),
x diffusion ability of tramp elements in the solid and the liquid states,
x solidification conditions.
The effect of crystal segregation can be seen particularly well, for example, through
“Chevrons” (zigzag cracks) during cold extrusion (Figure 3.32). This occurs when
the critical cooling rate is reduced so much (through the local enrichment of
elements) that a martensite transformation occurs.

97
3 Iron Alloys

1 cm

Figure 3.32: Chevron cracking, which occurs during cold extrusion through
segregation-determined martensite transformation.

Removal of crystal segregation


In contrast to ingot segregation, the extent of crystal segregation can be reduced
through diffusion annealing. However, it is often not practical to carry out a
diffusion annealing on a cast ingot because when large primary grains are involved,
the diffusion paths can be too long. Therefore, the material is usually hot formed
before diffusion annealing is done. The stronger this prior deformation is, the
shorter the necessary holding time generally is at a constant diffusion annealing
temperature.
The necessary annealing time is adjusted according to the degree of segregation, the
tolerable degree of residual segregation after annealing, the rate of diffusion of the
segregating element, and the diffusion path (distance of the concentration maximum
= f(degree of forming)). After diffusion annealing, normal annealing is necessary at
least once in order to obtain a fine grained structure. The difficult and costly
process of diffusion annealing, which requires hours of annealing time, is generally
only used for high-quality structural components and materials requiring top
quality, and a high amount of alloying elements, such as tool steels.

98
3.3 Segregation

3.3.6 Secondary segregation


Primary crystal segregation, which develops during progression through the
liquid/solid two-phase field, can lead to carbon segregation in the solid state if
continued cooling of the steel occurs, or if the temperature is held constant within a
certain temperature range. During hot forming, the dendrites and the residual
segregated interdendritic regions are elongated in the direction of deformation. The
extent to which they are elongated is dependent on the degree of deformation. The
areas of higher concentration (residual fields) are then no longer spherical or net-
shaped, but instead, have an ordered linear form. The resulting secondary structure
has an irregular structural quality due to the uneven distribution of carbon, and
tramp elements.
The influence of alloying elements on carbon segregation during continuous
cooling (structural banding)
The microscopic analysis of hot formed hypoeutectoid steels quite often reveals an
arrangement of bands, consisting of pearlite and ferrite, running in the direction of
the deformation of the secondary microstructure. These formations in the secondary
microstructure are the direct result of primary segregation (Figure 3.33).

20 μm

Figure 3.33: Secondary bands of a segregated ferritic-pearlitic microstructure.


For energetic reasons, during the cooling of a homogeneous austenite in the two-
phase field J + D, nucleation and growth of proeutectoid ferrite always occurs at the
austenite grain boundaries. Austenite that is inhomogeneous, due to crystal
segregation, behaves differently during transformation. In this case, the type and
direction of carbon segregation can influence the transformation (Figure 3.34).
The structural bands are formed due to local variations in the A 3-temperature caused
by alloying elements. Within the Fe-Fe3C system for hypoeutectoid steels alloying
elements are to be distinguished into those that decrease the A3-temperature as their

99
3 Iron Alloys

concentrations increase (austenite formers), and those alloying elements that


increase the A3-temperature as their concentrations increase (ferrite formers).

Figure 3.34: Structural band formation resulting from silicon segregation during
continuous cooling.
Ferrite formers such as, silicon, molybdenum, and chromium, increase the A 3-
temperature. If an austenite cools in the J + D field, in which carbon is
homogeneously distributed, but the alloying elements are inhomogeneously
distributed, then proeutectoid ferrite forms first of all in the areas with higher
concentrations of Si, Mo, and Cr. Carbon diffuses in front of the austenite-ferrite
reaction-front in the areas depleted of alloying elements, where the primary
structure consists of dendrites. After austenite transformation is complete, a ferritic
structure is present in the regions of the alloying element rich residual fields, and a
pearlitic structure in the dendritic regions. If the steel is hot formed and not
quenched, a microstructure consisting of ferrite and pearlite bands is formed. This
type of banding is called silicon banding (Figure 3.35, left).
Austenite formers such as manganese and nickel, decrease the A3-temperature. The
proeutectoid ferrite precipitation begins in the regions of the dendrites with lower
concentrations of these elements, because in those regions the A 3-temperature is
higher. In areas with higher manganese or nickel contents, a pearlitic structure is
present after transformation, while in the dendritic area a ferritic structure is present.
This type of banding is called manganese banding (Figure 3.35, right). The
metallographic evidence which type of banding prevails is often based on the
position of inclusions, which, with few exceptions, are formed in the residual melt.

100
3.3 Segregation

If, for example, a sulfide is found within a ferrite band, silicon banding can be
assumed, or vice versa.
Such carbon segregation has not yet been proven in hypereutectoid steels, because
the austenite grain boundaries are much more preferred over the segregation regions
as sites for proeutectoid carbide precipitation.
T T
R

A3 D A3 D

T % Si T % Mn

R D

D
R

%C %C

pearlite in D pearlite in R

ferrite in R ferrite in D
D: dendrite (low alloying element content) R: residual field (high alloying element content)

Figure 3.35: Diagram of ferrite-pearlite banding: silicon banding (left) and


manganese banding (right).

Influence on structural banding


Due to structural banding, anisotropy of properties occurs. In particular, difficulties
in cutting and unfavorable mechanical properties can develop under transverse or
perpendicular load. The creation of a band-free or low-banded structure is possible
through either the avoidance of crystal segregation, which is nearly unrealizable, or
more realistically, through a targeted adjustment of ferrite and austenite forming
elements or a heat treatment. Hence, two possible methods are relevant: accelerated
cooling and diffusion annealing.

Combination of ferrite and austenite forming elements


The displacement of the Ar3-temperature in the areas of the former residual melt as
compared to the dendritic areas can be compensated if the influence of ferrite
formers is balanced by austenite formers. Then, the carbon segregation takes no
longer place during cooling below the A r3-line. The combination of elements

101
3 Iron Alloys

depends on the mass-percentages and the segregation coefficients. This effect is


used especially for manganese and silicon containing quenched and tempered steels.

Accelerated cooling
During the formation of ferrite bands from inhomogeneous austenite, carbon has to
diffuse away from the transformation front. This segregation can be hindered by an
accelerated cooling. The temperature control has to be selected such that the ferrite
formation begins in the slowly transforming bands before carbon enrichment takes
place in the transformation-favorable areas due to advanced ferrite formation. In
principle, this accelerated transformation can occur isothermal or through
continuous cooling. During a diffusionless transformation from austenite
(quenching), all diffusion of carbon is suppressed, and thereby, a segregation of the
material can be avoided (band-free structure after hardening). In contrast, during a
diffusion-controlled transformation, it is important to adjust a band-poor or band-
free ferritic-pearlitic microstructure. In this case, the continuous cooling rate has to
be carefully chosen to hinder carbon diffusion, while ensuring that the steel
transforms into ferrite and pearlite. In Figure 3.36, such cooling rates are marked
on the CCT-diagram of the non-alloyed quenched and tempered steel Ck45 for
various rod diameters.
The removal or reduction of structural banding achieved in this way is reversible.
This means that through renewed austenitising and slow cooling, banding occurs
again.

Figure 3.36: Cooling curves for various rod diameters superimposed on


continuously recorded CCT-diagram of the steel Ck45.

102
3.3 Segregation

Influence of alloying elements on carbon segregation during annealing in the


two-phase field ferrite-carbide
Investigations on Si-Cr-Mn-Mo-Ni model-alloys showed that the carbides formed
during tempering below the Ac1-temperature become more stable as the
concentration of elements that normally accumulate in the carbides increases in the
solid solutions. Carbides are stable in areas locally enriched with elements that
reduce the activity of carbon. On the other hand, in areas containing alloying
elements that increase the activity of carbon, carbides are unstable and dissolve.
The released carbon then diffuses, even against its concentration gradient (up-hill-
diffusion), into areas in which alloying elements are present that reduce its activity.
As a result, carbon segregates, while the alloying elements remain in the iron
matrix, due to their relatively low diffusion rate. The driving force behind this
diffusion is an activity gradient, but not a concentration gradient as is assumed in
the simplified diffusion Law. Therefore, the alloying elements chromium,
manganese, and molybdenum accumulated in the residual fields create favorable
conditions for the formation of carbides. The carbides formed in the low alloying
element zones (dendrites) dissolve and agglomerate with carbides formed in the
enriched zones (residual fields), which leads to a spheroidisation of these carbides.

Influence of alloying elements on carbon segregation in the two-phase field


austenite - carbide
Along with the characteristic ferrite-pearlite bands in hypereutectoid steels,
undesirable carbide bands or networks form in tool steels, which can result in
flaking, excessive and uneven warpage, and hardening cracks. The carbide bands
result from the formation of ledeburitic clusters or networks in the residual fields
due to crystal segregation during solidification. Through hot deformation, the
carbide clusters in the as-cast structure are crushed and stretched, more or less
strongly, in the direction of deformation. Carbide clusters also form as a result of
carbon segregation during holding in the austenite-carbide two-phase field. During
the cooling of a hypereutectoid steel from the homogeneous austenite field into the
austenite-carbide two-phase field, proeutectoid carbides segregate out
predominantly on the austenite grain boundaries. At an increased rate of cooling,
they take on a needle-shaped form in a Widmannstätten structure. When annealing
a martensitic structure, an even distribution of undissolved carbides does not
develop. The resulting carbon segregation can occur in two ways:
x The inhomogeneous distribution of alloying elements changes the solubility
of carbon in austenite: manganese increases the ability of austenite to
dissolve carbon and silicon reduces it.
x As a consequence of the enrichment of alloying elements, carbon segregation
results from the difference in stability of the carbides in various areas.

103
3 Iron Alloys

In hypoeutectoid steels, carbon segregation in the phase field ferrite and carbide is
determined by the stability of the carbides. Both the activity of carbon in austenite
and the stability of the carbides have an effect when annealing hypereutectoid steels
in the phase field austenite and carbide while these two influences work against one
another.
Silicon reduces the ability of austenite to dissolve carbon. In silicon rich areas the
local carbon concentration is low. Simultaneously, silicon increases the activity of
carbon (effective carbon concentration), which thereby reduces the activity gradient
necessary for the diffusion of carbon out of the austenite into the carbides. Since
silicon is not soluble in cementite, the silicon content is high at the phase boundary
austenite/cementite, causing an effective delay or even suppression of carbide
precipitation.
Manganese increases the solubility of carbon in austenite. In manganese rich areas,
an increased concentration of carbon is present. Therefore, carbide precipitation is
favored.
The influence of alloying elements on carbon solubility overlaps with their
influence on carbide formation. This follows from the solubility of the elements in
the carbide, and their influence on carbon activity at the phase boundary
carbide/austenite.
However, research on alloy models has shown that the influence of alloying
elements on the solubility of carbon in austenite is clearly hidden by the stabilising
effect of the carbides (Table 3.3).
Mn Si Cr Mo Ni
C solubility + - - - +
C activity - + - - +
carbide formation + - + + -
in residual fields

Table 3.3: Influence of alloying elements on carbon solubility, activity, and carbide
formation in residual fields of hypereutectoid steels (+ increase,
- decrease).

3.3.7 Examination of segregation


Generally, segregation can be qualitatively proved through metallographic etching
techniques. In the past, quantitative verification of segregation was achieved
through chemical analyses of samples with varying areas of segregation. Today, the
state of the art for quantitative evaluation is microprobe analysis. In the following,
the most commonly used methods for the testing of segregation are presented.

104
3.3 Segregation

Pickle tests
Rough, pre-ground discs or block halves are pickled with an acid (e.g. hydrochloric
acid) or acid mixture (e.g. nitro hydrochloric acid) suitable for the steel. Apart from
the development of the casting structure, zones of segregation appear due to the
variance in acid attack determined by the electrochemical potential. Pickled discs
are often used as a quality check both at the start of operation, and during
production in strand casting (Figure 3.37).

Heyn etch
In Heyn etching, roughly ground samples are etched with a diluted copper
ammonium chloride solution. The areas of residual melt appear dark, while the
dendrites are bright. This etching process allows for a fast process inspection,
although the verification of the segregation of phosphorous and sulfur is only
possible with low resolution.

Figure 3.37: Pickled disc from a horizontal section of a continuous cast strand
(Material 1.4501). Recognisable are the finely crystallised outer layer,
the dendritic middle layer, and the solidified globular core.

Oberhoffer etch
A polished sample is etched with a solution consisting of water, ethyl alcohol,
hydrochloric acid, iron chloride, copper chloride, and tin chloride. Opposite to the
Heyn etch, the residual melt appears bright and the dendrites dark. With use of this
etchant, the casting structure, as well as macro- and microsegregation can be seen.
The course of the fibers can be developed in hot formed steels, and therefore,
conclusions can be drawn regarding the deformation or finishing process. In order
for the etchant to react, the steel must contain a certain amount of phosphorus
(Figure 3.38).

105
3 Iron Alloys

Figure 3.38: Example of an Oberhoffer etch; recognisable are extrusion lines of the
deformation process and phosphorus segregations.

Baumann print
Photographic paper, soaked in a 5% sulfuric acid solution, is pressed with the
emulsion side down on the surface of a pre-ground sample. The following reactions
occur between the photo paper and the sulfur-containing area of the material:
H2SO4 + FeS o FeSO4 + H2S,
H2SO4 + MnS o MnSO4 + H2S.
The resulting hydrogen sulfide reacts with the silver bromide of the photo paper:
H2S + 2 AgBr o Ag2S + 2 HBr.
After a reaction time of one to five minutes, the paper is fixed and washed with
water. Sulfur segregations appear brown to black-brown as shown in Figure 3.39.
Qualitative statements about the solidification processes can be made on the basis of
the sulfur segregation which developed.

106
3.3 Segregation

Figure 3.39: Baumann print of a raw continuous cast strand.


Atom Probe Tomography
Atom Probe Tomography (APT) covers the finest scale of 3D chemical analysis
available for complex alloys. Its application enables us to study the segregation
features at nano- and atomic scale. An example of Cr nano- segregation in high
carbon bearing steel 100Cr6 is shown in Figure 3.40. A local overview of the atom
distribution in a 60nmh60nmh126nm volume in a mixture microstructure of
bainite (B) and martensite (M) is indicated. The Cr atoms segregate in form of a
nano-sized tube across the B/M boundary in the material. This indicates that the Cr
nano- segregation is formed prior to the formation of bainite and martensite
microstructure and the Cr nano- segregation is not eliminated by the previous
homogenization and austenitization treatments.

107
3 Iron Alloys

Figure 3.40: 3D atom map showing the Cr nano- segregation in high carbon
bearing steel 100Cr6.
Atom Probe Tomography (APT) is a characterization technique enabling spatially
resolved chemical analyses of materials at sub-nanometer resolution (ᇞx=ᇞy ≈ 0.2
nm and ᇞz ≈ 0.1nm). Figure 3.41 shows the Local Electrode Atom Probe
(CAMECA, LEAP 4000X HR) at IEHK, Aachen, Germany. The instrument is
equipped with a local electrode, a wide-angle reflectron, a high-speed delay line
detector system as well as an ultrafast laser with a wavelength of 355 nm.
Comparing to the conventional atom probes, such an instrumental design provides
fast data acquisition rates (up to 2 Mio ions/min), large analysis volumes and high
compositional accuracy. Large volumes containing up to several hundred millions
of atoms can be probed within a few hours and the impurity concentrations as low
as few tens of ppm can be detected.

108
3.3 Segregation

Figure 3.41: Local Electrode Atom Probe (CAMECA, LEAP 4000X HR) at IEHK,
Aachen, Germany.

109
3 Iron Alloys

3.4 Internal Cleanliness in Steel


Every steel contains, to a greater or lesser degree, substances that have an entirely
different composition to that of their surrounding matrix, and which form a distinct
phase boundary area.
One speaks of a non-metallic inclusion when this second phase has a non-metallic
character, and a metallic inclusion when this second phase has a metallic character.
The inclusion is said to be exogenous if it enters the steel from outside. If the
inclusion is the product of a reaction that occurred either in the melt or during
solidification, then it is called endogenous.
The endogenous non-metallic inclusions have a large influence on the
characteristics of steel. The most important properties of endogenous inclusions are
summarised in Figure 3.42. These are either deoxidation products (oxides) or
alloying compounds (sulfides or carbonitrides). In the last few years, steel users
have placed extreme requirements on the purity of steel, meaning the absence of
non-metallic inclusions in steel. The production of steels with the highest grade of
purity, in order to meet these demands, has been made possible by secondary
metallurgical processes.
The following types of inclusions are distinguished:
x SS: sulfide inclusion in linear form,
x OA: oxide inclusion in dissolved form (aluminiumoxides),
x OS: oxide inclusion in linear form (silicates),
x OG: oxide inclusion in globular form.
The appearance of exogenous metallic inclusions is relatively seldom. These can be
cases of incompletely dissolved alloying additions, such as ferro-chromium. Most
of the time these inclusions are either refractory material that found its way into the
melt, entrained slag or casting powder remnants.

110
3.4 Internal Cleanliness in Steel

Slag/inclusion SS OA OS OG
type
Composition MAS2 CA2
M=MnO, C=CaO MnS Al2O3
CaS M2A2S5 CA
S=SiO 2 , A=Al2O 3 CA6

microhardness HV 170 2200 unknown 1100


930
Hot deformability

2200
deformation index* ~1 0 v = f(T) 0
800°C...
v=0
1100°C...
v=1
micro cracks micro cracks
defects no defects possible no defects probable
thermal expansion 18.1 8 5 5
Quenching stresses

coefficient 14.7 2 6.5


Din 10 K
-6 -1 8.8

[matrix (850-50°C)
-6 -1
=12.5 10 K ]

phase boundary micropores tensile stress tensile stress tensile stress

E-modulus, GPa
147 390 146 115
contact stresses

[matrix: 210GPa]

increase in stress
1.2 2.5 1.2 1.8

* F0 - cross section of the block a - length of the inclusion


F1 - cross section of the semiproduct b - thickness of the inclusion

Figure 3.42: Physical properties of non-metallic oxidic and sulfidic inclusions.

111
3 Iron Alloys

3.4.1 Internal cleanliness


The internal cleanliness of steel is characterised by the number, size, composition,
shape, amount, and distribution of non-metallic inclusions. Non-metallic inclusions
are found in the macroscopic range from 1.5 to 10 mm. In the microscopic range,
the inclusions reach the resolution limits of light microscopy.

Processes to determine the macroscopic internal cleanliness


The internal cleanliness of steel has reached a very high qualitative level with the
commonly used metallurgical processes of today. The implication of this for the
inspection of materials is that inclusions have to be found that occur at a very low
frequency and with an uneven distribution throughout the steel. A thorough
inspection, for example, using ultrasonic continuous testing, is only possible with
certain types of products. There are two standardised processes in use for
determining the macroscopic internal cleanliness of steel, the blue brittleness test
and the stepped torsion test.
In the blue brittleness test, an unnotched sample is broken at temperatures between
300 and 400°C. The non-metallic inclusions that do not tarnish, in comparison to
the steel, can be seen with the naked eye or with a minimum amount of
magnification. The inclusions are classified into ten classes according to the
guidelines in the Stahl-Eisen-Prüfblatt 1584.
In the stepped torsion test, cylindrical sections of various thicknesses are made by
turning a sample. By analysis of the surface of these sections with the naked eye,
the total length and number of non-metallic inclusions can be determined. The
sample preparation is, hereby, labor intensive, while the investigated sample volume
is relatively low.
According to a newly developed method of testing, the “surfboard” method,
samples from cast strands are transversely rolled at an angle perpendicular to the
direction of casting. In the case of round strands, a 130 mm high cylindrical sample
is rolled out to a 13 mm thick sheet, such that the diameter of the sample in one
direction, from the bottom to the top, is stretched to five times its original value,
while in the perpendicular direction, the diameter is doubled (Figure 3.43). As a
result, the inclusion band on the inner curve of the strand is pressed flat in the
direction of casting and is concentrated in a small area. In sections A and B, the
upper side relevant to the internal cleanliness, is subject to an ultrasonic test. This
method of testing has the advantage that a comparatively large sample volume is
investigated. Due to the defined deformation, the inclusions form a surface
favorable for the reflection of ultrasonic waves. The “surfboard” method has
proved itself a well suited test method for describing the macroscopic internal
cleanliness. However, it must be noted that samples are usually taken from raw cast
strands, and consequently, the test results are not generally valid, but rather

112
3.4 Internal Cleanliness in Steel

dependent on the chosen proportion of deformation from raw block to prepared


sample.

Figure 3.43: Sketch of a specially deformed sample for ultrasonic testing to


determine the macroscopic internal cleanliness.

Processes to determine the microscopic internal cleanliness


The processes to determine the microscopic internal cleanliness are defined in
DIN 50602. To determine the internal cleanliness, at least six samples of a certain
size from a single mould charge must be examined. The inclusions are then
registered and classified through comparison with a photo series.
A cleanliness value K (K 0 to K 8) can be calculated by adding all of the inclusions
with a certain weighting on the sample surface under investigation; this is also
known as calculating the internal cleanliness according to the “K” process. Often
the “M” process serves as a sufficient evaluation, whereby only the maximal
number of inclusions present in a given sample is specified.

113
3 Iron Alloys

Figure 3.44 shows a small section of the photo series used in the testing of steels
for non-metallic inclusions according to DIN 50602.

Figure 3.44: Small section of the photo series used in testing of steels for non-
metallic inclusions according to DIN 50602.
Figure 3.45 shows an analysis table for the inclusions found in six samples of a
stainless steel billet. When the sample surface area is converted to 1000 mm², a
total “K 4”-value is reached, which characterises the internal cleanliness.

114
3.4 Internal Cleanliness in Steel

Sample Area of the Inclusion Number of the inclusion with respect to the code number Multiplication
number micro- type and subtotal
section according 0 1 2 3 4 5 6 7 8
in mm² to the
table Factor fs
0.05 0.1 0.2 0.5 1 2 5 10 20 S *) O*)
1 450 SS 3 1 1 - - 10
OA 5 1 - - -
OS 2 - - - - 10
OG 1 - - - -
2 400 SS 4 2 - - - 8
OA 3 1 1 - -
OS 2 1 - - - 21
OG 2 - 1 - -
3 350 SS 2 1 - 1 - 14
OA 4 2 - - -
OS 1 1 - - - 12
OG 1 - - - -
4 600 SS 5 - - - - 5
OA 8 1 - - -
OS 1 1 - - - 15
OG - 1 - - -
5 250 SS 1 1 1 - - 8
OA 3 1 - - -
OS 1 - 1 - - 14
OG 1 1 - - -
6 300 SS 4 1 - 1 - 16
OA 2 2 1 - -
OS 2 - 1 - - 21
OG 1 1 - - -
sum 2350 subtotal S: 61 O: 93
*) S = suflide; O = oxide K 4-value **) S: 26 O: 40
**) converted to a microsection area of 1000 mm² total K 4-value 66

Figure 3.45: Example of an analysis to determine the “K 4”-value.

3.4.2 Sulfide inclusions


Sulfur dissolves in liquid iron; however, its solubility is very low in the solid state.
Sulfur precipitates out on the grain boundaries during solidification as iron sulfide
(Fe1-xS), an eutectic component with iron (Figure 3.46). The eutectic has a low
melting point of 988°C.

115
3 Iron Alloys

Sulfur content in mass-%


0 10 20 30 40 50 60 70 80 90 100
1600
1538°C
(GFe) L1
1394°C
1400
1365°C L1 + L2
L1
1200 (JFe) 1188°C L2
1062°C
54.2 ~99.7
Temperature in °C

1000 988°C 1 bar S 2


927°C 44.6
~99.94
912°C
Magnetic transformation Fe1-x S
800 770°C 55 743°C
697
617°C
600 (D Fe)
FeS 2(pyrite)
444.5°C
400 B.P.
315°C

FeS2(marcasite)
200
138°c 115.22°C

(S)
0
0 10 20 30 40 50 60 70 80 90 100
Sulfur content in atomic-%
Figure 3.46: Phase diagram iron-sulfur.
In the presence of oxygen, the FeS-FeO-eutectic can be formed with a melting point
of 988°C. When heating into the temperature range of hot forming (900-1200°C),
liquid phase appears on the grain boundaries, which leads to cracks during a plastic
deformation. This failure mechanism is called red shortness.
In steels of today, the danger of red shortness is reduced through the lowering of
oxygen and sulfur contents and the addition of manganese. Manganese has a higher
affinity for sulfur than iron, and forms manganese-sulfide with a melting point of
1610°C. The amount of sulfides formation depends on the sulfur content and its
maximal solubility. As can be seen in Figure 3.47, the solubility of sulfur is about
0.05% in the J-field at 1370°C. If the melt is rapidly cooled from high
temperatures, the sulfur will remain in supersaturated solution or in an extraordinary
finely distributed form. The maximum solubility is greatly reduced through the
addition of manganese.

116
3.4 Internal Cleanliness in Steel

Figure 3.47: Section of the phase diagram Fe-S (top) showing the influence of
manganese on the solubility of sulfur in J-iron (bottom).

3.4.3 Oxide inclusions


Next to sulfur, oxygen is the most significant impurity in steel. Oxygen appears in
the form of oxide inclusions, which can be of both, microscopic and macroscopic
size. The measurement of the occurrence of oxide inclusions generally results in
two peaks in the distribution curve (Figure 3.48). The curve for the microscopic
inclusions follows a Gaussian distribution. However, macroscopic inclusions can
no longer be said to follow a Gaussian distribution, since today they are isolated

117
3 Iron Alloys

cases and therefore, cannot be described statistically. The number of macroscopic


inclusions is in the range of a few permille of the number of microscopic inclusions.

Deoxidation - reoxidation products


Number of grains

Microscopic inclusions

Macroscopic inclusions
I
II

0 20 100 200 1300


Average grain diameter in μm
Figure 3.48: Distribution curve of oxide inclusions in an Al-killed (deoxidised)
steel.
Figure 3.49 shows different forms of oxide growth in dependence on the relation of
local activities of oxygen and deoxidizing metal. This scheme is based on
investigations of the growth morphology of aluminum oxides, manganese oxides,
and silicates.

a[O] a[Me]

Figure 3.49: Forms of oxide morphology and their dependence on the proportion of
local activities of oxygen and deoxidizing metal.

118
3.4 Internal Cleanliness in Steel

The process begins with a globular-shaped growth at low concentrations of the


deoxidizing metal. As the activity increases, the globular shape becomes instable
and changes to a rosette-like shape and finally forms dendrite-like growths. The
branched growth changes into a more compact form with further increasing activity
of the deoxidizing metal.

3.4.4 Influence of calcium on the formation of non-metallic inclusions


In aluminum killed steels, solid inclusions, such as Al 2O3, can be deposited on the
nozzle during strand casting, which can affect the castability of the steel. With
calcium-metallurgy, inclusions are modified (Inclusion Shape Control) and removed
from the steel, along with other undesired tramp elements. Dissolved calcium reacts
with the aluminum oxide inclusions and results in a multitude of calcium-
aluminates. In Figure 3.50, various non-metallic inclusions are shown in the three-
component system CaO-SiO2-Al2O3. The deformation behavior changes with the
chemical composition. The pure oxides SiO2 and Al2O3 are difficult to deform.

Deformation behavior of the inclusions


0 100 non-formable

10 90 poor deformability
good deformability
%

20 80
in

very good deformability


nt

30 70
nte

Inhomogeneous inclusions
SiO
-co

40 60 with just one deformable phase


aO

2
-co
+C

50 50
nte
nt
nO

60 40
+M

in
%

70 30
O
Fe

80 20

90 10

100 0
0 10 20 30 40 50 60 70 80 90 100
Al2O3-content in %

Figure 3.50: Chemical composition and deformability of oxide inclusions.


Small amounts of added calcium can affect the castability, because inclusions of the
type CaO˜6Al2O3 with a melting point of 1850°C and inclusions of the type
CaO˜2Al2O3 with a melting point of 1750°C can form. These inclusions are present
in the solid state in liquid steel. With the increased addition of calcium, inclusions
of the type CaO˜Al2O3 and 12CaO˜7Al2O3, with melting points below 1600°C, can

119
3 Iron Alloys

form. They are in the liquid state at the casting temperature of steels and therefore,
improve castability (Figure 3.51).
If the addition of calcium is increased even more, then calcium sulfide inclusions
are formed. At temperatures necessary for steel production, these inclusions are
found in the solid state and, therefore, hinder the casting process. The aluminates,
found in the solid state at the casting temperature, join with calcium to form liquid
calcium-aluminate inclusions. The number of manganese sulfide inclusions is
reduced, and spherical CaS-CaO-Al2O3 inclusions are formed, which are not
deformable in the rolling process.
max. max.
TL
1600°C

CaS in inclusions
Dissolved Al 2O 3

Al2O3

“Window”
Inclusions
completely liquid CaS

0 0
0 max.
Amout of Ca-addition
Figure 3.51: Change in inclusion composition through the addition of calcium.

Behavior of non-metallic inclusions in the hot deformation process


Typical examples of the various forming behaviors of non-metallic inclusions are
shown in Figure 3.52. The steel matrix, containing oxide inclusions, can be
understood as a composite. Since a force is carried across the grain boundary, the
deformability of both components is relative and plastic flow is low. Generally, the
inclusions do not withstand the shear stress during deformation and behave in a
fragile manner and brittle fracture. The fragments from crushing during
deformation are dispersed in the flowing metal matrix. Hard, crystalline inclusions
remain undeformed; coral-shaped structures are broken up, and agglomerates are
scattered.
The plasticity of the inclusions in comparison with the plasticity of the steel
considerably influences the form of the inclusions after deformation. Whenever the
deformability of the inclusions and the metal are not equal, the inclusions are a
potential source of defects in the material. The deformation stabilities of the matrix
and inclusions follow different temperature dependencies. This phenomenon can be
described by the deformation index (Q), which gives the ratio of the inclusion
extension to the extension of the steel matrix at a certain temperature:

120
3.4 Internal Cleanliness in Steel

b
2 ln
Q a (3.3)
F0
3 ln
F1
where F0 = cross section of the block, F1 = cross section of the semiproduct, a =
length of the inclusion, and b = thickness of the inclusion.
The softer the inclusions behave in comparison to the matrix, the higher the
deformation index Q becomes. Figure 3.53 shows the temperature dependence for
the deformation index of manganese-silicate, manganese-sulfide, alumina, and
calcium-aluminate. If the deformation index is greater than one, then the extension
of the inclusion is greater than the extension of the matrix; if the deformation index
is less than one, then the extension of the matrix is greater and pores can develop.
Cast Modification
After hot deformation After hot deformation
condition with Ca

Carbides

Silicates

Al 2O3 CaS Al O
2 3

CaO

Titanium
carbides and
nitrides

Oxides and.
oxi-sulfides

Sulfides
MgS

Figure 3.52: Deformation behavior of non-metallic inclusions in steel.

121
3 Iron Alloys

Figure 3.53: Influence of temperature on plastic deformation of various inclusions


in a steel matrix.
The deformation index of inclusions containing either alumina or calcium-aluminate
does not exceed the value 0.1 in the temperature range up to 1300°C. With
decreasing temperatures, the deformability of manganese-sulfides increases in
relation to the matrix, while contrastingly, that of manganese-silicates decreases
steeply at approximately 1000°C. This steep decrease is strongly influenced by the
content of manganese, calcium and iron in the silicate.

122
3.5 Further Readings

3.5 Further Readings


Dantzig, J.A.; Rappaz, M.:
Solidification
EPEL Press, 2009

Flemings, M.C.:
Solidification processing
McGraw-Hill, 1974

Kurz, W.; Fisher, D.J.:


Fundamentals of Solidification
4th ed., Trans Tech Publications Ltd, Juni 1998

Kubaschewski, O.:
Iron-Binary Phase Diagrams
Springer Verlag, Berlin, Heidelberg, New York, Düsseldorf, 1982

Leslie, W.C.:
The physical metallurgy of steels
Herndon, Va. TechBooks 1991

Massalski, T.B.:
Binary Alloy Phase Diagrams
2nd ed., Volume 1 + 2, ASM Materials, Park Ohio, 1996

Porter, D.A.; Easterling, K.E.; Sherif, M.:


Phase Transformations in Metals and Alloys
3rd ed., CRC Press, February 2009

123
3 Iron Alloys

Verein Deutscher Eisenhüttenleute (Ed.):


Steel
Volume 1: Fundamentals
Verlag Stahleisen, Düsseldorf
Springer Verlag, Berlin, Heidelberg, New York, Tokyo, 1992

Arata, Y.; Matsuda, F.; Katayama, S.:


Solidification crack susceptibility in weld metals of fully austenitic stainless
steels
Trans JWRI 5 (1976), pp. 35-51

DIN 50602:
Metallographische Prüfung des Gehaltes nichtmetallischer Einschlüsse in
Stählen mit Bilderreihen
October 2003

DIN EN 10020:
Begriffsbestimmung für die Einteilung der Stähle
July 2002

Stahl-Eisen-Prüfblatt 1580:
Stufendrehversuch zur Prüfung von Stählen auf makroskopische
nichtmetallische Einschlüsse
December 1970, Verlag Stahleisen

Stahl-Eisen-Prüfblatt 1584:
Blaubruchversuch zur Prüfung von Stählen auf makroskopische
nichtmetallische Einschlüsse
December 1970, Verlag Stahleisen

124
4 Phase Transformations

4 Phase Transformations
The great variety of properties in steels is primarily due to the possibility of
adjusting the microstructures of steels over a wide range. Allotropic phase
transformations, the phase transformations with crystallographic structural changes,
and precipitation processes in which a two-phase microstructure develops from a
supersaturated solid solution are very significant. The dependence of the phase
transformation on the variables temperature, concentration, and pressure is graphed
in a phase diagram.
In many technological processes the JoD transformation is of great importance
because it occurs at temperatures used in heat treatments and hot deformation
processes. This transformation can progress near-equilibrium, as described in the
phase diagram, or it can at first be suppressed during rapid cooling and then take
place in the temperature range of supercooled austenite. This transformation can no
longer be described in a phase diagram. Much more to this subject will be
discussed in the time-temperature-transformation diagrams in Chapter 5.
The phase transformations of solid bodies can be divided into diffusion-controlled
and diffusionless transformations. The diffusion-controlled transformations are
thermally activated, while the diffusionless transformations are athermal and can
only be controlled through supercooling. The phase transformations that can appear
during cooling from the austenitic phase and the resulting microstructures are given
in Table 4.1. The phases ferrite and carbide, as well as the phase mixture pearlite,
are formed by diffusion-controlled processes. During diffusionless phase
transformations, atoms move only over distances that are significantly smaller than
the length of one lattice space. Therefore, these transformations can also occur at
lower temperatures where diffusion is no longer possible.
Typical temperatures for diffusion-controlled transformations are between the
equilibrium temperature and around 500°C; typical cooling rates are < 160 K/s. If
the transformation occurs without diffusion by means of a shearing process, then the
martensitic phase, which forms at high cooling rates below 500°C, is created as
transformation product. A combination of diffusion-controlled and shear
transformation can appear within a wide range of the cooling rate. Such
transformations are called bainitic transformations.
The metallographically visible microstructures of steels are strongly dependent on
the prevailing boundary conditions of the phase transformations. Usually, an
austenitic microstructure serves as starting point for technical heat treatments from
which stable transformation products are created during cooling, although, the grain
size and homogeneity of austenite also influence the transformation product.
According to the cooling rate, as well as supercooling below the equilibrium
temperature, equiaxial ferrite and pearlite, bainitic ferrite and pearlite, bainite –

125
4 Phase transformations

possibly even in combination with bainitic ferrite or martensite – as well as lath


martensite can be created from austenite.

Characteristic Type of phase transformation


I II III
1. Mechanism diffusion-controlled bainitic diffusionless
2. Microstructure ferritic-pearlitic bainite martensite
irregular D
equiaxial D
+ carbide- granular bainite
+ normal lath martensite
poor to upper bainite
pearlite
pearlite
3. Parameters:

3.1 Transformation approx. 900 to approx. 640 to


below 510 °C
temperature 640 or 510°C 250 °C
3.2 Cooling rate very slow up to approx. approx. 160 to higher than
160 °C/s 3000 °C/s 3000 °C/s
Table 4.1: Phase transformations while cooling from the austenitic state of an
unalloyed steel with approx. 0.08 mass-% carbon.
Three groups of steel transformations are distinguished according to their formation
temperature and the resulting microstructure: the transformations into pearlite,
bainite, and martensite. The temperatures given in Figure 4.1 are only a hint at the
respective transformation ranges. These ranges are strongly dependent on the
content of alloying elements and the cooling conditions. The term “shear” describes
a cooperative atomic movement, which is reflected in lattice shearing.
Along with these allotropic phase transformations, precipitation also plays an
important role in adjusting the microstructures of steels. Precipitates are secondary
phases that form below the solubility level and are not iron modifications. With
this, the iron matrix does not experience any change in its crystallographic structure.
In pure iron-carbon alloys, carbides are very important, particularly, the iron-
carbide Fe3C, cementite, and its modifications.
In Table 4.2 phase transformations are compared with other metallographic
phenomena. Changes in density, crystal structure, crystallographic orientation, as
well as remaining deformation of the initial body are important characteristic
properties. The external form of deformation, for example, is recognisable as a
relief on an originally polished surface. The ferritic-pearlitic, the martensitic, and
the bainitic phase transformations, as well as precipitation from the supersaturated
solid solution are more closely examined in the following sections.

126
4 Phase Transformations

Kind of
Microstructure transformation Transformation range
900

ferrite, carbide, diffusion in J pearlite range


700 pearlite then transformation
Temperature in °C

500 diffusion in J
bainite bainite range
then shear
then diffusion in D and J
300

martensite shear martensite range


100 then diffusion in D

0
Figure 4.1: Formation of the most important microstructures in unalloyed steels.

Characteristic properties

Dislocation Mechanical Diffusion- Martensitic


Changes Precipitation
gliding twinning controlled transformation
in:
transformation

Density no no yes yes yes

Crystal no (in the


no no yes yes
structure matrix)
Crystallo-
no (in the
graphic no yes yes yes
matrix)
orientation
External
form
yes yes no yes no
(relief
formation)
Table 4.2: Characteristic properties of various metallographic phenomena.

127
4 Phase transformations

4.1 Ferritic-Pearlitic Transformation


4.1.1 Morphology
If a material is cooled near-equilibrium after austenitising to a temperature below
the Ac1-temperature, the closed packed Jsolid solution (fcc, austenite) transforms
in accordance with the iron-cementite phase diagram. The Jsolid solutions
transform into the less closed packed D-solid solutions (bcc, ferrite) and into
orthorhombic cementite. Depending on the carbon content of the steel, the
metallographic microstructure consists of ferrite and cementite.

steels with a very low ferrite (D-mixed crystal) and


carbon content (< 0.02%C) tertiary cementite

hypoeutectoid steels proeutectoid ferrite (D-mixed crystal)


(< 0.8%C) and pearlite (eutectoidic cementite)

eutectoid steels pearlite


(0.8%C)

hypereutectoid steels pearlite and secondary cementite


(>0.8%C) (mainly on the grain boundaries)

Figure 4.2 shows typical product microstructures of the ferritic/pearlitic


transformations. The microstructural component ferrite is formed by diffusion-
controlled transformation into pearlite. Ferrite (Latin “ferrum”) is the
metallographic term for the cubic body-centered D-solid solution, in which carbon
is interstitially dissolved. In its typical shape, it is seen as a sequence of polyhedral-
shaped surfaces in the two-dimensional representation of the flat metallographic
sections (Figure 4.2A). For near-equilibrium cooling, the carbon solubility of the
D-lattice is represented by the line P-Q in the iron-cementite phase diagram (Figure
3.15) as a function of temperature. The maximum carbon solubility in bcc iron is
0.02 mass-% at 720°C while the carbon solubility in bbc iron at room temperature is
very low (~10-5 mass-%).

128
4.1 Ferritic-Pearlitic Transformation

Figure 4.2: A: ferritic microstructure of a low carbon steel,


B: pearlitic microstructure of an eutectoid Fe-C steel,
C: ferritic-pearlitic microstructure of a hypoeutectoid steel,
D: pearlitic microstructure with grain-boundary cementite in a
hypereutectoid steel.
The most important structural characteristic of pearlite, formed at higher carbon
concentrations, are alternating plates (“lamella”) of ferrite and cementite, whereby,
the properties of pearlitic materials are dependent on the interlamellar spacing O. In
hypereutectoid steels, the cementite forms on the austenite grain boundaries. With a
decreasing formation temperature, the pearlitic microstructure can be divided as
follows.

coarse lamellar pearlite

fine lamellar pearlite


decreasing
lamellar spacing O
very fine lamellar pearlite (sorbite)

very fine lamellar pearlite (troostite)

A steel that only undergo a pearlitic transformation after austenitising has a


microstructure that can consist of the following phases.

129
4 Phase transformations

ferrite and tertiary cementite

ferrite and pearlite increasing


carbon content
pearlite

pearlite and secondary grain boundary cementite

Since significantly more ferrite is present in pearlite than in cementite, ferrite forms
the matrix, in which cementite is found as a plate-like structure (lamellar pearlite;
Figure 4.2B). The typical color of ferrite and cementite is white. Therefore,
pearlite, as a mixture of two white crystal types, must also be white. Generally, if
under lower magnification, pearlite appears dark, this is due to a shadow effect of
the incoming light and is not a coloration of this phase (Figure 4.3). The name
“pearlite” comes from the golden yellow sheen often appearing in light microscopy,
which resembles mother-of-pearl.
Incident light

Figure 4.3: Sketch of a pearlite cross section. The incoming light creates a shadow
behind the protruding cementite lamella (c = cementite, D ferrite).
Above all, iron-carbon alloys with less than 0.8% carbon precipitate out primary
carbon-poor D-solid solutions during cooling, as soon as the curve of the Fe-Fe3C-
diagram drops below the GS-line. As a result, the remaining J-solid solutions
become enriched with as much as 0.8% carbon, at 723°C, and afterwards,
decompose in eutectoid pearlite (Figure 4.2C).
On the contrary, iron-carbon alloys with a carbon content between 0.8 and 2.06%
precipitate out primary carbon-rich Fe3C crystals during cooling, as soon as the as
the curve of the Fe-Fe3C-diagram drops below the ES-line. As a result, the
remaining J-solid solutions become depleted of carbon and decompose isothermally
at 723°C into eutectoid pearlite (Figure 4.2D).
The iron-carbide Fe3C precipitates out during near-equilibrium cooling as a more or
less closed network on the austenite grain boundaries due to favorable nucleation

130
4.1 Ferritic-Pearlitic Transformation

conditions. It contains 6.67% carbon according to the iron-carbon diagram and is


known in metallography as cementite. In the orthorhombic unit cell, with lattice
constants a = 0.4517 nm, b = 0.5079 nm, and c = 0.673 nm, there are four Fe3C
molecules; in other words, there are 12 iron atoms and 4 carbon atoms (Figure 4.4).
Each carbon atom is surrounded by four iron atoms in a tetrahedral structure. The
tetrahedrons are connected with one another alternating either along an edge or at a
corner. Cementite is extraordinarily hard (800-1150 HV10); it has a lower density
than iron (UFe = 7.86 g/cm3, U Fe C = 7.4 g/cm3) and is magnetic at room temperature.
3

Figure 4.4: Position of iron atoms in the cementite crystal. The atoms g and d are
surrounded hexagonally by the atoms h, e, f and a, b, c.
Cementite appears as an independent microstructural component in the following
three different structural forms, all having the same chemical composition.

primary cementite created from the melt


h l
secondary cementite precipitation from austenite

tertiary cementite precipitation from ferrite

131
4 Phase transformations

4.1.2 Formation of the ferritic microstructure


Under industrial circumstances, steels are often cooled from austenite to room
temperature. In hypoeutectoid steels, the first transformation occurs at the A 3-
temperature. In Figure 4.5 a part of the iron-carbon phase diagram is shown.
950

Austenite

a)
Temperature in °C

Ferrite + Austenite b)
c)
d)

Ferrite Ferrite + Cementite

500
0 1.0
Carbon content in %

Austenite Austenite Austenite Pearlite

Ferrite- Ferrite
nuclei
a) b) c) d)

Figure 4.5: Section of the metastable iron-carbon diagram (top). Sketch of the
microstructural development from austenite (a) through the
precipitation of proeutectoid ferrite (b and c), until pearlite (d) for an
alloy with 0.5 mass-% carbon (bottom).
For a given binary alloy with a mass content of 0.5% carbon, a slow cooling is
plotted on the graph. At point a) the alloy is fully austenitic. The transformation of
austenite begins when the A3-temperature is reached at point b). Ferrite nucleates
on the grain boundaries of austenite. This transformation is called proeutectoid
ferrite precipitation. As cooling continues, ferrite grows along the grain boundaries.
The growth process is characterised by the diffusion of carbon out of the ferrite into
the austenite, since a maximum of 0.02 mass-% carbon is soluble in ferrite. This
causes a respective increase in the carbon content on the phase boundaries of the
austenite. At point c) a microstructure is present in which ferrite completely covers

132
4.1 Ferritic-Pearlitic Transformation

the austenite grain boundaries. At this point, just before the eutectoid
transformation, the microstructure consists of up to almost 40% ferrite. A further
cooling below the eutectoid temperature leads to a transformation of the remaining
60% austenite into pearlite (point d). Figure 4.2 C shows a microstructure which
was formed this way.
During proeutectoid ferrite precipitation, there are local changes in chemical
composition. These changes in chemical composition can be seen in the iron-
carbon diagram. If the conode belonging to point c) is added to the two-phase
region, then the carbon content in austenite cJ as well as the carbon content in ferrite
cD can be read from the graph (Figure 4.6). If the carbon profile is schematically
plotted at this temperature, then these carbon contents correspond to the maximum
and minimum values in the profile. Carbon piles up on the phase boundaries in
austenite, because diffusion in austenite progresses more slowly than in ferrite.
Within the grains, the carbon content decreases to the initial value c0.

Figure 4.6: Section of the metastable iron-carbon diagram (left) and schematic
representation of carbon enrichment on the austenite-ferrite phase
boundary at point c) (right).
Usually, industrial steels are not only binary iron-carbon combinations, but alloys
consisting of multiple components. The addition of further alloying elements
greatly influences the proeutectoid ferrite precipitation. Hence, a ternary Fe-C-X
alloy can contain ferrite and austenite with different carbon contents, while at the
same time, the Fe and X contents in the ferrite and austenite are almost or
completely identical. In general, different mechanisms can be distinguished that
describe the participation of substitutional alloying elements in the proeutectoid
ferrite precipitation:

133
4 Phase transformations

1) Ferrite growth with participation of the alloying element X, assuming local


equilibrium (Local Equilibrium – LE),
2) Ferrite growth without participation of the alloying element X, assuming local
equilibrium (Local Equilibrium Non Partitioning – LENP),
3) Ferrite growth without participation of the alloying element X, assuming local,
metastable equilibrium (Paraequilibirum – PE).
Under local equilibrium conditions (mechanism 1), both carbon and the alloying
element X can diffuse. The growth rate of ferrite is determined by the diffusibility
of the alloying element X, which is much lower than that of carbon. The
appearance of this mechanism depends significantly on the alloying composition.
Under the same assumption of local equilibrium, mechanism 2 allows for a much
more rapid growth than mechanism 1, because the reaction depends only upon the
diffusibility of the interstitial dissolved carbon atoms. In front of the mobile phase
boundary, however, there is either an enrichment or depletion of the alloying
element X, which takes the form of a concentration peak. Whether there is an
enrichment or a depletion depends upon whether ferrite or austenite forming
elements are concerned (see Chapter 3).
Mechanism 3, which assumes a paraequilibrium on the phase boundary, is
characterised by a complete “freezing” of the substitutional alloying element.
Therefore, it is a prerequisite that the ferrite growth rate is so high that the
substitutional alloying element does not participate in the reaction, and hence, the
atomic ratio Fe/X remains constant. Carbon diffusion is, however, possible. It
controls the transformation without any participation from the alloying element X.
Transformation takes place when an atom jumps from one phase into another, and
the number of jumps is thermally activated. For the phases to grow without
changing the concentration, the following applies for the forward jump of particles
(J to D):
Q1

z1 z ˜ e R˜T
(4.1)
and for the jump backwards (D to J):
Q2

z2 z˜e R ˜T
(4.2)
where z = number of atomic jumps and Q = activation energy per jump. The
x
mobility rate W of the phase boundary results from the difference between z 1 and
z2 :
x
ª Q1

Q2
º
W | z1  z 2 z ˜ «e R˜T  e R˜T » (4.3)
¬ ¼

134
4.1 Ferritic-Pearlitic Transformation

x 
Q1
ª 
R˜T º
Q Q
2 1

W z ˜e R˜T
«1  e » (4.4)
¬ ¼
where Q2 – Q1 = 'G:
x 
Q1
ª 
'G
º
W z˜e R ˜T
«1  e R ˜T
» (4.5)
¬ ¼
Q1

The first term z ˜ e R˜T represents the mobility and diffusibility of the atoms. The
ª 
'G
º
second term «1  e R˜T » is the driving force behind the phase formation.
¬ ¼
x
At a given temperature, the mobility rate W is independent of time so that the
growth of the new phase is constant and only dependent on temperature. As
supercooling increases, 'G increases and with that, the driving force of the
transformation increases. In contrast, atomic mobility decreases with sinking
Q
temperature ( 1 increases), and the transformation becomes more difficult.
R ˜T
x
During the JoDtransformation, W reaches a maximum. If the transformation
x
progresses from DoJ, then W does not reach a maximum, but rather increases
continuously with temperature (Figure 4.7).

Figure 4.7: Graph of a diffusion-controlled transformation with J as the high-


temperature phase and D as the low-temperature phase.

135
4 Phase transformations

4.1.3 Formation of the pearlitic microstructure: transformation with


simultaneous precipitation
At the eutectoid temperature of 723°C and a carbon concentration of 0.8%, the three
phases austenite, ferrite, and cementite are in equilibrium. No transformation force
exists in either the direction of the pure austenitic state or in the direction of the
phase mixture of ferrite and cementite. An austenitic transformation is
thermodynamically possible only after the temperature drops below A1 during near-
equilibrium cooling. The J-solid solution decomposes according to the eutectoid
reaction:
J-solid solutionoD-solid solution + Fe3C
The nucleation of the phase mixture pearlite occurs on the austenite grain boundary
or on the phase boundary of austenite with a proeutectoid precipitation (ferrite or
cementite). The first unit of pearlite consisting of a ferrite or cementite grain begins
primarily at the grain boundaries and interfaces. During this nucleation, a carbon
diffusion inevitably takes place because of the great difference in carbon
concentration between the J-phase, the D-phase, and cementite. Therefore, the
pearlite transformation belongs to the diffusion-controlled, discontinuous reactions.
It is also called a double reaction, because one partial reaction step involves the
formation of the D-phase, and a second, the formation of the carbides.
If, for example, the first nucleus that forms and has the ability to grow is a ferrite
nucleus, then carbon precipitates into the surrounding austenitic matrix. This is
because the carbon solubility of the D-solid solution is lower than that of the J-solid
solution. The accumulation of carbon on the phase boundary between ferrite and
metastable austenite leads to the formation of orthorhombic cementite, which has a
relatively high percentage of carbon atoms. At the existing temperature, this
carbide, due to the nature of the cementite lattice, provides a more energetically
favorable site than the lattice vacancies of the supersaturated austenite. Therefore,
an activity gradient develops between cementite and austenite, through which, due
to carbon diffusion, the cementite nucleus can continue to grow (Figure 4.8).
On the side of the carbide grains, the austenite becomes depleted of carbon, which
increases the probability of ferrite nucleation. This is how the lamellar
configuration of a pearlite island forms. The lamellae that are parallel to one
another and whose crystallographic orientation is the same are called a “colony.” If
a ferrite or cementite nucleus coincidentally assumes another orientation, then a
pearlite colony with a new lamellar direction forms (Figure 4.8-5). The direct
observation of the orientation of the colonies in relation to austenite is complicated
by the fact that the colonies possess fixed crystallographic relationships to the
austenite grains into which the pearlite colony does not grow.

136
4.1 Ferritic-Pearlitic Transformation

A linear time function exists for the growth of faces and edges, from this follows
that the growth in volume of a colony is dependent on t3. What is exceptional about
the pearlite growth mechanism is that carbon does not have to diffuse over long
distances in the austenite. Instead, it is distributed by short diffusion flows such
that cementite and ferrite form next to one another because neighboring fields have
differing relative carbon concentrations.
Figure 4.9 shows the carbon concentration curve and clarifies the mechanism of
material movement at the growth-front (incoherent pearlite/metastable austenite
phase boundary). On the ferrite/metastable austenite phase boundary, carbon
accumulation can be seen due to the repulsion of carbon by ferrite. Carbon
depletion occurs on the cementite/metastable austenite phase boundary, due to
“carbon removal” by the growing cementite lamella.
The given proportions would result in a stop of lamellar growth, due to the growth-
hindering effects of the carbon distribution at the phase boundaries. However, this
is not the case, because of a concentration compensation through cross-diffusion.
This brings about the diffusion of excess carbon atoms from the ferrite/austenite
phase boundary to the cementite/austenite phase boundary and therefore, further
lamellar growth is possible, i.e. an advancement of the phase boundary into the
metastable austenite.

Fe 3C DFe
1) 2) Fe3 C 3)

Initial Fe3C nucleus Fe3 C plate full-grown, D-Fe plates now full-grown,
D-Fe now nucleated new Fe3 C plates nucleated

4) 5)

Formation of a new Fe3C nucleus New pearlite colony


with a different orientation on the at an advanced stage
surface of a continuously of growth
growing pearlite colony
Figure 4.8: Nucleation and nucleus growth of pearlite.

137
4 Phase transformations

Figure 4.9: a) Schematic representation of pearlite and the cross-diffusion in


austenite, indicated by the arrow; b) Local carbon distribution in front
of pearlite in the section D-D’; c) Concentration ratios on the
cementite/austenite and ferrite/austenite phase boundaries.
Cross-diffusion is described by Fick’s first law:
CJD  CJ
J=D˜ c
(4.6)
O
where J = diffusion flow, D = diffusion coefficient = D0 exp(-Q/RT), O lamellar
spacing of pearlite, CJD  carbon concentration in austenite on the JD phase
boundary, and CJc = carbon concentration in austenite on the J/cementite phase
boundary.
The processes of nucleation and nucleus growth during pearlite formation can be
understood macroscopically as the temporal change in the amount of pearlite in the
entire microstructure and can be described mathematically with the help of the
JMAK Equation (compare with Equation 4.41, Ch. 4.4.1 theoretical basics).
During the non-equilibrium cooling of hypoeutectoid steels, which have a ferritic-
pearlitic microstructure at equilibrium, both an increased supercooling and an
increased cooling rate result in a larger decrease in the percentage of proeutectoid
ferrite in the entire microstructure. A carbon content above 0.5% leads to a
complete pearlite formation. This is visible in the extension of the GS and ES lines
in the Fe-Fe3C phase diagram (Figure 4.10). With that, proeutectoid precipitation
is suppressed, and pearlitic transformation is favored.
To the left of the extrapolated ES line (area I) supercooled austenite is
supersaturated with regard to the proeutectoid ferrite. To the right of the

138
4.1 Ferritic-Pearlitic Transformation

extrapolated GS line (area III), the supercooled austenite is supersaturated with


regard to the proeutectoid cementite. Therefore, in the areas I and III, the
transformation begins with proeutectoid precipitation. The supercooled austenite is
supersaturated with both ferrite and cementite only in area II. The transformation
directly involves the formation of pearlite. Therefore, with sufficient supercooling,
a pure pearlitic microstructure with a carbon content starting at 0.5% can be found.
During proeutectoid ferrite formation, carbon enrichment takes place at the
ferrite/austenite phase boundary, which gradually raises the carbon content of the
metastable austenite up to that of the eutectoid composition. With increased
supercooling, the diffusion of carbon from the phase boundary into the austenite
becomes more difficult. Therefore, it piles up on the phase boundary. If the
accumulation reaches a critical point, then the formation of ferrite stops, because the
nucleation conditions for pearlite are fulfilled, although the austenite at the phase
boundary has not yet reached the eutectoid concentration. In this way, an iron-rich
pearlite is created.
This is how, for example, the steel C45 with around 50% ferrite and 50% pearlite, in
equilibrium with one another, can be transformed to 100% pearlite by increasing the
cooling rate (increased supercooling). This is used industrially in the production of
wires with a patenting treatment in a lead bath (Figure 4.11 a-c).

Figure 4.10: Schematic supercooling diagram of pearlite.

139
4 Phase transformations

a) furnace cooling b) cooling in air c) cooling in a lead


x x
T | 14 K/min T | 10 K/s bath
x
ferrite/pearlite pearlite T | 500 K/s
pearlite/grain
boundary cementite
Figure 4.11: Influence of cooling rate on the amount and form of pearlite in a steel
with 0.45 mass-% C.

Temperature dependence of nucleation and nucleus growth


The nucleation rate is controlled by the supercooling 'T. As the temperature
decreases (increased supercooling), the thermodynamic instability of austenite
increases and the diffusion rates of iron and carbon decrease. The surface formation
of a nucleus benefits from supercooling. An increased supercooling raises the
energy required to form an interface. Normally a lamellar configuration forms since
the pearlite formation is a discontinuous coupled reaction. A lamellar configuration
forms because the diffusion path of carbon and, therefore, the time to pearlite
formation is shorter than for a globular carbide formation in a ferritic matrix. The
greater the supercooling is, the finer the pearlitic structure becomes. The energy for
interfacial formation is taken from the driving energy, i.e. in the more extreme case.
The entire driving energy is used up on interfacial formation. In reality, part of the
energy must be used for nucleation.
The observation for diffusion-controlled transformations, that the nucleation rate
can be derived from the opposing behavior of the diffusion coefficient and the sum
of potential nuclei, can be applied for pearlite formation. The nucleus growth rate
gives the mobility rate of the product-phase/metastable austenite phase boundary
and is mainly determined by diffusion and interfacial stresses. The structure and
consequently, the mechanical properties of pearlite are a direct result of the
effective nucleation rate and nucleus growth rate during the transformation of the
given steel. Figure 4.12 shows the calculated curves of the nucleation rate and
nucleus growth rate for an eutectoid steel (0.78% C, 0.63% Mn) as a function of the
transformation temperature.
The nucleation rate and nucleus growth rate constantly increase over the
investigated temperature range (A1 to 550°C). At low transformation temperatures,

140
4.1 Ferritic-Pearlitic Transformation

nucleation dominates because the nucleus growth rate only increases slightly due to
the difficulty of diffusion.
725

700
Temperature in °C Nucleus growth rate

650

600

550 -5 -4 -3 -2 -1
10 10 10 10 10
Nucleus growth rate in mm/s

-4 -2 0 2 4
10 10 10 10 10
Nucleation rate in nuclei/mm³s
Figure 4.12: Nucleation rate and nucleus growth rate of an eutectoid steel,
dependent on temperature.

Lamellar spacing
The mechanical properties of pearlite are significantly determined by the distance
between the lamellae. The lamellar spacing O describes the distance between two
lamella axes. The lamellar spacing visible in a pearlite section is dependent on the
angle at which the surface cuts the ferrite-cementite disk stack (Figure 4.13). The
true lamellar spacing O is only visible in those pearlite grains whose lamellae
coincidentally lay perpendicular to the plane of the cut. If the angle D between the
plate perpendicular and the surface of the cut is >0°, then the apparently larger
lamellar spacing OD is given by the equation:
O
OD = (4.7)
cos a
The apparent lamellar spacing becomes larger as the angle of intersection
Dincreases. At D q the cos D= 1, i.e. OD O Beginning at around D = 80°, OD
grows very rapidly (Oq OOq OOq O 

141
4 Phase transformations

Figure 4.13: True and apparent lamellar spacing in pearlite.


A constant value for the true lamellar spacing O can only be spoken of in
approximation because a slight statistical scatter around the average of O is seen.
The equation below is generally valid:
Oa = 1.65 Omin (4.8)
where Oa = average lamellar spacing, and Omin = minimum distinguishable lamellar
spacing.
Independent of the steel composition the lamellar spacing becomes smaller as the
pearlite formation temperature decreases (increased supercooling). It can be
estimated using the effective thermodynamic and kinetic variables (Figure 4.14).
During the growth of a lamella with two interfaces over the distance dx in the
direction of the J-phase, the following boundary area enthalpy must be supplied:
'GB = 2·W·V·dx (4.9)
where 'GB= interfacial enthalpy, W = the width of the lamella, V the specific
interfacial energy and dx = growth length.
The enthalpy of volume released during the transformation JoD + Fe3C, produces
the following (total volume) gain in enthalpy:
ΔH v ·'T
'EV = O·W·dx·'GV = O·W·dx· (4.10)
TE
where 'EV = total gain in enthalpy (with regard to volume), 'GV = enthalpy gain
through transformation, 'HV = enthalpy of volume, 'T =TE-TT, TE = equilibrium
temperature, and TT = transformation temperature.
Setting 'GB = 'EV results in the following:
2σ 2σ TE
O= = (4.11)
ΔG v ΔH v ·'T

142
4.1 Ferritic-Pearlitic Transformation

Since Oa = 1.65·Omin | 2·O the supercooling 'T must become larger, and the true
lamellar spacing is then:
4σ ˜ TE 4V ·TE 1 const .
Oa = = ˜ = (4.12)
ΔH v ˜ 'T 'H V 'T 'T
The value for O depicts the minimum lamellar spacing. In this extreme case, growth
would stop. From this equation, we learn that O becomes smaller as supercooling
increases. To estimate the lamellar spacing practically, the following empirical
formula is used:
15
O| (4.13)
'T

Figure 4.14: Dependence of the lamellar spacing O in pearlite on the formation


temperature.
4.1.4 Influence of lamellar spacing on mechanical properties
The mechanical properties of pearlite are set mostly through lamellar spacing and
hence, through supercooling. Additionally, they can be determined by the size of
the pearlite colonies, which are reduced by decreasing transformation temperature
and smaller austenite grain size. For instance at the same lamellar spacing, the
reduction in area increases and the transition temperature decreases. (Figure 4.15).
For the same yield strength values, the toughness properties improve as the pearlite

143
4 Phase transformations

fraction increases. A fine lamellar microstructure has the same effect on the
mechanical properties as a fine grained microstructure.
As the carbon content increases or cooling accelerates, the percentage of pearlite in
the transformed microstructure increases with supercooling. Compared to ferrite,
the harder pearlite, with increasing percentages of carbon, has a higher yield point
and the steels are stronger. Limits to the mechanism of strengthening through
pearlite are determined, when exceptional toughness properties, especially good
notch impact toughness, are needed for use at low temperatures. In such cases, the
lowest possible pearlite content must be determined, in order to attain a low
transition temperature and to protect from brittle fractures. The resulting loss of
strength due to the necessary reduction of pearlite must be compensated through
other strengthening mechanisms.
Both a reduction in pearlite colony size and a smaller lamellar spacing increase the
reduction in area in unalloyed and low-alloyed steels. Since lamellar spacing
decreases as supercooling increases, the cementite lamellae formed at high
temperatures in a steel with 0.8% carbon are significantly thicker than the lamellae
formed at lower temperatures (Figure 4.15, curves 1 and 3 a-c). Thick lamellae are
not deformable and, therefore, lead to a very low reduction in area. In contrast, very
thin lamellae are easily deformable and show a significantly higher reduction in area
with concurrently increasing yield strength values (Figure 4.15, curves 1 and 3).
In addition a small solony size increases the reduction in area.

144
4.1 Ferritic-Pearlitic Transformation

Lamellar spacing O in μm

2
Reduction in area Z in % Yield strength R e in N/mm
0.40 0.15 0.10 0.07 0.05
1000 0.8% C
2 1
0.6%C
800
0.4%C
600
0.2%C
400

200

75

60

45

30 3a
3b 1
15
3c
0
0.05 0.10 0.15 0.20 0.25 0.30

Lamellar spacing Oin μm


1: Values for pearlite in a pure Fe-C alloy with approx. 0.8% C
2: Values for very fine pearlite with 0.2-0.8% C; 0.25% Si and 0.7% Mn
3: Values for pearlite with various pearlite colony diameters
(3a: approx. 14 μm, 3b: 25-40 μm, 3c: approx. 140 μm)
Figure 4.15: Yield point and reduction in area as a function of pearlite lamellar
spacing.
4.1.5 Influence of alloying elements on pearlite formation
Pearlite formation changes with the addition of alloying elements to binary iron-
carbon alloys. Alloying elements change the free enthalpy of transformation and by
this the equilibrium temperature for pearlite formation. For a given temperature of
transformation this will change the supercooling and consequently the growth rate
of pearlite. Figure 4.16 indicates first with increasing Ni and Mn content the
growth rate is slightly reduced in the transformation temperature range 700-500°C,
while a Mo addition sharply reduces pearlite formation.

145
4 Phase transformations

In general, a specific distribution between ferrite and austenite develops for each
alloying element. The gain in free transformation energy is largest when the
distribution of alloying elements reaches that of the equilibrium state. Time can be
a determining factor for the diffusion of alloying elements, especially at small
supercooling values with relatively low driving forces. An influence on the activity
of carbon, moving from austenite to carbide, is possible. During pearlite formation,
an accumulation or depletion of alloying elements can occur on the phase
boundaries, due to differences in the solubility for the alloying elements amongst
the phases.
Generally, substitutional alloying elements result in a delay in the transformation to
pearlite, corresponding to a slow-down in the pearlite formation. They lead to a
clear separation of pearlite from bainite in TTT-diagrams. Some of these elements
are, amongst others, manganese, nickel, molybdenum, chromium, vanadium, and
boron. The deceleration of pearlite formation is exploited whenever a bainitic or
martensitic microstructure with, for example, a higher hardenability, should be
produced instead of a pearlitic microstructure.

Figure 4.16: Growth rate of pearlite as a function of chemical composition at


various temperatures.
As with all austenite-stabilising elements, nickel and manganese delay the pearlite
transformation, reduce the A1-equilibrium temperature, and lead, therefore, to a
lower supercooling 'T. With regard to the mechanism of the local equilibrium
model, the lowering of A1 can be understood as a reduction of the 'Cc value, which
is the concentration difference of carbon over the width of the ferrite border, which
is proportional to a reduction in the growth rate. The influence of manganese is
principally the same as that of nickel. The reduction in growth rate is especially
great as the nickel content increases at large supercooling values (Figure 4.17).

146
4.1 Ferritic-Pearlitic Transformation

Moreover, nickel also changes the lamellar spacing (Figure 4.18). With greater
additions of nickel, the lamellar spacing becomes larger at a constant transformation
temperature, since the supercooling becomes smaller through the decreased
A1-temperature (from 723°C without Ni to 655°C for 3% Ni). The elements
molybdenum and tungsten reduce the pearlite growth rate without a significant
change in the A1-temperature. Chromium has the same effect, but it does increase
the A1-temperature.

Figure 4.17: Pearlite growth rate as a function of nickel content and supercooling
below the A1-temperature.

Figure 4.18: Lamellar spacing of pearlite as a function of transformation


temperature.

147
4 Phase transformations

4.1.6 Special forms of pearlite

Very fine lamellar pearlite (sorbite and troostite)


If the transformation rate increases greatly due to a large supercooling (cooling rates
up to 200 K/s), then the pearlite formation takes place in a temperature range with
recalescence, i.e. reheating of the sample by the released heat of transformation.
The resulting very fine lamellar pearlite, which has lamellae that are hardly visible
with an optical microscope under even the strongest magnification (Figure 4.19),
was earlier referred to as sorbite.
At cooling rates above 200 K/s, the supercooling and heat dissipitation are so large
that a reheating of the sample does not occur, and so pearlite forms at around
500°C. This very fine lamellar pearlite is formed around the maximum
transformation rate and is no longer visible through the optical microscope (Figure
4.20). It was formerly referred to as troostite.

Figure 4.19: A fine lamellar microstructure (sorbite) under light microscopy.

148
4.1 Ferritic-Pearlitic Transformation

Figure 4.20: A micro-lamellar microstructure (troostite) under the optical


microscope.

Degenerated pearlite
Pearlite formation generally proceeds according to the double-reaction mechanism.
Nucleation and nucleus growth of ferrite and cementite lamellae proceed parallely
and are, with regard to carbon distribution, coupled with one another. By slight
supercooling below the equilibrium temperature A1 and/or through the addition of
alloying elements, the reaction mechanism of the pearlite formation can be changed
such that ferrite and carbide, more or less independently of one another, precipitate
out in a non-lamellar grainy formation. Since the amount of released energy is low
for small supercooling values, the available interfacial energy also becomes smaller.
Because the interfacial energy of a sphere is smaller than that of a lamella, a
spherical structure usually forms in such cases. This creates a microstructure that is
called degenerated pearlite (Figure 4.21).

149
4 Phase transformations

Figure 4.21: A ferritic-pearlitic microstructure (degenerated pearlite) under the


optical microscope.

Non-lamellar pearlite
High alloyed steels cannot be described by a binary iron-cementite phase diagram;
according to present the multiple material systems, the carbide phase Fe3C is
replaced by one of several types of special carbides, which mostly crystallise into a
lattice structure different to that of cementite. Of the resulting special carbides,
especially M6C, Mo2C, W2C, VC, and M3C lead to non-lamellar microstructures.
Normally, these microstructures have a significantly finer structure than lamellar
pearlite and first become visible with an electron microscope. The microstructure is
no longer called pearlite, but rather ferrite and carbide (Figure 4.22).

Figure 4.22: A ferritic microstructure with globular cementite (non-lamellar


pearlite) under the optical microscope.

150
4.1 Ferritic-Pearlitic Transformation

4.1.7 Application of ferritic–pearlitic steels


There is a wide range of use for pearlitic microstructures in the field of (pearlitic)
railway steels, where materials with high pearlite content generally offer good
resistance against the roll-wear occurring between wheel and rail. A suitable heat
treatment improves these steels by reducing the lamella thickness. The smaller
lamella thickness leads to a better notch-impact toughness and a reduction in area
value, which ultimately minimises wear. This is exploited, for example, on high-
speed routes through the use of steels called “head hardened rail steels”, with an
approximately eutectoid composition and a micro-lamellar microstructure formation
in the rail head. The rolled microstructure of steels usually consists of ferrite and
pearlite.

151
4 Phase transformations

4.2 Martensitic (athermal) Transformation


4.2.1 Morphology
At large supercooling values, the phase transformation experiences an increased
restriction to diffusion as the temperature sinks. Since the driving force behind the
phase transformation simultaneously increases, entire areas of austenitic crystals
suddenly shear into a ferritic state, without changes in the local alloy composition
occurring due to diffusion processes. The crystal structure which forms is called
martensite, named after the German scientist A. v. Martens, 1850-1914. This man
directed the Royal Materials Testing Office, which was founded in 1904 and was a
predecessor to the present Institute for Materials Testing and Research in Berlin
(Bundesanstalt für Materialprüfung, BAM).
The changes in volume and in configuration associated with this transformation
cause strong elastic lattice distortions, which naturally counteract the
transformation. At high carbon content values (> 1.4%) and low transformation
temperatures, these distortions can be reduced only to a small degree through
gliding and recovery mechanisms, so that the transformation from austenite to
martensite does not always fully proceed. Metallographically, martensite forms in
plates. Before etching, the transformed crystalline regions are already recognisable
on a polished surface as clearly outlined relief structures. The sheared structure of
the martensite crystals in the “retained austenite”, which has not yet transformed,
becomes visible after etching as shown on the left in Figure 4.23.
At lower carbon content values (<< 0.5%) and higher transformation temperatures,
the elastic lattice stresses can be removed more easily through dislocation
movements. Martensite then no longer consists of single crystal plates but rather of
lath-shaped crystals that are bunched into blocks and oriented in various directions.
Lath martensite is shown on the right in Figure 4.23. Laths are several microns long
and are 0.1 to 0.5 microns thick. At an average carbon content (0.5 to 1.4%), both
stress-reducing mechanisms can take place.
4.2.2 Formation of the martensitic microstructure
Martensite transformations, which lead to the microstructures shown in Figure 4.23,
have a series of characteristics that clearly distinguish them from diffusion-
controlled transformations:
1. They usually occur athermally at low temperatures. They begin suddenly
when a certain temperature is reached during cooling and proceed cascade-
like in a fraction of a second when cooling continues. If the cooling is
stopped, then the further transformation also stops, although “para”-
equilibrium has not yet been reached.
2. Single atoms do not exchange places, instead only clusters of atoms
exchange places in a coordinated, coupled movement (cooperative shear

152
4.2 Martensitic (athermal) Transformation

movement of atoms). These complex lattice deformations are achieved


through shear and tensile forces. Since no diffusion takes place, the
positions of the iron and carbon atoms relative to one another remain the
same during the shearing of the fcc-austenitic lattice to the bcc-martensitic
lattice (Figure 4.24).
3. A relationship between the crystallographic orientation of the cubic face-
centered lattice of austenite and the cubic body-centered martensite exists.
4. The coordinated atomic movement causes an increase in volume
(approximately 3% in cubic body-centered martensite) and a marked
alteration in shape of the transformed crystal region (Figure 4.24). These
alterations in shape are noticeable as a relief structure on the surface.
5. Elastic distortions in the crystal volume result from the shape alterations.
These distortions can be reduced through slipping or twinning. Due to the
elastic distortions and the high displacement density or twin border density,
the martensitic phase is harder than the mother phase.

a) Plate martensite b) Lath martensite


Figure 4.23: Different morphologies of martensite in Fe-C alloys.

153
4 Phase transformations

Tensile

Shear

Habit plane

= fcc = bcc
Figure 4.24: Transformation of the cubic face-centered lattice to the cubic body-
centered lattice through shear and tensile forces.

Nucleation and nucleus growth


If the cooling rate is large enough to stop all diffusion-controlled phase
transformations, then austenite and martensite have the same free enthalpy at a
concentration- and pressure-dependent temperature T=TE (Figure 4.25).
However, to begin the martensitic transformation a certain nucleation enthalpy
('GA-M) is necessary. In order to supply this energy, a supercooling over the
temperature difference 'T is necessary. The temperature at which the martensite
transformation begins is called the martensite start temperature M s. It is
independent of the cooling rate, but is noticeably influenced by increased carbon
content values and alloying additions.
Contrary to other transformation steps, during martensite formation, no incubation
period worth mentioning can be determined. Probably, “pre-formed” martensite
nuclei form in the metastable austenite on the existing displacements through
displacement reactions. They are enlarged through the growth and formation of
“transformation displacements,” which result in a transformation of the austenitic
lattice to the martensitic lattice. The time needed for a martensite crystal to form is
10-7s.

154
4.2 Martensitic (athermal) Transformation

GM= Volume enthalpy


of martensite

'GA-M GA= Volume enthalpy


of austenite
Enthalpy GA, GM

TE= Equilibrium temperature


between the two phases
'G A-M=0

GM
'T= Supercooling
'T
GA MS = Martensite start temperature
(beginning of martensite formation)

MS TE Temperature
Figure 4.25: Graph showing the influence of temperature on the free enthalpy of
austenite and martensite.
Figure 4.26 shows the lath-package structure of a former austenite grain
schematically. Such a grain is divided into several packages with parallel lancets.
Again these packages are subdivided in blocks that can contain two groups of
interspersed K-S-laths.

Figure 4.26: Schematic illustration of the block-package structure of a prior


austenite grain.

155
4 Phase transformations

In the following Figure 4.27 the microstructure of a 9Ni steel is illustrated. In the
austenite area the material is annealed and subsequently cooled. The color
illustration in the lower part of the figure was measured by EBS IPF mapping. The
black and white illustration depicts a 3D figure by means of LOM after etching with
1% Nital.

Figure 4.27: Microstructure of 9Ni steel


Color: EBS IPFD mapping
Black and white: 3D figure by LOM after etching with 1% Nital

The crystallographic model of martensite formation


Martensite formation at relatively low transformation temperatures leads to high
transformation stresses because of the already considerable yield criterion value of
austenite. In order to at least keep the unavoidable elastic volume and boundary
area energies to a minimum, a special mechanism to describe the formation of
martensite is required. A martensite crystal, in its typical form at higher carbon
percentages, forms in sheets in order to achieve an energy minimum. The
crystallographic model described here should specify a transformation process, in
which:
1. The face-centered cubic lattice of the initial phase with the known lattice
constant aA converts to the body-centered cubic lattice of the new phase with
the known lattice constant aM, and
2. The plate-shape of the new phase supplies an invariant habit plane, meaning
at least one plane of separation between austenite and martensite that remains
undistorted in itself and does not rotate in space during the transformation.

156
4.2 Martensitic (athermal) Transformation

With these conditions, the austenite-martensite transformation can be divided into


two steps: the lattice-changing deformation (Bain) and the lattice-retaining
deformation (invariant deformation).
E. C. Bain formulated the “Bain model” (Figure 4.28). With the help of this model,
it is possible to fulfill the first demand for the lattice-changing deformation. The
starting point is a pair of face-centered cubic unit cells, with the lattice constant a A,
that already contain a virtual tetragonal space-centered lattice (Figure 4.28a).
The virtual tetragonal unit cell has the lattice constants cV = aA and aV = aA/ 2
(Figure 4.28b). These measurements do not, however, agree with the real lattice
measurements of martensite. In order to obtain the exact lattice parameters of
martensite, the c-axis must be compressed by approximately 20% through a
homogeneous deformation, and the a-axis must be enlarged by 12% (Figure 4.28c).
A tetragonal distorted martensite forms with the constants c M and aM. During
transformation, the original coordinate system XYZ is rotated to the coordinate
system X’Y’Z.
The tetragonal lattice distortion is supported by the occupancy of some octahedral
vacancies in the iron lattice with interstitial dissolved carbon atoms. Since carbon
atoms prefer to occupy the larger octahedral vacancies in the z-plane, martensite
becomes more tetragonal as the carbon content increases.
While the Bain model sufficiently describes the deformation of the crystal lattice
during the transformation from the fcc to the tetragonally distorted bcc lattice, the
description does not, however, contain an invariant habit plane. Since such a plane
is found experimentally, a further deformation step must follow. This second
transformation step must have the characteristic that the structure of the previously
produced martensite lattice is preserved.
Crystal slipping through dislocation and twinning are such lattice-retaining or
lattice-invariant alterations in shape. Figure 4.29 schematically shows both of
these variations of lattice-retaining deformation.
Through crystal slipping and twinning, it is possible to compensate the lattice
distortion that would be produced by pure lattice-altering transformations, so that
the second requirement for an invariant habit plane is fulfilled.

157
4 Phase transformations

Figure 4.28: The transformation of austenite into martensite according to the Bain
model: a) two face-centered cubic cells (austenite) with embedded
tetragonal cell; b) tetragonal cell from a); c) martensite cell.
Both deformations can be macroscopically described as shearing, in which the
shearing plane is the same as the slipping- or twinning plane and the direction of
shearing is the same as the slipping- or twinning direction. Figure 4.30 offers an
overview of the possible lattice distortions from transformation shearing.

Figure 4.29: Formation of martensitic plates, shown are two types of lattice-
retaining deformations: a) slipping and b) twinning.

158
4.2 Martensitic (athermal) Transformation

Bain-transformation

Slipping

Twinning

Figure 4.30: Types of lattice distortions from transformation shear.


Figure 4.31 shows the calculations of the volume change during martensite
transformation (Fe+1%C).

Figure 4.31: Equations of the volume change during martensite transformation


(Fe+1%C).
From a crystallographic point of view, the orientational relationships between
austenite and martensite as well as the habit plane are of special importance.
Table 4.3 shows some of these relationships.

159
4 Phase transformations

The first orientational relationship describes the transformation as specified by the


theoretical Bain model. According to this relationship, from one austenite crystal,
only three martensite crystals could be created.
Experimentally, the Kurdjomov-Sachs relationship was discovered first. In
association with a habit plane close to {111}A, it is valid for low-alloyed steels with
a mass content of carbon up to 0.5%. At levels of up to 1.4% carbon, this
relationship is coupled with a {225}A habit plane.
Table 4.4 shows the features of α´ martensite in Fe-Ni alloys.

Table 4.3: Crystallographic orientational relationships due to martensite


formation.

Martensite Thin Plate Lenticular Butterfly Lath


Types

D’ D’ D’
J D’
Morphology J
J

Formation
temperature, Low temperature High temperature
Ms
Accommodation Twin and single slip Twin and dislocation slip on Multiple slip
Twin
process dislocation limited slip dislocation
{3 10 15}γ {5 5 7} γ
Habit plane {3 10 15} γ {2 2 5} γ
{2 5 9} γ {1 1 1} γ
Orientation relations
G-T N-W, G-T or K-S N-W, G-T or K-S K-S
with γ

Table 4.4: Features of α´ martensite in Fe-Ni alloys.

160
4.2 Martensitic (athermal) Transformation

Morphology Substructure Habit Orientation Ms


plane Relationship*

Lath (Tangled) (111)A K-S High


dislocations
Butterfly (Straight) (225)A K-S
dislocations
and twins
Lenticular (Straight) (259)A N-W
dislocations or or
and twins (3 10 15)A G-T
(Mid-rib)
Thin-plate Twins (2 10 15)A G-T Low

* K-S: Kurdjumow-Sachs relationship, N: Nishiyama relationship, G-T: Greninger-Troiano relationship

Table 4.5: Connections among microstructures, habit planes and orientation


relationships for four kinds of α´ martensite.
Table 4.5 lists the connections among microstructures, habit planes and orientation
relationships for four kinds of α´ martensite.
As shown in Figure 4.32, the (110)-surface of martensite lays parallel to the (111)-
surface of austenite. Similarly, the [111]-direction of martensite is parallel to the
[110]-direction of austenite. As a result of this, 24 martensite crystals, each with a
different orientation, can be created from one austenite crystal.

aA
<110>A
<111>M
(111) (110)M
A

Figure 4.32: The crystallographic relationship according to Kurdjomov-Sachs.


In steels with a high carbon content, or in substitutional solid solutions, for example
Fe-Ni alloys, the Nishiyama-Wassermann relationship can be observed. The habit

161
4 Phase transformations

plane in this case is {259}A. Twelve different variations of martensite exist for this
orientational relationship.
It follows from the invariant habit plane, that the total deformation can consist of
only one change in length perpendicular to and one shearing parallel to this plane.
Quantitatively, this change in length is equal to the increase in volume during the
austenite-martensite transformation. It is without exception around +3 %. The
macroscopic shearing angle is calculated and measured for most alloys in the
magnitude of around 10°.
If the martensite formation proceeds at higher temperatures, i.e. with a low alloying
content, then the basic unit of the martensitic microstructure is no longer plate-
shaped, it is instead made of flattened laths that are packed very closely into layers
along the {111}A plane and furthermore, into solid blocks. Accordingly, this type
of martensite is called lath, block, or massive martensite.
This variation of martensite formation can be explained by the fact that the elastic
lattice stresses formed at the high transformation temperatures are reduced on a
large scale through slipping and recovery processes.
Figure 4.33 clarifies these circumstances. At a low carbon content, the M s-
temperature is relatively high, so that the starting stress for slipping is smaller than
that for twinning. Consequently, for the lattice-retaining deformation in austenite
and martensite, only slipping processes, which are necessarily associated with an
increase in the displacement density, come into question.
On the other hand, at a sufficiently high carbon content and a relatively low Ms-
temperature, the starting stress for twinning is smaller than that for slipping, so that
twinning becomes the decisive process during the lattice-retaining deformation of
plate martensite.

162
4.2 Martensitic (athermal) Transformation

Ms(c2) Ms(c1)

c1<c2

Initial stress Twinning

Plate Lath Slipping


martensite martensite

Temperature
Figure 4.33: Influence of the temperature dependency of the stress necessary for
slipping and twinning on the martensite morphology at various
transformation temperatures, carbon content c1 < c2.
4.2.3 Influence of alloying elements on martensite formation
The composition of the alloy and the way the alloying elements are included, as
either substitutional or interstitial alloying elements in the unit cell, determine a
series of properties that can decisively influence the morphology of martensite.
Most alloying elements that go into solution in austenite lower the Ms-temperature,
with the exception of cobalt and aluminum. The interstitial dissolved elements,
such as carbon or nitrogen, have by far the largest effect. The effect of interstitial
and substitutional dissolved elements on the martensite start temperature is
described by the empirical equation:
M s (qC ) 550  350(%C )  40(%Mn)  35(%V )  20(%Cr )  17(%Ni )
(4.14)
10(%Cu )  10(%Mo)  8(%W )  15(%Co )  30(% Al)

Influence of carbon content on martensite formation


When carbon is alloyed to iron, the transformation process from the J- to the D-
lattice is hindered by an impairment in the compression process. During this lattice-
changing deformation, the impairment occurs in the direction of the c-axis.
Figure 4.34 shows the respective lattice constants as dependent on the carbon
content. The tetragonal body-centered martensite cell can be quantitatively
described by the axis ratio cM/aM:
c M /a M 1.000  0.046 C% (4.15)

163
4 Phase transformations

This equation requires that the body-centered lattice structure is distortion-free at a


carbon content of zero mass-%.
While the extension of the austenite lattice progresses isotropically with increasing
carbon content (Figure 4.34), the extension of the lattice constants of the body-
centered cubic lattice is anisotropic. The constant c M increases strongly with
increasing carbon content, whereas the constant a M decreases. This behavior is due
to the direction-dependent occupation of the lattice vacancies in iron.

Figure 4.34: Lattice constants of austenite, tetragonal martensite, and axis ratio c/a
of tetragonal martensite as dependent on carbon content.
In both the face-centered cubic and the body-centered cubic lattices, the carbon
atoms occupy the octahedral vacancies. They are, then, surrounded by the six iron
atoms that are their closest neighbors. Figure 4.35 shows the position of the
octahedral vacancies in a martensite lattice. In the face-centered cubic lattice of
austenite, the six iron atoms form an isotropic octahedron around the interstitial
included carbon atom. In contrast, the octahedrons of the body-centered cubic

164
4.2 Martensitic (athermal) Transformation

lattice are shortened along the z-axis. The inclusion of carbon atoms in the
vacancies along the z-axis is, therefore, associated with a significantly larger change
in the atom spacing of iron than inclusions in the vacancies along the x- or y-axis.
Due to the diffusionless transformation, the carbon atoms retain their positions
during the lattice-changing deformation. This leads to a distortion of the body-
centered cubic lattice in the z-direction. The z-axis is associated with the lattice
constant cM of the tetragonally distorted body-centered martensite.

y z y Fe - atom
x
C - atom
cM
Octahedral vacancy
y z
z
x x
y z
z z

x
x
y z y y
aM
x

aM

Figure 4.35: Octahedral vacancies in the tetragonally distorted body-centered cubic


martensite lattice. Vacancies in the z-direction are fully occupied
with carbon atoms.
Figure 4.36 shows the influence of carbon content on the relative change in volume
during martensite formation. Since the face-centered cubic J-iron is atomically
more closely packed than the body-centered cubic D-iron or the tetragonal body-
centered martensite, a growth in volume occurs during the transformation from
austenite to martensite. As a result of the increasing tetragonality, with a raised
carbon content, the relative change in volume increases as well. Due to the
increased lattice distortions, a stronger supercooling becomes necessary to supply
the required enthalpy of volume for the transformation of martensite to proceed.
Consequently, the Ms- and Mf-temperatures are shifted to lower temperatures as the
carbon content increases.

165
4 Phase transformations

3.8

Relative change in volume in %


3.6

3.4

3.2
0 0.4 0.8 1.0

Carbon content in mass-%


Figure 4.36: Influence of carbon content on the relative change in volume during
martensite formation.
Figure 4.37 depicts the martensite start temperature Ms over the carbon content of
unalloyed steels. The martensite finish temperature Mf describes the end of the
transformation. The shaded area indicates the scatter band of the respective
temperature.
At a carbon content of approximately 0.5 mass-%, Mf reaches room temperature.
Therefore, at any carbon content larger than 0.5 mass-%, the martensite formation is
incomplete when room temperature is reached, and as a result, metastable retained
austenite appears.
The lower part of the figure schematically shows how the formation of martensite
progresses, with respect to the volume proportion, at low and high carbon contents
during continuous cooling. The transformation takes place discontinuously within
small ranges of volume and within very short periods of time, such that there are
temperature and time intervals where no transformation takes place. The
transformed sample volume at any point in time is clearly a function of the
temperature reached between Ms and Mf. It is decisive that, at low levels of carbon,
a complete austenite transformation occurs, while at high levels of carbon, the
austenite transformation does not progress to completion, even with cooling to very
low temperatures.
The percentage of martensite at higher carbon contents remains less than 100%,
even at temperatures below Mf. Alongside martensite, retained austenite is present
in the microstructure. If cooling is interrupted and then continued between M s and
Mf, a decrease in the Mf temperature can result. This process of reducing the Mf is
caused by thermal stabilisation, which probably occurs due to the fact that the

166
4.2 Martensitic (athermal) Transformation

interstitial dissolved carbon atoms form Cottrell clouds during holding and
temporarily block potential martensite nuclei.
800

Temperature in °C 600
Martensite start temperature
400

200
Martensite finish temperature
0

-200
0 0.5 1.0 1.5
Carbon content in mass-%

100 100
Martensite fraction in Vol.-%

50 50

0 0
Ms Ms
20°C Mf Mf 20°C
Temperature in °C
Figure 4.37: Influence of carbon content on the Ms- and Mf-temperatures as well as
the influence of transformation temperature on the fraction of
martensite for low and high carbon contents.
In reference to the metallographic appearance, with an increase in carbon content a
transition from lath to plate martensite occurs (Figure 4.38). In steels containing up
to approximately 0.5 mass-% carbon, pure lath martensite is present. Between 0.5
and 1.0 mass-% carbon, a transitional range exists consisting of lath and plate
martensite. At more than 1.0 mass-% carbon, plate martensite is exclusively
present. At the same time, the amount of retained austenite increases when
martensite transformation does not completely progress.

167
4 Phase transformations

Figure 4.38: Influence of carbon content on the relative volume fraction of lath and
plate martensite (LM, PM) in Fe-C alloys, with consideration to Ms-
temperature and retained austenite fraction (RA).

Figure 4.39: Driving force necessary to form lath and plate martensite.
Seen schematically, the transition from lath to plate martensite can be indicated by
the influence of the alloying content on the driving force necessary for martensite
formation (Figure 4.39). As the alloying content rises, the required driving force
continuously increases. For plate martensite, however, the increase is slower, such

168
4.2 Martensitic (athermal) Transformation

that this form is energetically favorable beginning at a certain level of alloying


content.

Influence of other alloying elements on martensite formation


Figure 4.40 shows the influence of the element nickel and the cooling rate on the
formation of martensite. Each curve is characterised by two plateaus. The upper
plateau is attributed to the massive transformation, and the lower plateau to the
martensitic transformation. The respective critical cooling rates are signified by the
grey regions. They decrease as the alloying content increases, while the Ms-
temperature simultaneously decreases, too. The massive transformation also
occurs without diffusion, but in contrast to the martensitic transformation, it lacks a
defined crystallographic relationship and a change in form. Both transformations
have in common that the chemical composition does not change.

Figure 4.40: Transformation temperature as a function of cooling rate and alloying


content of nickel.
The tetragonal distortion of the martensite lattice in Fe-C alloys is altered by the
addition of substitutional dissolved alloying elements. An example of this is given
in Figure 4.41 where the dependence of cM/aM on the carbon content is shown in
comparison for Fe-Ni-C-, Fe-Al-C-, Fe-Mn-C- and Fe-C-alloys.
It is assumed that only in the alloys Fe-Ni-C and Fe-Al-C with relatively high Ni- or
Al-contents all dissolved carbon atoms are to be found in the z-direction. In
contrast, in Fe-C-martensitic alloys, approximately 80% of the dissolved C-atoms
should be found in the z-direction, with the rest appearing in the x- and y directions.
In Fe-Mn-C alloys, less than 80% of the carbon atoms are expected in the z-
direction.

169
4 Phase transformations

The observed decrease in the cM/aM-ratios in Fe-Mn-C alloys, in comparison with


Fe-C alloys, can be explained by an increase in the volume fraction that is twinned.
In contrast, the very large tetragonality of the Fe-Ni-C and Fe-Al-C alloys, as
compared to the Fe-C-alloys, is clearly related to a long-range ordering of the
substitutional dissolved atoms, which strongly restricts, or even fully suppresses
twinning in austenite.

Figure 4.41: Influence of carbon content on the cM/aM-ratio of martensite in ternary


iron-based alloys.
Relative change in volume in %

0.6

0.5

0.4

0.3
Si
Mn
0.2 Ni
Cr
0.1 Mo
Co
0 1.0 2.0 3.0 4.0 5.0
Alloying element content in mass-%
Figure 4.42: Relative change in volume during austenite-martensite transformation
as a function of alloying element content.

170
4.2 Martensitic (athermal) Transformation

In Figure 4.42 the influence of substitutional alloying elements on an increase in


volume is shown for the transformation of austenite to martensite.
4.2.4 Influence of deformation on martensite formation
The influence of stress on the martensite transformation expresses itself as an
increase in the Ms-temperature. The stress field of a plastic or elastic deformation is
superimposed on the stress field of the martensite nucleus, and strengthens it.
Deformation leads to an increased martensite start temperature, which is given as
Md. Above this temperature, austenite deformation does not lead to martensite
formation (Figure 4.43).

Figure 4.43: Dependence of martensite start temperatures Ms and Md on Ni-content


(steel with 0.04% C, 18% Cr, Ni up to 12%), degree of deformation at
Md -measurement amounts to 50%.
If the Md-temperature of steel is high, then one speaks of low austenite stability or a
high martensite formation activity. This Md-temperature is difficult to determine
experimentally, which led to the introduction of the M d30-temperature. At this
temperature, an D´-martensite fraction of 50% is reached in the single-axis quasi-
static tensile test at 30% deformation. This value is used almost exclusively to
quickly determine the austenite stability in austenitic stainless steels, for example
X5CrNi18-10. Table 4.6 gives selected examples of the approaches developed by a
few authors for the calculation of this characteristic temperature. The approaches
differ above all in their estimation of the effect of the alloying elements nickel,
niobium, and copper. The different valuations of the element nickel can, thereby,
lead to varying results for the Md30-temperature. According to Sjöberg, on the other
hand, the alloying elements nickel and nitrogen weigh heavily in the determination
of the Md30-temperature.
This effect is of interest for high-alloyed stainless austenitic chromium-nickel
steels, as well as for newly developed automobile steels (known as TRIP steels).
The extent of deformation that is necessary to initiate the austenite-martensite

171
4 Phase transformations

transformation at room temperature can be enlisted as stability criteria for austenite.


If a cold deformation of more than 80% is necessary to produce a certain amount of
martensite in a steel, then this steel can be seen as austenitically stable. The
martensitic transformation is supported above all by tensile stresses and to a lesser
degree by compression stresses. In contrast, hydrostatic pressure makes the
martensite transformation more difficult, because it counteracts the characteristic
increase in volume.
Another value has been accepted for describing austenite stability in low-alloyed
TRIP steels (for example, TRIP700 with approximately 0.2% C, 1.5% Mn and Si-,
Al-alloying additionally). In this case, the M sV-temperature is used, which can be
determined with the help of the Bolling-Richman method. This temperature defines
the boundary temperature between the actions of the stress induced transformation
mechanism i.e. the transformation of retained austenite to martensite up to the yield
point, and the deformation induced transformation mechanism, i.e. transformation
in the plastic range. By means of Figure 4.44, the single sample method according
to Bolling and Richman will be briefly discussed.
Author Formula Unit
Md30=413-462*(%C+%N)-9.2*(%Si)-8.1*(%Mn)-
Angel °C
13.7*(%Cr)-9.5*(%Ni)-18.5*(%Mo)
Md30=497-462*(%C+%N)-9.2*(%Si)-8.1*(%Mn)-
Gladman °C
13.7*(%Cr)-20*(%Ni)-18.5*(%Mo)
Md30=551-462*(%C+%N)-9.2*(%Si)-8.1*(%Mn)-
Nohara °C
3.7*(%Cr)-29*(%Ni+%Cu)-18.5*(%Mo)-68*(%Nb)
Nohara, Md30=Md30(Nohara)-1.42*(GB-8), (GB corresponds
°C
(modified) to the grain size number according to ASTM)
Md30=608-515*(%C)-821*(%N)-7.8*(%Si)-12*(%Mn)-
Sjöberg °C
13*(%Cr)-34*(%Ni)-6.5*(%Mo)

Table 4.6: Approaches of various authors to calculate the Md30-temperature.

172
4.2 Martensitic (athermal) Transformation

Figure 4.44: Single specimen method according to Bolling-Richman to determine


the MsV-temperature (shown here for the steel TRIP700).
In this test, a tensile test specimen is loaded at a defined temperature to the point at
which it begins to yield plastically. The load is then removed from the tensile test
specimen, the test room is cooled by 10°C and the tensile sample is strained again to
the point of plastic yield. The pre-deformation and decreased temperature results in
a higher yield stress. Furthermore, at a defined temperature (here, at 20°C), a
characteristic progression of the stress-elongation curve becomes apparent, which
resembles the appearance of a distinct yield point with a lower and upper yield
point. This change in the progression of the stress-elongation curve around the
yield point is induced by the transformation of retained austenite into martensite.
The decrease in strength is the result of shearing due to the austenite-martensite
transformation. The test temperature, at which the transition from the continuous
yield point meets the upper and lower yield point, is the MsV-temperature.
4.2.5 Mechanical properties of martensitic steels
The increased hardness and brittleness of steels after quenching from austenite is
the result of martensite formation. In this regard, the hardness is a function of the
carbon content of the steel. In non-iron alloys, as well as iron-nickel alloys,
martensitic transformation does not lead to such great increases in strength.

173
4 Phase transformations

900

65
800

Rockwell hardness (HRC)


700 60

Vickers hardness (HV)


Martensite in
Fe-C alloys
600 55

500 50
45
400 40
35
300 30
25
Austenite in 20
200
Fe-Ni-C alloys

100

0 0.2 0.4 0.6 0.8 1.0


Carbon content in mass-%
Figure 4.45: Influence of carbon on the hardness of martensite and austenite.
Figure 4.45 shows the large influence of the carbon content on the hardness of
martensite, as compared to the relatively low effect of carbon content on the
hardness of austenite that is stable at room temperature. The shaded region of the
hardness curve of martensite signifies the influence of the increasing retained
austenite content.
The great hardness of martensite has many causes. Decisive is the complex
structure of martensite. The tetragonal body-centered lattice with interstitial
dissolved carbon that is created through shearing leads to a high dislocation density
and twinning.
The high strength of martensite is determined through:
strengthening through elastic stresses,
strengthening by dislocations,
strengthening through lath, plate, and twinning boundaries,
strengthening of the solid solution by interstitial dissolved atoms.
The main causes of the high strength of martensite are strengthening of the solid
solution and strengthening by dislocations (Figure 4.46). Up to approximately
0.2% carbon, the improvement in strength is accomplished mainly through a high
displacement density, since the greater part of the carbon precipitates out as
carbides due to the high temperature of martensite formation. At high carbon
content levels, the hardening of the solid solution through carbon atoms takes on

174
4.2 Martensitic (athermal) Transformation

more importance, because it lowers the martensite formation temperature, and the
fraction of dissolved carbon in the martensite lattice increases.

Figure 4.46: Strengthening by solid solution of carbon and by dislocations in


martensite (C and Ni are adjusted such that a constant M s-temperature
exists).
For practical purposes, the quenched hardness of martensitic steels is of importance.
Figure 4.47 shows the progression of the (martensite) hardness of iron-carbon
alloys (the influence of alloying elements can be considered negligible) in the iron-
rich side of the iron-carbon diagram. This graph (also called "Burns Curve") shows
that hardness steadily increases until a peak is reached by quenching from the
homogenous austenite field to the eutectoid composition. The fall in the curve at
higher levels of carbon content (in the hypereutectoid field) can be explained by an
increasing amount of (softer) retained austenite in the microstructure. During
quenching in water, the temperature does not fall below the M f temperature because
of the high carbon content of austenite. The hardness can be improved if the
temperature is dropped below Mf through deep-quenching. The hardening of
hypereutectoid steels, as commonly carried out industrially, takes place from the J +
Fe3C two-phase field. The hardened microstructure then consists of martensite and
very hard cementite.

175
4 Phase transformations

Figure 4.47: Hardness of martensite as well as Ms- and Mf-temperatures as a


function of carbon content; the shaded area signifies the possible
existence of retained austenite.

176
4.2 Martensitic (athermal) Transformation

4.2.6 Special forms of martensite and martensite formation


Cubic martensite
Alongside the mostly tetragonally distorted body-centered cubic lattice of the iron-
carbon martensite, cubic martensite can also be found in a quenched microstructure.
The cubic martensite appears when a low content of interstitial alloying elements is
present, but most especially in iron-nickel, iron-manganese, iron-chromium, and
other binary and ternary iron alloys containing substitutional elements.

Hexagonal (HH-) martensite


Hexagonal or H-martensite (Figure 4.48) occurs in steels containing more than 10%
manganese. It also forms during cooling or after a plastic deformation in carbon-
containing alloys of iron with nickel, manganese, chromium, and cobalt.

Figure 4.48: Hexagonal martensite.


The H-martensite possesses a hexagonal closed-packed structure with the following
axis ratio:
c M /a M 1.616 (4.16)
The transformation from the cubic face-centered austenite lattice to the hexagonal
lattice is associated with a reduction in volume of about 2.1%. This is in contrast to
the behavior usually observed during martensite transformation.

Thermoelastic martensite
In alloyed steels beside Fe-C martensite, reversible martensite transformations
appear as well. Here, the martensitic phase is in thermal equilibrium with an
austenitic phase. The austenite-martensite transformation during cooling begins
when the Ms- temperature is reached, and the martensite-austenite transformation
begins during heating when the As-temperature is reached.
Usually, the martensite-austenite retransformation first begins after a strong
superheating above the equilibrium temperature TE. This retransformation often

177
4 Phase transformations

progresses, similarly to the martensite formation, with a coordinated atomic


movement. Figure 4.49 shows the temperature hysteresis of the austenite-
martensite transformation cycle, which occurs with the removal of elastic lattice
distortions through slipping processes. Deformation facilitates the transformation
so that the hysteresis decreases. The deformation-induced martensite start
temperature Md is higher than Ms, and the deformation-induced austenite start
temperature Ad is lower than As.

Figure 4.49: Change in electrical resistance as a function of temperature during a


reversible Fe-Ni martensite transformation.
If the lattice distortions produced during transformation cannot be removed through
plastic deformation, then martensite remains coherently connected with the
surrounding austenite. This can be seen in, for example, Fe-Pt and Fe-Ni alloys.
During a long annealing treatment, a superlattice with higher strength forms. The
transformation and retransformation of martensite then occurs practically without a
temperature hysteresis. When the temperature is raised, the martensite formed
during cooling transforms immediately back into austenite as a result of elastic
stress, although the test temperature is still below the equilibrium temperature T E.
This type of behavior is called thermoelastic.

Isothermal martensite formation


Most of the martensite transformation occur athermally. However, there are a few
examples of isothermal martensite formation. It is thus observed, in some alloys,
that an isothermal martensite formation begins after completion of cooling and
holding. This type of martensite formation can usually be seen in steels and alloys
with low martensite start temperatures, around and even below room temperature,
for example in binary iron-nickel alloys with 30% nickel.

178
4.2 Martensitic (athermal) Transformation

It is possible to completely suppress the athermal martensite transformation in iron-


nickel-manganese alloys by rapidly quenching to -200°C. Martensite forms during
isothermal holding at intermediate temperatures. Transformation diagrams can be
determined for this type of formation, which never progresses to completion as
shown in Figure 4.50.
-80

-100
Temperature in °C

-120

1% 5% 10% 20% 30%


-140

-160

First fractions of
-180 martensite
(0.2%)
-200
10 50 100 500 1000 5000
Time in s
Figure 4.50: Time-temperature-transformation diagram for isothermal martensite
formation in a Fe-Ni-Mn alloy.
4.2.5 Application of martensitic steels
Steels with a martensitic microstructure have a multitude of different applications.
This is due the high hardness that can be realized by this type of microstructure. In
the following, two examples of martensitic steels are described.
Cutlery is among the items of everyday use. Martensitic chrome steels are used as a
standard material for blades. Figure 4.51 shows a hollow-handle knife.

blade
cutting edge back handle
Figure 4.51: Example of a hollow-handle knife: the blade possesses a martensitic
microstructure, whereas the handle is austenitic.

179
4 Phase transformations

This knife contains a forged blade, which is set in the handle and made of a
martensitic chrome-steel (material number 1.4034; (| 0.45% C, 14% Cr)). Because
of the high chromium content, these steels are among the acid and heat resistant
stainless steels. However, they are not weldable, and after a heat treatment,
martensite appears as the transformation microstructure. The hardness of this
martensite is over 250 HV10. The handle itself is made of a soft, easily deformable
material. In most cases, an austenitic chromium-nickel steel is used (material
number 1.4301; (| 0.035% C, 17.5% Cr, 9% Ni)).
Another example is taken from high strength steels, which can have tensile
strengths over 1200 MPa (approximate). The Marageing steels belong to the high
strength steels. These steels contain 18% nickel, 8 to 12% cobalt, 4 to 5%
molybdenum and 0.4 to 1.6% titanium.
A special feature of the iron-nickel phase diagram is exploited in setting the
appropriate properties. Already during slow cooling from the austenitising
temperature (approximately 800°C), carbon-poor alloys with around 18% nickel
transform to a martensitic microstructure of nearly cubical form at a temperature of
around 250°C. The retransformation to austenite occurs during heating up again
first at temperatures above 500°C.
The previously described transformation hysteresis (Figure 4.49) allows for the age
hardening of steels, which begins at a temperature of around 480°C, just before
initiation of the retransformation. When the steel is quenched, the martensite
created is supersaturated with alloying elements. Heating leads to ordering
processes and pre-precipitations, which cause considerable age hardening. The
advantage of age hardening is that the material gains high strength while retaining
good toughness properties.
Such Marageing steels are used in the aerospace industry for tool and machine parts
that are placed under high demands. Examples can also be found in the recreational
industry (Figure 4.52).

Figure 4.52: Head of a golf club, the striking surface consists of a Maraging steel
plate.

180
4.3 Bainitic Transformation

4.3 Bainitic Transformation


4.3.1 Morphology
Bainite, a two component microstructure decomposed from austenite, forms at a
temperature between the pearlite and martensite range (550 to 250°C). In the range
overlapping with pearlite, the formation of bainite resembles that of pearlite; and in
the area overlapping with martensite, the formation resembles that of martensite.
Therefore, it is understandable that there are continual transitions during the
microstructural development from the pearlite range to the bainite range, as well as
from the bainite range to the martensite range. Since bainite was discovered in
1930, three definitions are proposed, namely the (1) microstructural, the (2) overall
kinetics and the (3) surface relief definition. According to these definitions bainite
shows the following characteristics:
x Diffusional, non-cooperative, competitive edgewise growth of two
precipitate phases formed during eutectoid decomposition where the
minority phase appears in non-lamellar form;
x Own C-curve on a TTT-diagram, lying partially below but extensively
overlapping that of pearlite in Fe-C and some Fe-C-X alloys, lying entirely
below the pearlite C-curve in appropriately alloyed steels;
x Consists of precipitate transformation plates, resulting in an invariant
plain strain (IPS) surface relief effect, which forms by shear, i.e.
martensitically, at temperatures usually above the martensite start
temperature Ms.
The main difference between the three definitions is that for the first one, two
product phases (e.g. ferrite and cementite) are required, whereas according to the
definitions (2) and (3) bainite can consist entirely of ferrite.
Although many of the characteristics of bainite especially the morphology and the
shape deformation were found to be similar to those of martensite, the bainitic
transformation is unlike the martensitic one as it also involves partitioning of
carbon. This leads to the controversial discussions on whether the transformation
mechanism is diffusion-controlled or displacive. In any case bainitic phase
transformations consist of 2 steps: J Æ D transformation mostly by shear + C
diffusion that leads to C partitioning between J and D or the precipitation of
carbides. The sequence of the 2 steps differs between low C and high C steels. In
low C steels the transformation takes place first and C diffusion occurs in the
bainitic ferrite leading to small precipitates (e.g. H – carbide). In high C steels
cementite precipitation takes place first in the austenite, followed by transformation
of the C depleted austenite.
The description of the different bainite morphologies is a rather difficult task as
there is a large variety of terms that are used to describe the different kinds of

181
4 Phase transformations

bainite because many of the morphology characteristics are similar to those of


martensite as well as to those of ferrite. There are various different perspectives on
the bainite morphology, e.g. fine acicular bainite, coarse acicular bainite and
granular bainite. The expressions “fine acicular,” “coarse acicular,” and “granular”
signify the spacing as well as the arrangement of the carbides within the bainite,
which are important for the mechanical properties. In the case of acicular structure,
the bainitic ferrite is in the form of long, stretched laths, which appear in parallel
groups. If the carbon-rich components and the bainitic ferrite are only partially
distinguishable, then the microstructure is said to be fine; otherwise, the
microstructure is called coarse. If no continuous needles appear in groups, then the
microstructure is classified to be granular.
In practice, bainite manifests in two distinctly different microstructures: upper and
lower bainite. In upper bainite, carbon-rich components, such as carbide,
martensite, and/or retained austenite, appear between the laths of the bainitic ferrite.
In lower bainite, carbides precipitate within the ferrite plates. Both types of bainite
can appear in the microstructure concurrently. Differentiation between these types
of microstructures is hardly possible with an optical microscope, since the structure
can be so fine that the carbon-rich components can hardly be distinguished from the
bainitic ferrite. Therefore, Transmission Electron Microscopy (TEM) is usually
applied to differentiate upper bainite and lower bainite. Figure 4.53 shows TEM
micrographs of bainite microstructures in 100Cr6 steel. The lower bainite and
upper bainite in 100Cr6 steel exhibit distinct morphological features. In the lower
bainite, as shown in Figure 4.53 (a), nano-sized carbides precipitate within bainitic
ferrite plates and the particles precipitate with an angle of approximately 60eto the
longitudinal direction of bainitic ferrite. The boundaries of the bainitic ferrite
subunits are indicated in Figure 4.53 (a) in terms of dashed green lines. In upper
bainite, the bainitic ferrite adopts a lath-type shape, unlike the plate-like
morphology of bainitic ferrite in lower bainite. As is shown in Figure 4.53 (b),
nano-sized carbides are precipitated between the bainitic ferrite laths in the upper
bainite. In upper bainite, several carbides precipitate from the grain boundary (GB)
shown at the right bottom of Figure 4.53 (b). The carbides in the upper bainite are
much longer and thicker than the carbides in the lower bainite. The TEM images
reveal that the average size of nano-sized carbides in lower bainite is about 30nm in
width and about 80nm in length, while in upper bainite, the coarsened carbides are
even as large as 100nm in width and 0.5 - 1μm in length.

182
4.3 Bainitic Transformation

Figure 4.53: TEM micrographs of 100Cr6 steel (a) lower bainite, (b) upper bainite;
αB: bainitic ferrite, θ: cementite and GB: grain boundary.

Figure 4.54: Comparison of the simulated nano-sized carbide precipitation within


lower bainite microstructure at 260 °C in 100Cr6 steel using multi-
phase field approach with the experimental observation by TEM. (a)
multi-phase field simulation (b) TEM bright field micrograph. The
colour bar in (a) ranges from 0 wt.%. to 7 wt.%.

The phase transformation kinetics and the microstructure evolution during bainite
formation can be simulated by means of multi-phase field simulation approach.
Figure 4.54 displays the simulated and the experimental-observed carbide
precipitation within the lower bainite microstructure at 260 °C in 100Cr6 steel. The

183
4 Phase transformations

nano-sized carbide precipitation within the lower bainite microstructure tends to


adopt a single crystallographic variant in a bainitic ferrite plate and this is different
from the carbide precipitation within the tempered martensitic microstructures,
where multiple crystallographic variants are preferred.

4.3.2 Formation of the bainitic microstructure


Bainite can be classified according to the metallurgical processes in carbon-poor
bainite (B1), upper bainite (B2), and lower bainite (B3) (Figure 4.55). Figure 4.56
shows a carbon-poor bainitic microstructure, known as “acicular ferrite”. This type
of microstructure consists of ferrite laths and islands of carbon-enriched austenite,
which either completely or partially transforms into martensite. During continuous
cooling or isothermal transformation, mixtures of upper and lower bainite may be
obtained.
Temperature

B1 B1 : Low carbon bainite = ferrite laths with


B2
carbon enriched retained austenite
or martensite fields
B2 : Upper bainite = ferrite-cementite
Ms two-phase structure
B3
B3 : Lower bainite = ferrite plates with finely
dispersed cementite precipitates

Carbon content
Figure 4.55: Schematic classification of the formation ranges of various bainite
microstructures.

100 Pm

Figure 4.56: Carbon-poor bainitic microstructure (“acicular ferrite”).

184
4.3 Bainitic Transformation

Upper Bainite
Upper bainite forms in the upper temperature range of the bainite field, between 550
and 400°C. The progression of transformation can vary, depending on the carbon
content of the steel. In steels with high carbon content, the transformation begins
with the diffusion-controlled precipitation of cementite, Fe3C. As a result of this,
the austenite is depleted of carbon, so that the M s-temperature is reached and
transformation occurs through shearing, according to the martensite mechanism.
In steels containing lower levels of carbon, the driving force for the formation of
cementite is smaller. The austenite, from which the ferrite laths form, does not
“wait” for carbon to diffuse out, but rather transforms according to the martensite
mechanism. After this first step, a second, diffusion-controlled sub-step of the
transformation occurs. Due to the low carbon solubility and the still existing ability
of carbon atoms to diffuse, cementite precipitates out between the ferrite plates.

Lower Bainite
Lower bainite forms in the lower temperature range of the bainite field between 400
and 250°C. Ferrite forms in the same way as in the case of upper bainite. Because
of the decreased transformation temperature, however, carbon diffusion is strongly
restricted, so that the carbon that is insoluble in ferrite cannot diffuse out of the
ferrite plates. As a result, in lower bainite, the diffusion-controlled sub-step of the
transformation reaction consists of a precipitation of carbide particles within the
growing ferrite plates. In doing so, carbides preferably assume an angle of
approximately 60° from the ferrite axis. This angle is a result of the preferred
nucleation on the intersection between the (101)-shear planes of ferrite with the
bainite/austenite phase boundary.
In lower bainite, instead of cementite, the more easily nucleated H carbide may
precipitate (Figure 4.57), or H carbide precipitation precedes the formation of Fe 3C.
The Atom Probe Tomography (APT) images in Figure 4.57 shows the 3D carbon
atomic map and 1D concentration profiles of lower bainite in 100Cr6 steel. It
provides a local overview of the carbon distribution in bainitic ferrite matrix and
carbides. The carbon-depleted region with a carbon content of <1.2 at.% represents
the bainitic ferrite, which is supersaturated with carbon. The carbon-rich regions
correspond to cementite Fe3C and ε carbide Fe2.4C. The particle with a carbon
content of approximately 25 at.% is identified as cementite and the particle with a
carbon content of approximately 29.4 at.% as ε carbide Fe2.4C, both owing to their
respective stoichiometries. The atom probe indicates the co-existence of H carbide
and Fe3C precipitation. The features of substitutional element distribution across
the bainitic ferrite αB/cementite interface and the bainitic ferrite αB/ε carbide Fe2.4C
interface hence infer that both, cementite and ε carbide Fe2.4C precipitation in lower
bainite occur under paraequilibrium mode. The paraequilibrium mode here refers to

185
4 Phase transformations

the phase transformation condition where only the interstitial elements diffuse, the
substitutional elements are frozen in the sublattice.

Figure 4.57: 3D carbon atomic map and 1D concentration profiles of lower bainite
in 100Cr6 steel. αB: bainitic ferrite, θ: cementite and ε: ε carbide
Fe2.4C; ROI: region of interest.
After long holding period, H carbides transform into the equilibrium phase Fe 3C. In
steels, the obvious reaction in an iron matrix is the transition between ε carbide/iron
and cementite (θ),
H-Fe2.4C + 0.6 Fe ⇄T-Fe3C
where Fe is either bcc iron in a bainitic-ferritic matrix at low temperatures or fcc
iron in austenite at higher temperatures. Figure 4.58 shows the Gibbs free reaction
energies between ε Fe2.4C and cementite as a function of temperature in a ferritic
and an austenitic matrix. Positive value of the Gibbs free energy indicates an ε
favored region and a negative value indicates a cementite favored regime. In lower
bainite, where the matrix is mainly bainitic ferrite, the formation of TFe3C and
HFe2.4C has nearly the same probability from a thermodynamic standpoint. In upper
bainite, where the matrix is austenite, however, the formation of cementite is clearly
preferred at any temperature. The theoretical calculations reveal that the formation

186
4.3 Bainitic Transformation

of HFe2.4C benefits from a ferritic matrix and thus ε carbide is more prone to
precipitate from lower bainite than from upper bainite.

Figure 4.58: Gibbs free reaction energies between ε Fe2.4C and cementite (Fe3C, θ)
as a function of temperature in a ferritic (left) and an austenitic matrix
(right). ITT: isothermal transformation temperatures.
A new classification scheme - “Bainite chart”
The bainite exhibits complex microstructures, containing bainitic ferrite matrix and
2nd phases, i.e. alloyed carbides, cementite, retained austenite, etc. At different
bainite formation temperatures, the 2 nd phases precipitate from the bainitic ferrite
matrix at different locations with the varied shapes, which leads to fairly different
mechanical properties of the steels. Based on the micro- and atomic morphological
features of bainite, a new classification scheme is proposed to achieve a possibility
for a full description of bainite microstructures. Following the new classification
system, so-called “bainite chart”, the bainitic microstructures can be classified not
only at micro-scale for industrial applications but also at nano-scale for nano
precipitation mechanism study.
According to the “bainite chart” classification, bainite microstructure consists of
two components: basic structure and substructures. The basic structure is mainly
bainitic ferrite (bcc crystal structure) with different forms, i.e. polygonal, granular,
Widmanstätten, etc. The substructures are the 2nd phases which precipitate at
different locations with varied shapes and types, i.e. alloyed carbides, θ-carbide
(cementite, Fe3C), ε-carbide (Fe2.4C), etc. Figure 4.59 shows a full description of
the two components involved in the bainitic microstructures. The application of the

187
4 Phase transformations

“bainite chart” includes two steps: 1) to identify the crystal structure and form of the
basic structure (bainitic ferrite); 2) to identify the location, type and form of the sub
structure (2nd phases). Two application examples are shown in Figure 4.60 and
Figure 4.61 to exhibit the new classification regimes of the bainitic microstructures
that are traditionally classified as upper bainite and lower bainite respectively.

Figure 4.59: A new classification scheme of bainitic microstructures - “bainite


chart” at IEHK.
Figure 4.60 indicates the classification of the bainitic microstructure formed at 500
°C in 100Cr6 steel by using “bainite chart”. The bainite morphology in Figure 4.60
(a) is classified as “B-L, S-I/θ-E” type and “B-L, S-I/θ-L” type. As is shown in the
TEM bright field image in Figure 4.60 (a), the Basic structure of bainite adopts
Lath-like shape (B-L) and the Substructure precipitates Intragranularly (S-I)
between the bainitic ferrite laths. The 2nd phase substructure is identified to be θ-
carbide using selected area diffraction technique and it adopts Elongated shape (θ-
E) or Lath-like shape (θ-L). The TEM bright field image in Figure 4.60 (b) shows
the lath-like bainitic ferrite basic structure and the θ-carbide precipitation from the
grain boundary with an elongated shape (B-L, S-B/θ-E) or lath-like shape (B-L, S-
B/θ-L). More specifically, the 3-dimensinal atom probe tomography (APT) in
Figure 4.60 (c) reveals the elemental distribution features within bainitic ferrite and
θ-carbide at 500 °C. In addition to the microstructural features observed by TEM,
the carbon atom map further shows the carbon clusters in bainite (C-C).

188
4.3 Bainitic Transformation

Figure 4.60: Bainitic microstructure classification by means of “bainite chart”:


Example1. Bainitic microstructure at 500 °C in 100Cr6 steel; (a)TEM;
(b)TEM; (c)APT
Figure 4.61 shows the classification of the bainitic microstructure formed at 260 °C
in 100Cr6 steel by using “bainite chart”. As is shown in the TEM bright field image
in Figure 4.61 (a), the Basic structure of bainite adopts PLate-like shape (B-PL) and
the Substructure precipitates Intragranularly (S-I) or from the bainitic ferrite platelet
boundaries (S-B). The 2nd phase substructures in the bainitic microstructure formed
at lower transformation temperatures exhibit in large varieties and forms. As is
shown in Figure 4.61 (a) and Figure 4.61 (b), the substructures precipitated within
the bainitic structure at 260 °C includes Cr carbide with a round shape (AC-R), θ-
carbide with a plate shape (θ-P*), martensite with a plate shape (M-P), retained
austenite with a film shape (RA-F). Furthermore, the 3-dimensinal atom map in
Figure 4.61 (c) reveals the θ-carbide and ε-carbide precipitations as 2nd phases with
a plate like shape. Similar with the bainite formation at 500 °C, the carbon clusters
within bainite (C-C) is also indicated.
As is indicated in Figure 4.61, the TEM suggests that the nano-sized carbides in
bainite exhibit elongated shapes with an average length of about 80 nm and an
average width of about 30 nm. However, the TEM observations only show these
nano-sized carbides in the form of two-dimensional projections. With the 3-

189
4 Phase transformations

dimensional morphological information provided by atom probe tomography (APT),


it is revealed that the nano-sized carbides in bainite are with a thickness of 3-10 nm.
Thus, the θ-carbide and ε-carbide precipitation within the bainite microstructure at
260 °C in 100Cr6 steel are classified as plate-like shapes (θ-P* and ε- P*) in Figure
4.61.

Figure 4.61: Bainitic microstructure classification by means of “bainite chart”:


Example 2. Bainitic microstructure at 260 °C in 100Cr6 steel;
(a)TEM; (b)TEM; (c)APT

4.3.3 Temperature range of bainite formation


Bainite is also described as a transitional microstructure because its range of
formation exists between pearlite and martensite. Therefore, a bainitic
microstructure shows properties of both neighbors. Bainitic transformations can
occur continuously or isothermally.
Figure 4.62 shows schematically the dependence of the martensite or bainite
formation from super-cooled austenite on the carbon content. Bainite forms at
higher temperatures than martensite because the required supercooling below
equilibrium is smaller for bainite than for the formation of martensite.

190
4.3 Bainitic Transformation

With regard to bainite formation, upper and lower bainite can be distinguished. As
described in Figure 4.63, the type of formation mechanism is determined by the
temperature and carbon content. In steels with low carbon content, of about 0.1%,
the transition from upper to lower bainite takes place at about 450°C. As the carbon
content increases, the temperature boundary rises, to around 550°C for
approximately 0.5% carbon, subsequently falling to around 350°C. This drop in
temperature begins at the point B, which is the intersection between the boundary
line and the extrapolated Accm line.

Figure 4.62: Diagram of martensite and bainite formation temperatures as a


function of carbon content; Bs = Bainite Start Temperature, Bf =
Bainite finish temperature, Ms = Martensite start temperature , Mf =
Martensite finish temperature.
The bainite transformation temperature range can be indicated in the Fe—C phase
diagram. There exists a boundary line between upper and lower bainite, which
results from the fact that with decreasing temperature, the diffusion of carbon
becomes so low that carbon can no longer diffuse for a distance larger than the
ferrite needles dimension. Thus, a transition takes place from upper to lower
bainite, with its differing formation mechanism.

191
4 Phase transformations

Extrapolated Accm-line
Temperature in °C

723°C
723

BB
ĉ 1
B 550°C
500 B
Bċ3
BBĊ2 350°C
300
0.5 0.8

Carbon content in mass-%


Figure 4.63: Influence of carbon content and temperature on the formation of
different types of bainite.
Depending on the temperature range and amount of carbon, the sequence of
transformation and partitioning/precipitation changes (Figure 4.63). In the field Bĉ
the transformation by shear is possible without pre-segregation of C, although, the
process usually overlaps with the diffusion-controlled transformation of pearlite.
With increasing C content pre-segregation becomes more likely and the boundary
line increases. In the field Bċcarbide precipitation begins first, since the carbon
concentration lies to the right of the Ac cm-line in the two-phase field of the iron-
carbon diagram. As a result, austenite becomes depleted of carbon and bainite
transformation begins when reaching the Bs temperature. This region of upper
bainite formation is limited by the 350°C line, as below this temperature C diffusion
is so hindered that carbide precipitation no longer occurs in austenite. Above B
corresponding to 550°C, no bainite transformation begins, since the carbon
concentration of austenite does not fall below the B S line. In the field BĊ bainitic
ferrite is supersaturated with carbon, and carbides form within the acicular ferrite.
In this case, the nucleation rate is decisive for the first precipitation phase.
From the plot of growth rate over formation temperature for iron-carbon alloys, it
can be seen that the growth curves have similar slopes, while the growth rate
decreases with increasing carbon content, due to a decreasing formation temperature
(Figure 4.64).
The influence of alloying elements, such as chromium and nickel, on bainite growth
rate is principally the same as the influence of carbon. A reduction in growth rate
due to alloying elements is probably because of a restriction in the mobility of the
phase boundary as a result of the accumulation of the alloying elements there.

192
4.3 Bainitic Transformation

Temperature in °C

300 250 200 150

-2
10
Growth rate in mm/minute
0.96% C

-3
10

1.16% C
-4
10

1.43% C

1.80 1.90 2.00 2.10 2.20 2.30


3 -1
Reciprocal temperature 1/T*10 in K

Figure 4.64: Bainite growth rate for Fe-C alloys.


4.3.4 Influence of alloying elements on bainite formation
Carbon has a large influence on the temperature range in which upper and lower
bainite forms. The Bs-temperature is shifted to lower temperatures and longer
holding times by the addition of alloying elements. The influence of alloying
elements on the Bs-temperature can be described by the following empirical
formula:
BS(°C)=830 - 270 (%C) - 90 (%Mn) - 37 (%Ni) - 70 (%Cr) - 83 (%Mo) (4.17)
The solubility of carbon in austenite is significantly greater than in ferrite, which
leads to a delay in the reaction kinetics. As the alloying content increases, a zone
without transformation can develop, in which nucleation is strongly restricted or
even impossible. Such a temperature range can be explained by the existence of
cluster-like segregations on the metastable austenite/bainite phase boundary that are
rich in alloying elements, such as Cr or Mo, as well as carbon. Within a critical
temperature range these segregations can block the growth of bainitic ferrite.
Experiments on Fe-C-X alloys (X: Mn, Mo, Cr, Ni, Cu, Si) show that materials with
critical content levels of Cr, Mo, and Mn lead to a temporarily strong hindrance for
diffusion, which prevents the completion of bainite transformation, although,
dilatometer curves show an apparent final transformation (Figure 4.65).

193
4 Phase transformations

Figure 4.65: Schematic of an incomplete bainite transformation in an isothermal


TTT diagram.
The addition of alloying elements usually results in a delay of the ferritic and
pearlitic reactions. Also, the bainitic reaction is shifted to lower temperatures,
which leads to a zone without transformation between the individual phase fields.
Independent of the influence of alloying elements, it is difficult to obtain a fully
transformed bainitic microstructure due to the neighboring martensitic
transformation. Materials with bainitic structures offer a variety of advantages,
such as a low carbon content, which improves weldability and other technological
properties. These properties can be varied over a wide spectrum. For doing so,
transformation diagrams are convenient that have a metastable austenite field
between the ferrite and bainite phase fields shown in isothermal TTT (time-
temperature-transformation) diagrams. This is best shown in isothermal TTT
diagrams with a metastable austenite field between the ferrite and bainite phase
fields, as shown in Figure 4.66. The ferrite phase field is delayed to a later time,
and the bainite transformation begins immediately upon reaching the B S
temperature. This displacement and separation of the phase fields is achieved for
example, by adding boron. The addition of only about 0.003% boron, delays the
ferrite and pearlite transformations, while the bainitic transformation is not
influenced. Precondition for this strong effect is that B exists in solid solution and is
not precipitated in form of BN.

194
4.3 Bainitic Transformation

Figure 4.66: Schematic of the influence of boron on transformation behavior.

Mechanical properties of bainitic steels


The mechanical properties of bainite are determined mainly by the size of the
bainite laths and to a lesser degree, by the size of the bainite bundles. The
transformation temperature is decisive for the lath width of bainite (Figure 4.67).

Figure 4.67: Influence of bainite formation temperature on bainite laths width.


The size of the bainite laths decrease with a decreasing transformation temperature
and is independent of the initial austenite grain size. The properties of toughness
are significantly influenced by the form of the bainite formation. Figure 4.68
shows the notch impact temperature curves for steel, for which intentionally close
cooling rates were chosen in order to obtain a finer bainitic structure. As the

195
4 Phase transformations

bainitic microstructure becomes coarser, the maximum of the notch-impact


temperature curve is shifted to a higher temperature.
160
A - Martensite t8/5 = 10s D
Notch-impact energy in J

140
B - Fine acicular bainite t8/5 = 30s
120 C - Coarse acicular bainite t8/5 = 100s
C
D - Granular bainite t8/5 = 300s
100

80
B
60
A
40

20

0
-120 -80 -40 0 40 80 120 160
7HVWLQJWHPSHUDWXUHLQr&
Testing temperature in °C
Figure 4.68: Influence of various microstructures on the notch impact temperature
curves of the normal annealed steel S690N.
Figure 4.69 shows the influence of the lath size of a bainitic microstructure on the
mechanical properties. The test samples were quenched in water immediately after
either various austenitisings or a 60% deformation of the austenite. In all cases, the
BS-temperature was 450°C. With an initial austenite grain size of 4, the yield
strength reaches 800 MPa, and with an initial austenite grain size of 8, the yield
strength even reaches 920 MPa.
The strength properties of bainite are closely connected to the transformation
temperature (Figure 4.70). Generally, with regard to the relationship between
structure and strength, it can be said that the strengthening mechanisms of grain
refinement are responsible for the properties of upper bainite (lath bainite).
Strength and toughness increase as the temperature of formation decreases, since
the carbides become smaller and more evenly distributed, and the ferrite sub-
structure becomes finer.

196
4.3 Bainitic Transformation

0
Austenite grain size number

Transition temperature in °C
4
-50

6
-100

-150
8
-200
800 900
Yield strength Rp0.2 in MPa

Figure 4.69: Influence of austenite grain size on mechanical properties in bainitic


microstructures.

Figure 4.70: Relationship between transformation temperature, microstructure, and


strength of low-carbon steels, according to Pickering.
The graph of the tensile strength of bainitic steels with low carbon contents shows
that the range of strength can vary greatly within the bainite field, as a result of the
changing share of the diffusion-controlled and martensitic mechanisms on the entire
transformation. During the production of structural components, the sole
transformation in the bainite field, “austempering”, is only used in special cases. As
a rule, a combination of martensitic quench-hardening and subsequent tempering
results in a higher toughness at the same strength (reduction in area, temperature-
dependent notch-impact energy), although, the danger of forming hardening cracks
increases.
The (isothermal) bainite transformation is preferred over the classic quenching and
tempering, when, in order to achieve a certain strength-toughness ratio, a material

197
4 Phase transformations

must be tempered in a temperature range where tempering brittleness is to be


expected. Moreover, it is used when hardening cracks are expected during
martensitic hardening.
Transformation in the lower bainite range, just above the martensite field, is useful
for sensitive steels, and especially for bulky components, when a temperature
balance over the diameter of the material can be achieved. The temperature balance
serves to reduce thermal stress in metastable austenite so that only transformational,
and no thermally-determined, stresses appear during the subsequent bainite
transformation. Since the transformation temperature is far above room
temperature, stresses can be better removed from the material than at room
temperature or just above it, as is the case during martensitic hardening.
Transformation in the bainite field (during continuous cooling) has gained
significance in relation to the thermomechanical treatment of low-alloyed
construction steels.
To increase strength in steels with greatly reduced carbon content (approximately
0.05% C), increasing the dislocation density in austenite can be used as another
mechanism besides the strengthening mechanisms of grain refinement, solid
solution, and dispersion hardening through lower end-rolling temperatures.
4.3.5 Widmanstätten ferrite
Widmannstätten ferrite is named after the microstructural form discovered in
metallic masses within meteorites by A. v. Widmannstätten (Figure 4.71).
Widmannstätten ferrite forms preferably at carbon contents between 0.2 and 0.4%,
and with relatively quick cooling from high austenitising temperatures, i.e. with
large austenite grains. An important characteristic of Widmannstätten ferrite is its
ferritic plates, which form at the austenite grain boundaries and grow into the
austenitic grain. These ferrite plates appear as needles in cross-sections. These
“needles” demonstrate an orientated relationship with the austenitic lattice
according to Kurdjumov-Sachs and an associated habit plane. No substructure is
present within the individual crystals, due to the absence or minimal presence of a
dislocation substructure. Due to the conditions for its creation, Widmannstätten
ferrite can be found in casting microstructures and welds. Compared to a globulitic
microstructure, the toughness properties of this microstructure are poorer.

198
4.3 Bainitic Transformation

Figure 4.71: Widmannstätten ferrite.


4.3.6 Acicular ferrite
Acicular ferrite is needle-shaped or rod-like, as shown in Figure 4.72. It has a very
high aspect ratio and appears in the microstructure as fully enclosed ferritic plates
without any intra-granular carbides precipitation. It normally does not grow in
sheaves but as single needles growing from the prior austenite grain boundaries or
starting at existing precipitations or inclusions.

Figure 4.72: Schematic sketch of the two different types of acicular ferrite and the
growth steps of acicular ferrite.
4.3.7 Application of bainitic steels

Railway steels
The Ofot railway in northern Norway (Figure 4.73) is used for the transportation of
ore. This 40 km long route is among the most difficult railway routes in the world,
since 30% of it travels through a mountainous region along fjords, with steep

199
4 Phase transformations

gradients and sharp curves. The total load on these tracks is between 15 and 22
million tons per year. The relatively steep gradient leads to very high braking
forces, which must be absorbed by the tracks. This, in turn, leads to extremely high
wear on the tracks. Until 1967 a pearlitic steel of grade 900 was used for the tracks,
but they had to be exchanged annually. Since 1978 a fine pearlitic Cr-V steel of
grade 1200 has been employed, with a useful life of up to 5 years. Bainitic steels
(grade 1400) have a lifespan of 8 years (Figure 4.74).

Figure 4.73: Ore train on the Ofot railway.

Figure 4.74: Wear of various railway steels in test curves with a radius of 300 m;
the tensile strengths of the steels are given.

200
4.3 Bainitic Transformation

Steels for construction of commercial vehicles


The construction of modern commercial vehicles is an important field of application
for the fine-grained steels in use today. Due to their excellent properties for
manufacturing and use, these steels, next to their good processability, show a
favorable ratio of load to dead weight. The steel grades used must often contain
different chemical compositions and be produced using optimised processes so that
the best possible cost/performance ratio for each specific application is guaranteed.
The main frame in freight trucks is a characteristic example for a component that is
produced by rolling or extrusion and finished from a 7 to 9 mm hot strip. These
parts must combine good deformability with high strength, especially for cyclic
wear. They are usually made from thermomechanically rolled steels with a
minimum yield strength of 500 MPa.
Water-quenched and tempered fine-grained steels meet the highest requirements for
strength and toughness values. They are rolled as heavy plates and subsequently
undergo a heat treatment above A3, and are subsequently quenched in water. The
fine-grained construction steels that are used are adjusted with regard to their
chemical composition, forming behavior, and heat treatment such that a minimum
yield strength is maintained in the range between 450 and 960 MPa. The group of
steels with strengths from 690 to 960 MPa has a large field of application for
commercial vehicles. The high strength allows for a reduced wall thickness and
savings in material and finishing costs. These steels are typically used for street
vehicles that transport heavy loads, railcars carrying heavy loads, crane
components, and earth movers.
After transformation, the microstructure of water-quenched and tempered steels
consists of martensite and fine bainite. This microstructure demonstrates a good
combination of strength, toughness, and deformability. Figure 4.75 shows that
bainitic hot strips, compared to ferritic-pearlitic strips, allow for minimum yield
strength of 800 MPa to be set.

201
4 Phase transformations

%C % Si % Mn % Al
Analysis
0.06 0.3 0.9 0.04
Yield strength Re and tensile strength Rm in MPa

900
Rm Bainitic steels
850
Re (0.18 % Ti, B - alloyed)
800

750 Ferritic-pearlitic
steels
Rm
700 (0.14 - 0.18 % Ti)

650
Re
600

550

500 Start of
the bainite
450 formation (Bs)

400
580 600 620 640 660 680 700 720 740

Coiling temperature in °C
Figure 4.75: Effect of ferritic-pearlitic or bainitic microstructures on strength
properties of hot strips; the bainitic microstructures are represented by
means of boron alloying and low hot strip coiling temperature.

202
4.4 Precipitating from a Supersaturated Solid Solution

4.4 Precipitating from a Supersaturated Solid Solution


In solid solutions, the solubility of interstitial or substitutional dissolved atoms in
the matrix is often a function of temperature. Usually, the solubility of these atoms
decreases with decreasing temperature. At near-equilibrium, meaning infinitely
slow cooling, microstructures can be set which correspond to the phases of the
phase diagram. However, in reality, industrially relevant cooling procedures do not
proceed near-equilibrium. The achievement of a phase-equilibrium is suppressed,
and a solid solution, supersaturated with interstitial and substitutional dissolved
atoms, exists at room temperature. In order to remove this supersaturation, the
forcibly dissolved atoms precipitate from the matrix. The supersaturation of
interstitial dissolved atoms can be reduced through precipitation. There are
adequate possibilities that diffusion can occur even at room temperature.
Precipitation takes place in three steps: nucleation, nucleus growth and coagulation.
The diffusion rate of substitutional elements at room temperature is generally so low
that precipitation does not occur.
Typical interstitial dissolved atoms in D-iron are carbon, nitrogen, oxygen and
hydrogen. With regard to precipitation processes, hydrogen and oxygen are
irrelevant. There is no supersaturation of hydrogen due to its high diffusability, and
oxygen is removed from steel through deoxidation processes during secondary
metallurgy.
Since the advancement of secondary metallurgy and the introduction of aluminum
as a deoxidation agent, nitrogen is not really significant in the processes of
precipitation from supersaturated solid solutions, because free nitrogen is
completely bound. Only in a few exceptional cases is free nitrogen available.
Therefore, the main focus in the following is on the precipitation of interstitial
dissolved carbon out of the supersaturated D-matrix. 
A pure iron-carbon alloy with 0.01% C is annealed at 720°C in the D-field until it is
completely homogenized and subsequently quenched in cold water. Two flat bar
tensile test specimens are produced from the quenched material. The mechanical
properties of sample A are immediately tested with the help of an uniaxial tensile
test. Sample B is held for 10 minutes at 60°C before testing. Figure 4.76 shows
the resulting stress-strain curves for both samples. Sample B shows, in contrast to
sample A, a distinct yield strength after 10 minutes of artificial ageing. This yield
strength is a phenomenon of ageing, which is discussed in detail in the following
sections.
During industrially relevant processes, cooling takes place so quickly that the
achievement of equilibrium conditions is suppressed, and carbon is frozen in a
supersaturated solid solution. The supersaturation can be removed through
accumulation of the forcibly dissolved atoms in lattice defects, or through
precipitation in the form of carbides. The associated time-dependent change in the

203
4 Phase transformations

material’s properties often results from this carbide precipitation and is termed
ageing.
The development of a distinct yield strength can be observed during ageing, along
with an increase in strength and a simultaneous decrease in ductility. Dislocation
movement is blocked by an accumulation of carbon atoms and micro-precipitations
in the dilated tension fields of the dislocations. This accumulation is called a
Cottrell cloud. With advanced ageing, carbide precipitation can occur after an
incubation period. After strength is increased through precipitation hardening, a
decrease in hardness can occur. This is due to the coagulation of precipitations,
which is often referred to as overageing.

Figure 4.76: Stress-strain curves for a steel with 0.01 mass-% C immediately after
quenching from the solution annealing temperature (A) and after 10
minutes of ageing (B).
4.4.1 Theoretical basics
As the temperature decreases, the solubility of carbon and nitrogen atoms in the
D-iron lattice decreases as well. The equilibrium solubility of these atoms in D-iron
(diluted solution) can be generally described with the following equation:
Q
cC / N (T ) B ˜ exp( ) (4.18)
RT
where cC/N = equilibrium concentration of C or N in D-iron in mass-%, B = constant,
Q = activation energy needed to bring 1 mol of C or N atoms into the D-matrix in J
mol-1, R = general gas constant (R = 8.3143 J mol-1 K-1), and T = temperature in K.

204
4.4 Precipitating from a Supersaturated Solid Solution

The values B = 2.55 and Q = 40614 J˜mol-1 are given for the equilibrium solubility
of carbon in an D-solid solution with metastable cementite Fe3C (in mass-%).
Equation 4.19 results when these values are applied. When applied to
thermodynamically stable graphite, in other words, pure carbon, the solubility of C
is significantly smaller:
 4850
cc (T ) 2.55 ˜ exp( ) (4.19)
T (K )
where cc(T) = equilibrium concentration of C in D-iron in mass-%, and T =
temperature in K.
The ratio of carbon atoms to iron atoms at room temperature is approximately 1 to
108, and at 723°C approximately 1 to 10³. The solubility for carbon and nitrogen is
graphed logarithmically, linearly and in accordance with Avrami, in Figure 4.77,
Figure 4.78, and Figure 4.79.

Figure 4.77: Solubility of carbon in D-iron in equilibrium with Fe3C as a function


of temperature, graphed on a half-logarithmic scale.

205
4 Phase transformations

Figure 4.78: Solubility of carbon in D-iron in equilibrium with Fe3C as a function


of temperature, graphed linearly.

Figure 4.79: Solubility of C in D-iron in equilibrium with Fe3C, graphed in


accordance with Avrami (log C a –1/T).
The solubility of carbon in D-iron is not only determined by the temperature and the
phase present at equilibrium. The presence of fine grains can raise the solubility,
Figure 4.80. For a grain size larger than 100 nm, a dissolved carbon content of
5 ppm requires a temperature of 300°C. As the grain size decreases, so does the
required temperature. At a grain size of 20 nm, approximately 5 ppm carbon is
already in solution at a temperature of around 80°C.

206
4.4 Precipitating from a Supersaturated Solid Solution

Figure 4.80: Influence of small grains on the solubility of carbon in the D-matrix.
According to the phase diagram, the microstructure of a sample with 0.01% C
consists of ferrite and cementite at room temperature (Figure 4.81). If this sample
is annealed for a sufficiently long period of time at T 1 = 720°C, then the carbides
dissolve and carbon goes completely into solution. A pure ferritic microstructure
(D-solid solution) is then present. After near-equilibrium cooling, the first
cementite precipitates form at T2 = 600°C. The concentration of carbon dissolved
in the D-iron would then proceed along the solubility line with continuing cementite
precipitation, until the carbon content cRT is reached at room temperature. In
industrial cooling processes, however, carbon cannot precipitate out quickly
enough. After rapid cooling to room temperature, a large amount of forcibly
dissolved carbon is present in the D-solid solution.
Since diffusion is possible for the interstitial carbon atoms at room temperature, the
supersaturation can be removed, and the microstructure eases towards a state of
equilibrium. This is the result of carbon precipitating out of ferrite. This type of
precipitation is dependent on the ageing temperature. At first, the carbon atoms
accumulate in the dilated tension fields of the dislocations, in which the expanded
lattices serve as nucleation sites. The carbon atoms arrange themselves in Cottrell
clouds, which evolve an enrichment and micro-precipitation in the nano-range. In
these expanded lattices, the distortion energy needed for inclusions is significantly
smaller than in undistorted lattices. The number of potential nucleation sites
increases with the dislocation density.

207
4 Phase transformations

Figure 4.81: Segment of the metastable Fe-Fe3C phase diagram.


Carbide precipitation also occurs simultaneously. At room temperature, low
temperature carbides precipitate out (possibly K-carbides). An ageing temperature
of around 200°C leads to the precipitation of H-carbides (Fe2.4C), and temperatures
of 300 to 400°C allow for cementite precipitation (Fe3C). Some carbides, which are
thermodynamicly stable in Fe-C alloys, are summarised in Table 4.7.
C-content in Crystal
Name Symbol Composition
mass-% structure
Cementite T (theta) Fe3C 6.69 orthorhombic

Hägg-carbide F (chi) Fe5C2 7.92 monoclinic

H-carbide H (epsilon) Fe2.4C 8.22 hexagonal

Eckstrom-Andock-
- Fe7C3 8.44 hexagonal
carbide

K-carbide K (eta) Fe2C 9.77 orthorhombic

Table 4.7: Overview of various iron carbides.

208
4.4 Precipitating from a Supersaturated Solid Solution

Precipitation occurs through nucleation (increasing number of grains), nucleus


growth (constant number of grains), and coarsening (decreasing number of grains).
The course of precipitation is shown in Figure 4.82. If the nucleation phase is
completed, and the critical nucleus radius has been reached after the incubation time
t0, then nuclei grow as a result of carbon diffusion from the supersaturated matrix to
the site of precipitation. The matrix quickly becomes depleted of supersaturated
carbon, and the size of the precipitated phase increases. After this steep rise, the
curve plateaus and runs asymptotically towards one. The precipitation process is
slowed by an increase in the length of the diffusion path, and a decrease in the
precipitation driving force, caused by the removal of the supersaturation. Finally,
no further carbon atoms precipitate out. The large precipitates present become
coarser at the expense of smaller precipitates. The driving force for coagulation is
the attempt to minimise energy-intensive phase boundary surfaces. On the other
hand, the elastic strain around pearlite becomes larger during pearlite growth, which
provides a resisting force against coagulation.
1
Precipitated amount W(t)

0
t0 t Ageing time log t

Figure 4.82: Schematic diagram of the course of precipitation.

Nucleation
The progression of precipitation depends strongly on the conditions for nucleation.
If there are many nucleation sites, such as displacements, grain boundaries, and
other lattice defects, then the process of precipitation is strongly accelerated.
Supercooling and supersaturation also have a strong influence on nucleation. The
exact relationship is explained in the following sections.
Nucleation as a function of supersaturation
The following derivation assumes some large simplifications, but represents the
influence of supersaturation on the rate of nucleation well.

209
4 Phase transformations

The amount of carbon that precipitates out of the supersaturated matrix can be
described with Equations 4.20 and 4.21. With the assumption that a spherical
nucleus precipitate is already present with the radius r0 at time t = 0, then the flow
of atoms through the carbide surface is:
dn 2
j D ˜ 4Sr0 (4.20)
dt
where n = number of precipitated atoms, t = time in s, jD = diffusion flux density in
cm-2s-1, and r0 = nucleus radius at the time t=0 in cm.
On the other hand, the depletion of carbon atoms from the matrix per unit time is
described in Equation 4.21:
dn dc m 4
˜ S x3 (4.21)
dt dt 3
dc m
where n = number of atoms having left the matrix, t = time in s, =
dt
concentration change in the matrix per unit time, and x = half of the average grain
spacing in cm.
Figure 4.83 shows the average distance between two carbides 2x. The average
matrix volume per precipitate is a sphere with radius x.

x x

Figure 4.83: Precipitation nuclei are surrounded by spherical volumes of the matrix
with the radius x.
The atoms that leave the matrix accumulate on the precipitates. Equation 4.22
follows from Equations 4.20 and 4.21:
dc m 4 2
˜ S x3 j D ˜ 4Sr0 (4.22)
dt 3
Equation 4.23 reflects the first law of diffusion (under the assumption that D z D(r),
one-dimensional view):
dc c0  cE
jD (r0 )  D r r0
D (4.23)
dr r0

210
4.4 Precipitating from a Supersaturated Solid Solution

where jD = diffusion flow density in cm-2s-1, D = diffusion coefficient in cm2s-1, c0 =


initial concentration in the matrix on the phase boundary, and cE = equilibrium
concentration in the matrix.
When substituted in Equation 4.22, then Equation 4.24 follows:
dcm 3D
 ˜ (c0  cE ) ˜ r0 (4.24)
dt x³
Under the assumption that the grains grow at a steady-state, in other words, that the
change in concentration in front of the grains is constant, and that D z D(t), the
concentration in the matrix is given by:
3D
cm  ˜ (c0  cE ) ˜ r0 ˜ t  const . (4.25)
x3
If time t = 0, then c = constant, and therefore, c = c0. At time
t = 0, depletion of the matrix has not yet begun. This leads to Equation 4.26:
3D
cm  ˜ (c0  cE ) ˜ r0 ˜ t  c0 (4.26)

Reformulated:
c0  cm
c0  cE
3D ˜ t ˜ r0

3 ˜ r0
x3
˜ Dt
2
(4.27)

From this we can conclude that the supersaturation c0  cm is proportional to:


the reciprocal of the grain spacing to the third power 1/x³,
the average distance of diffusion D ˜ t , and
the grain size r0 .
Nucleation rate as a function of supercooling
The nucleation rate N is strongly dependent on supercooling. This relationship
will be explained more thoroughly in the following. The nucleation rate follows a
typical Boltzmann relationship:
'Go
N ~ exp( ) (4.28)
kT
where N = nucleation rate, 'GO = nucleation work, k = Boltzmann constant, and T
= temperature.
The change in free enthalpy during the precipitation of a spherical nucleus with
radius r is given in Equation 4.29:

211
4 Phase transformations

4
'G(r ) Sr ³ ˜ ('g u  H el )  4Sr ² ˜ V (4.29)
3
where 'G(r) = change in free enthalpy, r = radius of the nucleus, 'gu = change in
the free molar energy of volume, Hel = elastic energy of distortion, and V = specific
interfacial energy.
When a nucleus forms, volume energy is gained and interfacial energy is used. The
difference in the molar volume of the precipitate, as compared to the mother phase
plays an important role during precipitation in the solid-state. The difference in
volume causes elastic distortions, whereby the elastic energy Hel increases with
increasing nucleus volume. The critical radius of nucleation rc is defined by
dG (r ) !
0 , as can be seen in Figure 4.84.
dr
2V
rc (4.30)
'g u  H el
16 V³
Ÿ 'G(rc ) 'G0 S (4.31)
3 ('g u  H el )²
Since elastic distortion is not a function of supercooling, it follows:
1
'G0 ~ (4.32)
('g u )²

Figure 4.84: Free energy of a spherical nucleus as a function of grain radius r and
nucleation conditions.
The change in the free molar energy of volume is given in Equation 4.33:
'g u 'h  T's (4.33)

212
4.4 Precipitating from a Supersaturated Solid Solution

where 'gu = difference in free molar energy, 'h = difference in molar energy, 's =
difference in molar entropy, and T = temperature.
In equilibrium, the molar free energies of both phases are equal:
!
'g u 0 (in equilibrium) (4.34)
As a result, Equation 4.33 becomes:
0 'h  TGG 's (4.35)
œ 'h TGG 's (4.36)
Equation 4.36 substituted into Equation 4.33 results in:
'g u (TGG  T )'s 'T's (4.37)
As a result, the molar free energy 'gu is:
'g u ~ 'T (4.38)
By substituting this new relationship in Equation 4.32, the following results:
1
'G0 ~ (4.39)
('T )²
§ 1 ·
and N ~ exp¨¨  ¸¸ (4.40)
© ( 'T )² ˜ R ˜ T ¹
Thus, small changes in supercooling cause large changes in the rate of nucleation.
The larger the supercooling, the greater the rate of nucleation will be.
The nucleation rates observed in reality, however, are much larger than the rates
calculated from Equation 4.28. This is because nucleation does not occur
homogeneously in free solid solution, but instead heterogeneously in existing lattice
defects, such as grain boundaries, precipitates, or dislocations. Accordingly, the
surface energy that must be supplied is reduced, and the applicable nucleation
energy is significantly lower (Figure 4.84).

Nucleus growth
The precipitation occurring after incubation can be described through the half-
empirical approach of Johnson, Mehl, Avrami, Kolmogorov (JMAK) for
transformation and precipitation processes, Equation 4.41:
­
° §t· ½
n
°
W (t ) 1  exp ® ¨ ¸ ¾ (4.41)
W
°̄ © ¹ °¿
where W(t) = precipitated volume fraction, t = time in s, W time constant in s, and
n =precipitation exponent.

213
4 Phase transformations

The following assumptions are made: at time t = 0 all nuclei already exist; all nuclei
are of the same size and spherical. When precipitation is complete, the number of
nuclei has not changed.
Precipitation kinetics strongly depends on nucleus density, supersaturation, and
supercooling. If many sites of nucleation are present, then the precipitation
exponent n can be given a value, for example, 2/3. If the number of nucleation sites
or the level of supersaturation changes, then the progression of precipitation can be
entirely different, with a value of, for example, n = 3/2. These examples of different
courses of precipitation are shown in Figure 4.85. The time constant W is
determined mainly through the diffusion coefficient, Equation 4.42.
1 3D(c0  cE )
(4.42)
W
1/ 3
cP R 2
where W = time constant, D = diffusion coefficient, c0 = initial concentration in the
matrix, cE = equilibrium concentration in the matrix, c P = equilibrium concentration
within the precipitating carbides, and R = attainable final radius of the grain.
1.0
Precipitated amount W(t)

n = 2/3 n = 3/2
0.8

0.6

0.4

0.2

0.0
0 500 1000 1500 2000 2500
tO Precipitation time in s
Figure 4.85: Progression of precipitation as a linear function of time for various
exponents of the JMAK equation.

Coarsening
As ageing time increases, the grains coarsen, whereby the number of grains
decreases. The driving force of this system is the effort to minimise the boundary
area. The coagulation of grains after the completion of precipitation, meaning the
growth of larger grains through the dissolution of smaller ones, can be described by
the Gibbs-Thomson Equation:
2:V
c(r ) cE exp( ) (4.43)
rRT
After reformulation:

214
4.4 Precipitating from a Supersaturated Solid Solution

2:V c(r )
RT ln (4.44)
r cE
where : = molar volume of the substance, V specific interfacial energy, r = grain
radius, R = general gas constant, T = absolute temperature, c(r) = equilibrium
concentration in front of a grain with radius r, and cE = equilibrium concentration in
front of a grain with a quasi-planar interface, which is the equilibrium concentration
according to the phase diagram.
This equation expresses how the concentration of the precipitated element on the
matrix/precipitate phase boundary is dependent on grain size. The larger a grain is,
the more the matrix is depleted between two spherical precipitates of differing size,
in the course of precipitation. The following stages can be distinguished:
Stage 1 (Figure 4.86): Both grains grow, due to an existing concentration gradient
in front of each grain. The precipitating element diffuses from the matrix to both
grains so that the concentration in the matrix cm(t) decreases, while the area around
the larger grain becomes more depleted than the area around the smaller grain, c(r1)
>c(r2).
Stage 2 (Figure 4.87): In the course of grain growth, the concentration cm(t) falls
below c(r1) of the smaller grain. Due to the present concentration profile, the
smaller grain dissolves while the larger grain continues to grow. The grains
coagulate. Coagulation is eventually stopped by the increasing distortion in the
matrix.

Figure 4.86: Schematic description of the growth of two differently sized grains:
c(r2) < c(r1) < cm(t).

215
4 Phase transformations

Figure 4.87: Schematic description of the growth of a larger grain at the expense of
a smaller one: c(r2) < cm(t) < c(r1).
4.4.2 Variables influencing carbide precipitation
Carbide precipitation is mainly influenced by the degree of supersaturation, the
diffusibility of the supersaturated atoms in the matrix, and thereby, the temperature
and alloying composition of the material.
The influence of supersaturation can be seen in Figure 4.88. According to the left
graph, three samples with different compositions were quenched to room
temperature after annealing at 723°C, and then subsequently aged at 140°C. Here,
the maximum damping value, determined with a Snoek-pendulum, is a measure for
the present amount of dissolved carbon. Therefore, the maximum damping value is
proportional to the dissolved amount of carbon.
As the initial carbon content rises, the precipitation driving force increases as a
result of increasing supersaturation, and with that, the nucleus density also
increases. Shorter diffusion distances ease precipitation, and the damping values
decrease more quickly with increasing supersaturation, which leads to a lower
residual carbon concentration.
Figure 4.89 shows the equations of athermal transformation according to Koistinen-
Marburger and Hsu-Hongbing.

216
4.4 Precipitating from a Supersaturated Solid Solution

Figure 4.88: Influence of the initial carbon content on the progression of


precipitation.

Figure 4.89: Equations (Calculations) of athermal transformations.

217
4 Phase transformations

Figure 4.90 shows three similar samples, which were quenched after annealing at
720°C to different ageing temperatures. As the ageing temperature increases, the
supersaturation and nucleus density decrease, but the grain growth proceeds more
quickly due to the greatly increased rate of diffusion. The matrix becomes depleted
of carbon more quickly.

Figure 4.90: Influence of ageing temperature and time on carbon precipitation.


Figure 4.91 shows the progression of precipitation as dependent on the alloying
composition at 100°C, plotted on a double-logarithmic scale. If carbon alone is
present, then the addition of manganese has no effect. Curve 1 applies to carbon
precipitation both with and without the addition of manganese. In manganese-free
steels, carbon precipitation is greatly accelerated by an increase in the amount of
nitrogen (Curves 2, 3, and 4). The addition of manganese would strongly delay the
precipitation of nitrogen, though this is not shown here. The precipitation of
nitrogen proceeds more quickly than that of carbon. This is due to the significantly
higher diffusion rate of nitrogen, as compared to carbon. If carbon and nitrogen are
present side by side, then the matrix is more highly supersaturated than is the case
when one element alone is present. Furthermore, nitrides serve as nucleation sites
for carbides and vice versa. From a solution supersaturated with both carbon and
nitrogen, the precipitation of carbon-nitrides is possible.

218
4.4 Precipitating from a Supersaturated Solid Solution

6
99
95 N+C 4
90
80
2 C+N
60 1

40 -ln (1-W(t))
20
-1
10 10
1 0.010%C (+ 0.025%Mn)
5
2 0.012%C + 0.017%N
3 N 1 C 3 0.025%N

-2
4 0.023%N + 0.008%C
1 10
1 2 3
1 10 10 10
Ageing time in min
Figure 4.91: Precipitation as a function of chemical composition, at 100°C.
4.4.3 Ageing
Ageing is the term for a time-dependent change in material properties. This is often
associated with carbide precipitation and a change in mechanical properties. Figure
4.92 shows the hardness of an unalloyed steel in dependence of the ageing
temperature. After annealing at 700°C for 20 minutes, the steel was quenched in
water and aged at temperatures from 35 to 250°C. Generally, the hardness increases
with a temperature of up to 75°C, before a decrease in hardness begins in the case of
long annealing. The lower the temperature is, the later the maximum hardness is
reached, though the increase in hardness is greater. This is a result of the
precipitation of very fine low-temperature carbides, which cause a large increase in
strength due to their small size. In the temperature range of 75 to 250°C, metastable
H-carbides (Fe2.4C) precipitate, which are replaced by cementite precipitates as the
duration of annealing increases. The H-carbides are larger than the low-temperature
carbides and, therefore, do not contribute so much to the increase in hardness.
Above 250°C, cementite (Fe3C) precipitates directly, which due to its size, does not
lead to an increase in strength. Hardness decreases, since the influence of the
strengthening of the solid solution with the precipitation of carbon out of the
supersaturated matrix decreases.
Figure 4.93 is an isothermal plot of the precipitation sequence for various carbides.
The type of precipitation is dependent on temperature. Low-temperature carbides
form at very low temperatures. As the temperature increases, H-carbides then
precipitate out, and finally, above 250°C cementite precipitates. The low-
temperature carbides first precipitated transform into H-carbides as the ageing time
increases, up to approximately 100°C. Through diffusion, which becomes easier at

219
4 Phase transformations

temperatures above approximately 180°C, the H-carbides are transformed to Fe3C as


the ageing time further increases.
200
Unalloyed steel 35°C
Vickers hardness HV

180
50°C
160

140
75°C

120 100°C

100 250°C 150°C

80
2 3 4 5 6 7
1 10 10 10 10 10 10 10
Time in s
Figure 4.92: Hardness as a function of time, at different ageing temperatures for an
unalloyed steel with 0.046 mass-% C.

Figure 4.93: Isothermal precipitation diagram for an unalloyed steel.


The hardness of an unalloyed steel with a significantly reduced carbon content of
about 0.01% is plotted as a function of ageing temperature in Figure 4.94. At 22°C
the maximum hardness is reached, as previously described, after a long ageing
period, caused by the precipitation of very fine low-temperature carbides. At 240°C
the curve shows a brief increase in hardness, which is probably due to the
precipitation of fine H-carbides. The H-carbides coarsen quickly, however, and
transform into coarse cementite particles. At 170°C two hardness peaks can be

220
4.4 Precipitating from a Supersaturated Solid Solution

distinguished. The first maximum is probably due to the precipitation of H-carbides.


The hardness then decreases due to the coarsening of these particles. The second
maximum can be explained by the precipitation of cementite, which forms finely
enough at this temperature to lead to an increase in hardness. As the cementite
grows, the hardness eventually decreases. At temperatures above 250°C the
strength continuously decreases as a result of the cementite that directly precipitates
at this temperature.
Ageing time in days
1 10 100 500
Vickers hardness HV3

Unalloyed steel
22°C
240°C 170°C

1 101 102 103 104 105 106


Ageing time t in minutes
Figure 4.94: Hardness as a function of ageing temperature and ageing time for an
unalloyed steel.
4.4.4 Application examples

Continuous annealing of steel strips


During the continuous annealing of deep-drawn steels with carbon levels lower than
0.01%, the annealing cycles I and II given in Figure 4.95 can be applied. The
individual microstructures developed in these cases are significantly different from
each other.
The first phase of the temperature-time curve, to which heating, holding at the
annealing temperature and the first phase of cooling belong, is responsible for
recrystallisation and grain growth. The rapid cooling to room temperature and
heating to the overageing temperature (II), or rapid cooling to the overageing
temperature and the subsequent overageing treatment (I) belong to the second phase
of the annealing cycle, which determines the carbide distribution in the
microstructure and the content on supersaturated, dissolved carbon. Above all, this
determines the ageing behavior of light sheets.

221
4 Phase transformations

During both annealing cycles, the cold-rolled steel strip is first heated to 830°C in
order to reduce the strain hardening of the rolled strip within a short time span
through recrystallisation. Recrystallisation leads to a fine-grained ferrite, whose
grain size increases with annealing time. Since the {111}-texture is preferentially
formed during annealing in the two-phase region, and is further supported by grain
growth during the first phase of cooling below A 1, a microstructure is created with
an r-value favorable for deep-drawing properties. In the second phase, the rapid
cooling is responsible for the precipitation and dispersion of cementite and with
that, the supercooling of the supersaturated carbon. As the cooling rate increases,
the precipitation driving force also increases. In turn, this reduces the necessary
overageing time. This is due to the smaller distance between the carbides, which
were formed through overageing. In annealing cycle II, carbon is frozen in a highly
supersaturated solution through strong supercooling. During subsequent ageing, the
carbon precipitates out even within the grains in a finely dispersed fashion.
Annealing cycle I demonstrates a smaller cooling rate. Even during cooling,
diffusion is already sufficiently possible for the precipitation of cementite on the
grain boundaries, which results in a smaller supersaturation than in cycle II.
Because of the lower precipitation driving force and the larger distance between
carbides, the overageing treatment to reduce the supersaturation of dissolved carbon
takes longer during annealing treatment I.
Grain boundary Cementite

[II] [I]
Annealing
830

700
[I]

[II] Overageing
350

0
0 120 240 360 480
Time in s
Figure 4.95: Continuous annealing cycles with isothermal overageing and the
associated cementite precipitation.
In Figure 4.96 the following relationships can be seen: an unalloyed steel with a
carbon content of approximately 0.1% is held for various lengths of time at
temperatures between 250 and 450°C and then quenched. Subsequently, the

222
4.4 Precipitating from a Supersaturated Solid Solution

increase in yield strength after an artificial ageing treatment at 100°C is determined.


The longer the overageing time is, the smaller the attainable increase in strength
becomes. As the ageing temperature increases, larger cementite particles form,
which, as a whole, decrease the level of strength. The higher the temperature at
which ageing occurs, the more carbon remains in supersaturated solution after
overageing and the subsequent rapid cooling. During an artificial ageing treatment,
this carbon can then precipitate out in the form of low-temperature carbides, which
increase strength. The resulting level of strength is proportional to the carbon
content left dissolved after overageing.

Overageing temperature
Increase in yield strength* in MPa

60 250 °C

450°C

400°C
40

350°C

20

100 200 1000 2000


Overageing time in s

*After artificial ageing (100°C/1h)

Figure 4.96: Ageing as a function of overageing time and overageing temperature.


Whether a continuously annealed fine sheet still shows ageing potential at the end
of a treatment depends on the extent of overageing, i.e. the reduction of carbon
supersaturation. Ageing can be disadvantageous if it already shows effects before
the cold deformation treatment finished. It can, however, be desirable if an increase
in strength is desired, as is the case with bake-hardening steels, which undergo
artificial ageing during varnishing after cold deformation.

223
4 Phase transformations

Bake-hardening effect
Case study: Bake-hardening effect

Figure 4.97: VW Scirocco with doors made of BH-steels


Bake-hardening steels are steels that are stable against ageing at room temperature,
and harden due to the effect of heat during the varnishing process. Figure 4.98
schematically shows the sequence of the increase in strength, first through
deformation and the associated cold strengthening, and finally through varnishing
under utilisation of the bake-hardening potential. The typical temperature range for
the hardening by varnishing is between 160 and 200°C. The main advantage of
these steels is that they can combine the favorable deformation properties of
conventional deep-drawing steels with the strength properties of high strength
steels, because the increase in strength first begins after deformation.
It is generally assumed that the bake-hardening effect is based on the anchoring of
dislocations through carbon atoms or very fine precipitates. Due to the higher
mobility of atoms during varnishing, carbon atoms arrange themselves in the dilated
tension fields of the dislocations. A prerequisite, however, is that a certain amount
of supersaturated carbon be dissolved in the material. Figure 4.99 reflects this
effect in a schematic stress-strain plot. The dislocation density can only be
increased through the deformation process without a heat treatment. During a heat
treatment, the carbon atoms diffuse to the dislocations and block them with a
“Cottrell cloud”. This results in a distinct yield point and an increase in strength.

224
4.4 Precipitating from a Supersaturated Solid Solution

Figure 4.98: Schematic of the bake hardening effect, by example of the production
of a car door.

Bake hardening B

After deformation,
for example, deep drawing
Stress

Dislocation

Initial state
Carbon atom and
fine precipitate
Strain
Figure 4.99: Metallurgical processes during the bake hardening effect.
The increase in strength can be determined through the supersaturated carbon
content, as shown in Figure 4.100. The upper limit for the amount of dissolved

225
4 Phase transformations

carbon in bake-hardened steels is around 15 ppm, which allows for a storage period
of a few months, in which no effects of ageing appear at room temperature.
Therefore, a carbon content is desired that guarantees ageing resistance at room
temperature, and induces a sufficiently high bake-hardening effect. A carbon
content between 5 and 10 ppm has shown itself to be optimal. With this, a BH 2
value between approximately 20 and 60 MPa can be obtained (see Equation 4.46).
The increase in yield strength through the influence of heat (170°C, 20 min) is
determined by two methods on tensile test specimens. For one, the increase in yield
strength through the influence of heat is measured, which supplies a BH 0 value in
accordance with Equation 4.45. Secondly, the increase in yield strength through the
influence of heat is determined after pre-deformation through a stretching of 2%,
which gives a BH2 value in accordance with Equation 4.46.
BH0 = ReL (after baking) – Rp0.2 (ReL) (in delivery condition) (4.45)
BH2 = ReL (after baking) – Rp2.0 (4.46)
Further explanations to determine the bake-hardening potential are given in the
Stahl-Eisen-Werkstoffblatt SEW 094.
70

60

50
BH2-value in MPa

40

30

20

10

0
0 5 10 15 20
Dissolved carbon in ppm
Figure 4.100: Influence of the amount of carbon dissolved in steel on the yield
strength increase, under the influence of heat (bake-hardening). The
amount of dissolved carbon was determined by means of damping
measurements (Snoek effect).

226
4.5 Further Readings

4.5 Further Readings


Aaronson, H.I.; Enomoto M.; Lee J.K.:
Mechanisms of Diffusional Phase Transformations in Metals and Alloys
1st ed., CRC Press, May 2010

Abbaschian, R.; Abbaschian, L.; Reed-Hill, R.E.:


Physical Metallurgy Principles
4th ed., Cengage Learning, 2008

Bhadeshia, M.K.D.M.; Honeycombe, R.W.K.:


Steels: Microstructure and Properties
3rd ed., Butterworth-Heinemann, 2006

Bhadeshia, M.K.D.M.:
Bainite in Steels
2nd ed., Institute of Materials, 2001

Burke, J.:
The Kinetics of Phase Transformations in Metals
1st ed., Pergamon Press, London, 1965

Gottstein, G.:
Physical Foundations of Materials Science
Springer Verlag, Berlin, Heidelberg, New York, Tokyo, 2004

Hillert, M.:
Phase Equilibria, Phase Diagrams and Phase Transformations, Their
Thermodynamic Basis
Cambridge University Press, Cambridge, 1998

Kostorz, G.:

227
4 Phase transformations

Phase transformations in materials


Wiley-VCH, 2001

Llewellyn, D.T.; Hudd, R.C.:


Steels: Metallurgy and Applications
3rd ed., Butterworth-Heinemann, February 1998

Porter, D.A.; Easterling, K.E.; Sherif, M.:


Phase Transformations in Metals and Alloys
3rd ed., CRC Press, February 2009

Verein Deutscher Eisenhüttenleute (Ed.):


Steel
Volume 1: Fundamentals
Volume 2: Applications
Verlag Stahleisen, Düsseldorf
Springer Verlag, Berlin, Heidelberg, New York, Tokyo, 1992/1993

Zackay, V. Fr.; Aaronson, M.I.:


Decomposition of Austenite by Diffusional Process
New York, London: Interscience Publ., 1962

Barbacki, A.:
The role of bainite in shaping mechanical properties of steel
Journal of Materials Processing Technology 53 (1995), pp. 57-63

Barrow, A.T.W.; Rivera-Doaz-del-Castillo, P.E.J.;


The ε ė η ė θ transition in 100Cr6 and its effect on mechanical
properties
Acta Materialia 60 (2012), pp. 2805-2815

228
4.5 Further Readings

Brown, D.; Ridley, N.:


Kinetics of the pearlite reaction in high- purity nickel eutectoid steels
Journal of Iron and Steel Institute 207 (1969), pp. 1232-1240

Borgenstam, A.; Hillert, M.; Ågren, J.:

Metallographic evidence of carbon diffusion in the growth of bainite

Acta Materialia 57 (2009), pp. 3242-3252

Caballero, F.G.; Miller, M.K.; Badu, S.S.; Garcia Mateo, C.:


Atomic scale observations of bainite transformation in a high carbon high
silicon steel
Acta Materialia 55 (2007), pp. 318-390

Cohen, M.:
The strengthening of steel
Trans. AIME 224 (1962), pp. 638-657

Fielding, L.C.D.:
The bainite controversy
Material Science and Engineering 29 (2013), pp. 383-399

Hillert, M.; Hoglund, L.; Ågren, J.:

Role of carbon and alloying elements in the formation of bainitic ferrite

Metallurgical and Materials Transactions A 35 (2004), pp. 3693-3700

Kaufman, L.; Cohen, M.:


Thermodynamics and Kinetics of Martensitic Transformations

229
4 Phase transformations

Progress in Metal Physics 7 (1958), pp. 165-246

Lipson, H.; Petch, J. N.:


The Crystal Structure of Cementite, Fe3C
Journal of Iron and Steel Institute 142 (1940), pp. 95-106

Mehl, R.F.; Hagel, W.C.:


The Austenite: Pearlite Reaction
Progress in Metal Physics 6 (1959), pp. 74-134

Nutting, J.:
The physical metallurgy of alloy steels
Journal of Iron and Steel Institute 207 (1969), pp. 872-893

Reynolds, W.T.; Li, F.Z.; Shui, C.K.; Aaronson, H.J.:


The Incomplete Transformation Phenomenon in Fe-C-Mo-Alloys
Metallurgical Transactions A 21A (1990), pp. 1433-1463

Reynolds, W.T.; Liu, S.K.; Li, F.Z.; Hartfield, S.; Aaronson, H.J.:
An Investigation of the Generality of Incomplete Transformation to Bainit
in Fe-C-X-Alloys
Metallurgical Transactions A 21A (1990), pp. 1479-1491

Shiflet, G.J.; Aaronson, M.I.:


Growth and Overall Transformation Kinetics above the Bay Temperature
in Fe-C-Mo-Alloys
Metallurgical Transactions 21A (1990), pp. 1413-1432

230
4.5 Further Readings

Song, W.; von Appen, J.; Choi, P.; Dronskowski, R.; Raabe, D.; Bleck,
W.:
Atomic-scale investigation of ε and θ precipitates in bainite in 100Cr6
bearing steel by atom probe tomography and ab initio calculations
Acta Materialia 61 (2013), pp. 7582-7590

Song, W.; Choi, P.; Inden, G.; Prahl, U.; Raabe, D.; Bleck, W.:
On the spheroidized carbide dissolution and elemental partitioning in
high carbon bearing steel 100Cr6
Metallurgical and materials transaction A, 45A (2014), pp. 595-606

Speich, G.R.; Leslie, W.C.:


Tempering of Steel
Metallurgical Transactions 3 (1972), pp. 1043-1054

Stone, H.J.; Peet, M.J.; Bhadeshia, H.K.D.H.; Withers, P.J.; Babu, S.S.;
Specht, E.D.:
Synchrotron X-ray studies of austenite and bainitic ferrite
Proceedings of the Royal Society A 464 (2008), pp. 1009-1027

Swanson, W. D.; Parr, J. G.:


Transformation in iron- nickel alloys
Journal of Iron and Steel Institute 202 (1964), pp. 104-106

Takahashi, M.:

Recent progress: kinetics of the bainite transformation in steels

Current Opinion in Solid State & Materials Science 8 (2004), pp. 213-217

231
5 Technical Heat Treatments

5 Technical Heat Treatments


According to the definition of the Internationalen Verbandes für die
Wärmebehandlungen der Metalle (IVW) and following the DIN EN 10052, heat
treatment is “a process, in the course of which a workpiece or a section of a
workpiece is purposely exposed to one or more temperature-time cycles and if
necessary, other additional physical or chemical influences in order to bring about
the desired characteristics of processing and/or use.” According to the nature of the
influences, the heat treatment processes can be divided into thermal, thermo-
chemical, and thermo-mechanical processes (Figure 5.1).

Figure 5.1: Important processes for the heat treatment of steel.


Among the thermal processes, annealing, hardening, and tempering treatments are
often differentiated. Annealing is generally concerned with influencing the
microstructure in order to improve the processing properties, such as machinability
or cold formability. Hardening and tempering are primarily used to optimize the
end-usage properties, such as hardness, toughness, or wear resistance. A change in
material properties is generally obtained by means of the following processes:
x Change in the size, shape, and orientation of microstructural components
(e.g. coarse grain annealing, soft annealing),
x Reduction of internal stress and change of its distribution
(e.g. stress-relief annealing, tempering),

232
5 Technical Heat Treatments

x Transformation of microstructural components, whereby a state of


equilibrium may be obtained in practice of normalizing, or may be not in
practice of hardening.
Every heat treatment consists of the following individual processes: heating to a
target temperature, maintaining the temperature over the entire cross-section for a
certain amount of time (holding), and cooling from the target temperature
(Figure 5.2). Therefore, the course of a heat treatment can be represented in the
form of a temperature-time diagram.

Target temperature

Surface Center Surface Center


Temperature

Temperature difference
surface/center

Heat penetration time


Heating time Holding time Cooling time
Complete heating time
Heating + holding time

Time

Figure 5.2: Temperature-time diagram of a heat treatment with characteristic


parameters.
It is important to consider that not only the properties of the material, but also the
size and shape of the workpiece, and the duration of the individual processes have
an influence on the selection of a heat treatment.
As the heating rate and the dimensions of the component increase, thermal
conduction leads to a larger temperature difference between the surface and center
of a component during heating (heating duration and duration of heat penetration).
Poor thermal conductivity, as can be found in high-alloyed steels, amplifies the
temperature difference and leads to the formation of thermal stresses. Since the
temperature difference can lead to distortions and stress associated cracking during
heating, the heating rate must be adjusted to suit the material and its dimensions.
The same is true for the cooling rate, which must be selected so that the desired

233
5 Technical Heat Treatments

microstructure can be achieved while the residual stresses remain relatively low.
The holding time is usually chosen based on experience.
For non-alloyed steels, the characteristic temperature ranges for heat treatment
processes are given as a function of carbon content in the iron-carbon diagram
(Figure 5.3). In steels with a larger amount of alloying elements, the temperature
ranges shift according to the change in the transformation temperatures.

234
5 Technical Heat Treatments

1600 1600
L+ G
1536
1500 G 1493°C L 1500
GJ
A4 1392 1400
Tempering colors
LJ
1300 1300 Yellow white

1200 Coarse grain


annealing
1200 Pale yellow
Diffusion annealing 1147
2.06

1100 1100 Yellow

J 1000
1000
Yellow red
Temperature in °C

A3 911
Nor A CM J + Fe 3C
ma
lisin 900
g and
har Pale red
den
ing
A2 769 800
J+ D
A1 723 Cherry red
Soft annealing
700 700
Dark red
D
Brown red
Stress-relief annealing 600
600
Dark brown
Recrystallisation annealing
500 500

D + Fe 3C 400
400 Tempering colors
Grey
Blue grey
Pale blue
300 300 Blue
Magenta
M S
Red
200 Yellowbrown
200 Yellow
Hypoeutectoid steel Hypereutectoid steel

100 100

20 20
0 0.5 0.8 1 1.5 2

Carbon content in mass-%


Figure 5.3: Temperature ranges of various annealing treatments for Fe-C alloys.

235
5 Technical Heat Treatments

5.1 Hardening

The goal of hardening is to produce a martensitic microstructure, since this


microstructure is characterised by high hardness. According to DIN EN 10052,
hardening also applies to cases in which a small amount of bainite, in addition to
martensite, appears in the microstructure. A transformation exclusively to
martensite is not always possible, depending on the steel and the conditions during
hardening.
5.1.1 Definitions

According to DIN EN 10052, two types of hardening can be distinguished:


1. Hardening by means of austenitisation and cooling, such that austenite
transforms completely or partially into martensite, and additionally, if the
case may be, into bainite. Here, steels with a carbon content of 0.3 mass-%
and higher are used.
2. Case hardening consists of carburising or carbonitriding and subsequent
hardening. In this case, steels with a low carbon content (<0.25 mass-%) are
used.
The ability of a steel to increase its hardness over a certain cross-section through
transformation into martensite is referred to as hardenability. This term makes
reference to the maximum achievable hardness under ideal conditions (hardening
capacity), and the distribution of hardness from the surface to the core of the
component (hardness penetration). The hardness penetration is defined generally by
the hardness penetration depth, which is the distance perpendicular from the surface
of a component to a defined microstructure (e.g. martensite) or when a predefined
hardness is reached.
After hardening, construction materials and tools are often subjected to additional
tempering, either to optimally adapt their strength to the stress and strain conditions
that they will be subjected to, or to minimise the risk of cracking during further
processing. Tempering at high temperatures establishes a relationship between
strength and toughness that is appropriate to the stress and strain conditions (Figure
5.4). After quenching and tempering, the properties of steels are determined by the
three processing steps (austenitisation, quenching, tempering), the chemical
composition and the dimensions of the component.
5.1.2 Quenching and tempering

According to DIN EN 10052, quenching and tempering is a combined heat


treatment that consists of the following steps:
austenitisation: heating and penetration of the heated component and
homogenisation of the microstructure,

236
5.1 Hardening

quenching: rapid cooling in a quenching medium (oil, water, air, etc.),


tempering: re-heating and holding at a temperature below AC1.
Figure 5.4 shows the progression of temperature against time during the quenching
and tempering process. The left side of the graph illustrates the austenitisation
process with a TTA diagram. The transformation process that occurs during
quenching is represented in the respective cooling curve and the TTT diagram of the
selected steel, as shown in the center of the figure. The right side of the figure
describes the progression of temperature over time during tempering.

Figure 5.4: Diagram of temperature progression during quenching and tempering.


There is a relationship between carbon content and the maximum possible hardness
when a microstructure consists almost completely of martensite (99.9%). However,
in this case only the amount of carbon that is dissolved in the austenite during
austenitisation is considered. Figure 5.5 shows this correlation. For example, in a
fully martensitic microstructure (99.9% martensite), a maximum hardness of 42
HRC can be reached at a carbon content of about 0.15 mass-% and a maximum of
65 HRC can be reached at a carbon content of about 0.6 mass-%. Higher carbon
levels do not lead to any further increases. Values between those given can be
approximated using Equation 5.1. Table 5.1 gives typical values to be substituted
for K, which are chosen according to the amount of martensite present in the
microstructure after quenching.
Maximum hardness = K + 50 * (mass-% C) r 2 in HRC (5.1)

237
5 Technical Heat Treatments

Martensite fraction in % K
100 35
80 30
50 23
Table 5.1: K values depending on the martensite fraction.
Martensite in Vol.-%
99.9
95
60 90
80
50
Hardness in HRC

40

Max. hardness according


to Burns
20
Hardness at different martensite
contents according to Hodge
and Orehoski

0
0 0.2 0.4 0.6 0.8 1.0
Carbon content in mass-%
Figure 5.5: Relationship between carbon content and maximum achievable
hardness, as dependent on martensite content after hardening.
This relationship is valid, independent of whether an alloyed or unalloyed steel is
being considered. If the microstructure includes components other than martensite
after hardening, then the above mentioned relationship can be used to estimate the
achievable hardness when the K values corresponding to Table 5.1 are inserted.
To approximate the maximum achievable hardness for a fully martensitic
microstructure (dashed line in Figure 5.5), the following equation is valid:
Maximum hardness 20  60 mass - % C in HRC (5.2)
The achievable hardness in the center of a workpiece is determined by the cooling
rate, which decreases from the surface to the core and leads to different
microstructures due to the austenite transformation in accordance with the TTT
diagram of the observed steel. One indication for a high hardening penetration
depth is when the transformation from austenite to pearlite and/or bainite is
suppressed at relatively high cooling rates. Subsequently, the desired core hardness
will be reached more easily. In contrast, a low hardness penetration depth is present
when the microstructure does not transform completely into martensite, even during
a very abrupt cooling.

238
5.1 Hardening

While the hardening capacity depends directly on the carbon content, the hardness
penetration depends on the type and amount of alloying elements. For the elements
that slow down the pearlitic transformation behavior of steels, the hardness
penetration will increase. Therefore, molybdenum, chromium, and manganese are
specifically used to increase the hardness penetration. Carbon also increases
hardness penetration, however, only slightly. Furthermore, hardening is influenced
by the austenite grain size: as the grain size increases, so does the hardness
penetration, since larger grains mean fewer transformation nuclei and therefore, a
larger transformation inertia.
Austenitisation
The hardening capacity of steels essentially depends on the carbon content
dissolved in the austenite before quenching. In order to choose the best
austenitising temperature, the following rules must be taken into consideration:
Non-alloyed hypoeutectoid steels: For these steels, the hardening temperature
should be about 30 to 50°C above the GS-line (A3-temperature) in the Fe-C diagram
(Figure 5.6). This guarantees the complete dissolution of the soft ferritic
microstructure components and a homogeneous carbon distribution.
Non-alloyed hypereutectoid steels: Hypereutectoid steels are austenitised above the
SK-line (A1-temperature) at about 780-800°C as shown in Figure 5.6. Figure 5.7
shows that for such steels with C > 0.8%, the M f-temperature is below room
temperature. Therefore, after quenching in water, a fully martensitic microstructure
cannot be produced from the homogenous austenite; retained austenite remains in
the microstructure, which reduces the hardness. After quenching from the
heterogenous γ + Fe3C area, the hardened microstructure consists of fine, needle-
like martensite with embedded, undissolved carbides and maybe small amounts of
retained austenite.

239
5 Technical Heat Treatments

Figure 5.6: Position of austenitising temperatures in the iron-carbon diagram.


The graph of hardness as a function of carbon content is shown for various
austenitising treatments in Figure 5.7. Curve 1 reflects hardness after quenching
from homogeneous austenite. Above a carbon content of 0.8%, hardness decreases
due to an increasing formation of retained austenite. Curve 2 shows the hardness
after quenching from the two-phase region (J+ Fe3C) at carbon contents greater
than 0.8%. Since carbide hardness is approximately equal to that of martensite, the
hardness of a steel does not change, although the amount of iron carbides increases
with increasing carbon content. Hardening from the J-range above the ES-line, i.e.
after complete dissolution of the existing carbide precipitations, would lead, on the
one hand, to an undesirable increase in grain size due to the high temperatures. On
the other hand, dissolving the existing carbide precipitations would increase the
amount of retained austenite due to the high levels of dissolved carbon. Curve 3
shows the hardness after quenching from homogeneous austenite and a complete
martensite transformation (e.g. by deep cooling). A small amount of retained
austenite in the J-structure can only be attained after austenitisation, if the steel is
quenched to temperatures substantially below the freezing point of water.

240
5.1 Hardening

Figure 5.7: Martensite hardness of pure iron-carbon alloys, dependent on carbon


content, austenitising temperature, quenching temperature (Burn’s
curve), and the course of the Ms/Mf-temperature; shaded area: region of
retained austenite appearance after quenching in ice-water. Bottom
graph: retained austenite content as a function of carbon content after
common hardening.
In alloys with carbide forming elements, high austenitising temperatures between
840 and 870°C are unavoidable, if a sufficient amount of carbides is to be dissolved
in the matrix. The embedded alloy carbides prevent a coarsening of the grains, but
as nuclei, favor the transformation into pearlite, and therefore, increase the critical
cooling rate. Although, when the carbides dissolve, they enrich the austenite with
alloying elements, which has the effect of decreasing the cooling rate. Finally, the
holding time during austenitisation must be selected such that a homogeneous
temperature distribution is achieved throughout the workpiece.

241
5 Technical Heat Treatments

For quenched and tempered steels, the austenitising temperature can be estimated
with the following equation; the content of alloying elements is given in mass-%:
TA >qC@ | 947  264 %C  8%Mn  45%Si  5%Cr
(5.3)
 74%Al  10%Mo  23%Ni  94%V
For unalloyed and alloyed steels with a carbon content above 0.8 mass-%, the
austenitising temperature is usually between 780 and 820°C, and for cast irons
between 850 and 880°C. High-alloyed tool steels are austenitised in the
temperature range between 950 and 1100°C for hot and cold work steels, or
between 1150 and 1230°C for high-speed steels. The appropriate temperature range
for every steel can be found in the quality standards, for example in DIN EN 10 083
for quenched and tempered steels, in DIN EN 10 085 for nitrided steels, in EN ISO
4957 for tool steels, or in other appropriate sources, such as steel manufacturer
catalogues or steel keys. Since grain growth, which can reduce the desired
properties of a hardened state, can be expected during austenising, austenitising
temperatures that are too high, and holding times that are too long are to be avoided,
if possible. Extremely high temperatures can cause irreversible damages to the
microstructure through melting.
Quenching
Quenching can take place using various mediums. In order to produce a fully
martensitic microstructure, the cooling rate in all parts of the workpiece must be
above the upper critical cooling rate of the material. The cooling rate that is critical
to suppress the formation of pearlite and bainite is mainly influenced by the
following factors:
1. Hardening temperature and holding time: an increase in temperature or
annealing time decreases the critical cooling rate.
2. Steel composition: the critical cooling rate decreases as the carbon content
(up to 0.9%) and alloying elements increase.
A reduction of the critical cooling rate corresponds to a shift in the TTT diagram to
the right. Figure 5.8 shows the influence of carbon content on the critical cooling
rate after austenitisation at 1000°C. The increase in the critical cooling rate at a
carbon content above 0.9% C is the result of carbides, which are not dissolved
during austenitisation. These undissolved carbides serve as preferred nucleation
spots for the pearlitic transformation, and, therefore, speed up this process. The
cooling rates in the different zones of the workpiece are dependent on:
x The specific heat capacity and the thermal conductivity of the steel,
x The size, shape, and surface quality of the component,
x The heat transfer in the workpiece/quenching medium boundary layer,

242
5.1 Hardening

x Type, concentration, temperature, and convection of the hardening medium


(with decreasing quenching effects in the following order: salt-water bath,
water, polymer solution, oil, hot bath, air).

Figure 5.8: Critical cooling rate, which when exceeded, results in a complete
martensitic transformation of plain carbon steels; austenitisation at
1000°C.
All desired quenching effects can be achieved by dissolving salts in water or by
applying acids, bases, or various types of oils and emulsions, as well as hot baths
(metal or salt baths). In industrial heat treatment facilities, aqueous polymer
solutions are being more frequently used as a quenching medium, replacing oils and
water. The most important advantages of polymer solutions over oils are:
x No danger of fire,
x Smoke or oil steam does not develop,
x No surface deterioration by oil,
x Greater quenching intensity as compared to oils, combined with the
possibility of using less expensive, low-alloyed steels, and

243
5 Technical Heat Treatments

x No degreasing of the hardened pieces.


The lower quenching intensity, compared to water hardening, results in fewer shape
distortions and a lower risk of crack formation. When using air or a hot-bath as a
hardening medium, the dissipation of heat from the workpiece occurs through
thermal conduction and convection. The cooling effect of these mediums is the
greatest just after cooling has begun, and decreases as the temperature of the
component also decreases. In water, aqueous solutions and oils, cooling takes place
in three steps (Figure 5.9). At the beginning of the quenching process, a closed
film forms around the component, which acts in an insulating manner (film phase,
Leidenfrost phenomenon). As a result, the heat flow to the quenching medium is
low (approx. 210 Js - 1 cm – 2 ), which is associated with a relatively low cooling rate.
At a certain temperature, i.e. the Leidenfrost temperature, the steam film breaks up,
resulting in a direct contact between the quenching medium and the surface of the
component (boiling phase). The intensive vaporisation of the bubbles that follows,
causes a sharp increase in the heat flow from the component (max. around 840 Js- 1
cm – 2 ), and accordingly, a high cooling rate. Below the boiling point of the
quenching medium, heat flow only occurs through convection (convection phase)
with an appropriately low heat flow level (approx. 85J s-1cm-2) and low cooling rate
(Figure 5.9, Figure 5.10). Adding salt causes not only a significant increase in the
cooling rate, but also suppresses or prematurely breaks up the steam film.
Therefore, the boiling process is more uniform than in water. Transformation
should be completed within the boiling phase in order to prevent a pearlitic or
bainitic transformation in the core of the hardening workpiece.
Figure 5.10 shows the cooling rate as dependent on the temperature of the
component. The maximum quenching effect is reached during the boiling phase. A
substantial influence on the quenching effect of the hardening medium is flow
within the bath, which bursts the steam bubbles that hinder heat flow.

244
5.1 Hardening

~800°C ~400°C ~200°C

Surface temperature [for oil]


Film phase Boiling phase Convection phase

Figure 5.9: Cooling phases in a liquid quenching medium, with a boiling point
below the hardening temperature.
400

Convection phase Boiling phase Steam skin phase

300
-1
Cooling rate in Ks

200

100

0 100 200 300 400 500 600 700 800


Temperature of the component in °C
Figure 5.10: Graph of cooling rate versus temperature for a silver sphere; paraffin-
based quick-hardening oil, viscosity 7.0 mm2/s at 50°C.
Currently, vacuum furnaces with integrated high pressure gas cooling are being
increasingly used for heat treatment. Reasons for the progressive spread of this

245
5 Technical Heat Treatments

quenching technology include improving the reproducibility of the cooling process


by eliminating the Leidenfrost phenomenon, ecological advantages, as well as
considerable savings in the finishing process by reducing refinishing expenses and
eliminating refining expenses. The heat transfer of gas is dependent on various
factors, such as, the physical properties of the gas, like thermal conductivity and
density. A second factor is the construction of the heat treatment facility, in terms
of the build up of pressure, gas flow, and gas return-flow cooling. Another
influencing variable is the shape and size of the workpiece.
Austempering
The goal of austempering is to produce a state within a material in which the
formation of ferrite and pearlite is avoided, and the partial or complete
transformation into bainite occurs. Bainite generally has a higher level of hardness
and strength than ferrite and pearlite. However, bainite does not always attain the
same high values for hardness as does the hardening process. As is the case with
hardening, austempering requires that the component be heated to and held at the
austenitising temperature. This brings about an austenitic state and dissolves a
sufficient amount of carbon in austenite. The lower limit of the temperature range
for hardening is usually chosen as austenitising temperature. In order to use the
effects from austempering, steels that have a sufficiently high hardenability must be
used. The high hardenability can be obtained by alloying with the elements
chromium, molybdenum, manganese, or boron. As a result, characteristic of
austempering is that the transformation of austenite into ferrite and/or pearlite
during cooling does not occur.
The bainitic state can be obtained either through cooling at a rate sufficient for the
continuous transformation exclusively into bainite, which can be taken from the
TTT diagram for continuous cooling, or through an isothermal transformation into
bainite. For an isothermal transformation, rapid cooling from the austenitising
temperature to a temperature just below 500°C is necessary. Only in certain
exceptions, however, should the temperature drop below the martensite start
temperature. This temperature is to be held until the transformation of austenite to
bainite is more or less complete. Subsequent cooling to room temperature can take
place at any rate.
Tempering
Tempering consists of the heating of a hardened workpiece to a temperature below
A1. During this, diffusion processes occur, which lead to a more stable, less brittle
microstructure. The goal of tempering is to improve the toughness values, although,
a reduction in strength usually occurs. With increasing tempering temperature,
tensile strength and yield strength decrease, while the reduction in area (Z), fracture
elongation (A), and notch impact energy increase. A representation of these
changes is referred to as a tempering diagram (Figure 5.11).

246
5.1 Hardening

2500 100

Rm
2000 80
R m and Rp 0.2 in MPa

A and Z in %
Rp0.2
1500
60

Z
1000 40

500 20
A

0 0
0 100 200 300 400 500 600 700
Tempering temperature in °C
Figure 5.11: Properties of the hardened steel 50CrMo4 as a function of the
tempering temperature.
Because the chemical composition of a steel after hardening corresponds to that
before quenching, all microstructural components (martensite and retained
austenite) are thermodynamically instable. Through tempering, a state closer to
equilibrium is achieved. This process can be divided into different temperature
ranges in which characteristic changes take place:
1st tempering level: from 100 to 150°C: H-carbides (Fe2.4C) precipitate out of
martensite if the carbon content is larger than 0.2%.
2nd tempering level: from 250 to 325°C: transformation of retained austenite to
bainite or martensite.
3rd tempering level: from 325 to 400°C: formation of cementite (Fe3C) and
transformation of H-carbides to Fe3C.
4th tempering level: above 400°C: recovery and recrystallisation of the
martensitic microstructure, during which defects, such as
vacancies and dislocations, are removed.
5th tempering level: above 450°C: formation of alloy carbides in high alloyed
steels.
The given temperature ranges can overlap and are determined by the content of
alloying elements and the heating rate. All of the processes that occur during
tempering are shown in Figure 5.12. If a steel is tempered at a high temperature or

247
5 Technical Heat Treatments

with a long holding time, then the microstructure consists of carbides


homogeneously distributed in a ferritic matrix.

Reduction in area Z in %
Yield strength Rp0.2

Z
1800 60
in MPa

1600 40
1400 Rp0.2 20

1200 0

10
austenite in %
Retained

0
Recovery through cell formation
5
Tempering step

Ostwald ripening of Fe 3C
4 Cr,Mo,V alloy carbides
3 Fe2.4C Fe3 C
2 Decomposition of retained austenite
1 Precipitation of Fe 2. 4C
0 C-enrichment at imperfections
100 200 300 400 500
Tempering temperature in °C
Figure 5.12: Changes in a hardened microstructure and mechanical properties of
the steel 55Cr3 dependent on the tempering temperature.
Depending on the chemical composition and the original microstructure, toughness
can be reduced by tempering in certain temperature ranges. In this case, the 300°C
and 500°C tempering embrittlement can be distinguished.
Tempering brittleness
When low-alloyed steels are tempered after hardening in a temperature range
between 250 and 350°C, fracture elongation and notch impact energy remain at a
minimum in spite of a constant reduction in strength (Figure 5.13).
Tempering brittleness, also known as 300°C embrittlement, correlates with the
formation of carbides in the third tempering step. It is increased through the
segregation of elements such as P, Sb, As, and Sn. During carbide precipitation on
the previous austenite grain boundary, these segregations are strengthened by the
rejection of the above mentioned elements from the growing carbides. The

248
5.1 Hardening

segregation of these atoms weakens adhesion on the previous austenite grain


boundaries, and cracks form, preferably at the carbides that precipitate out there. A
higher degree of steel purity minimises the extent of the 300°C embrittlement.
300°C embrittlement is also described as an irreversible tempering brittleness,
because of the difficulty in re-dissolving the carbides.

Figure 5.13: 300°C embrittlement in the tempering diagram of the steel


40NiCrMo6.
Some steels, especially Mn, Cr, CrMn, and CrNi steels, are sensitive with regard to
their notch impact energy, to the type of cooling after tempering, and reheating to
temperatures around 500°C. The strongest embrittlement is caused by the
segregation of phosphorus at the grain boundaries at temperatures around 500°C.
This phenomenon is reversible and can be undone by heating the component to a
temperature above 650°C so that the phosphorus is homogeneously distributed
again. Subsequently, the component must be rapidly cooled through the
temperature range of 600 to 350°C.
Corrective measures for tempering embrittlement are:

249
5 Technical Heat Treatments

1. Long term annealing just below AC1, followed by rapid cooling so that the
segregations are reduced during annealing. Since the critical temperature
range around 500°C is passed through quickly, they cannot form again. This
heat treatment can, however, only be applied to tempering-resistant steels.

2. Grain refinement increases the grain surface area so that the degree of local
enrichment is reduced.

3. Reduction of undesired impurities.

4. Alloying of elements that inhibit the phosphorous segregation by occupying


the grain boundaries (e.g. Boron).

5. Bonding of the dissolved phosphorus, for example, during formation of TiP


or Ti2C2P4.

Tempering resistance
Tempering resistance is the ability to retain qualities, such as high toughness,
hardness, and strength even at high tempering temperatures. Tempering resistance
is mainly determined by the content of alloying elements. As the tempering curve
in Figure 5.14 shows, the reduction in hardness during tempering decreases with
increasing levels of alloying elements. As a result, the tempering resistance
increases. Hot forming steels and high-speed steels are extreme examples of
tempering resistance: they are secondary hardened steels. After an initial decrease,
their hardness reaches a maximum between 500 and 600°C, which can even be
above the quenching hardness.
Figure 5.15 shows the hardness curves of Jominy tests (see Chapter 5.1.3) after
hardening at two different hardening temperatures, as well as after various
tempering treatments. The reduction in hardness is greatest on the quenched face
and decreases with an increasing distance from the end face. The tempering
resistance of martensite, which is present at the face, is the lowest. It increases
according to the formation temperature of the individual microstructures, over
bainite up to pearlite.
A comparison of both diagrams in Figure 5.15 shows the influence of the hardening
temperature on the tempering resistance. Raising the hardening temperature from
810 to 920°C not only increases the hardenability, but also the tempering resistance
in all areas of the microstructure. More alloy carbides are dissolved at higher
tempering temperatures, which increase the alloying content of austenite before
quenching. Therefore, more alloy carbides can precipitate out during tempering.
The improvement in tempering resistance at higher hardening temperatures is based
on an increased particle hardening through special alloy carbides.

250
5.1 Hardening

900

800

HS 6-5-2
700
Hardness in HV30

80 CrV 2

C 80 W 1
600

X 40 CrMoV 5 1

500

400

300
0 100 200 300 400 500 600
Tempering temperature in °C

Figure 5.14: Tempering hardness of cold-forming, hot-forming and high-speed


steels.

Figure 5.15: Influence of the hardening temperature on the tempering resistance of


the steel 51CrV4.

251
5 Technical Heat Treatments

5.1.3 Examination of hardenability

According to DIN EN ISO 18265, hardenability is usually measured using the


standardised end-quench test, also called the Jominy test. A cylindrical sample with
a diameter of 25 mm and a length of 100 mm is austenitised and then hung
vertically. The sample is quenched with a water-jet, which is set at a specific
pressure and jet-stream diameter, aimed perpendicularly at the lower face of the
sample (Figure 5.16). Through this, the heat of the sample is extracted almost
exclusively axially. The course of the cooling rate follows accordingly, as well as
the resulting progression of transformation.

Figure 5.16: Sketch of the end-quench (Jominy) test: test construction (top);
Jominy curve and measured hardness values against distance from the
quenched face-side (bottom).
Experimental procedure
1. Austenitisation of the sample,

2. Quenching from the face-side with a water-jet: the test construction and
sample dimensions are shown in the top half of Figure 5.16. The cylindrical
specimens will be machined to a depth of 0.4 mm for two oppositing falt
areas,

252
5.1 Hardening

3. Measurement of hardness on the surface along the axis of the cylinder as a


function of the distance from the quenched end, as shown in the bottom half
of Figure 5.16.

Experiment evaluation
Since heat flows in the direction of the rod axis, the cooling rate decreases with
increasing distance from the face. This leads to an increase in the amount of ferrite
and pearlite in the microstructure and to lower hardness values. Depending on the
different microstructures, the hardness of a sample at room temperature changes
along the length of its axis. The hardness of the quenched face (100% martensite) is
a measure for the hardening capacity. The influence of the composition of a steel
on the hardening capacity and hardness penetration is shown in Figure 5.17. While
the hardening capacity mostly depends on the carbon content, hardness penetration,
i.e. the change in hardness as dependent on the distance from the face, strongly
depends on the content of alloying elements. Steels with a high hardness
penetration, like the steel 51CrV4, have the full martensitic hardness even at large
distances from the face, while the hardness of steels with a lower hardeness
penetration, for example the steel C35, rapidly decreases.

Figure 5.17: Jominy curves of steels with different contents of carbon and alloying
elements.
Figure 5.18 shows the influence of the austenitising temperature on hardenability.
Austenitisation at a high temperature slightly raises the hardening capacity and it
distinctly raises the hardness penetration.
If the Jominy test is executed for several melts of a certain steel type, then the
individual Jominy curves can be combined into a hardenability scatter band, as is
shown per example of the steel 34CrMo4 in Figure 5.19. The scatter bands are

253
5 Technical Heat Treatments

defined in the quality standards of steels. They can be used in choosing a steel that
is suitable for a certain component.

Figure 5.18: Hardness progression for a Jominy sample with 0.9% C and 1.4% Cr
after quenching from various austenitising temperatures.

254
5.1 Hardening

Figure 5.19: Hardenability scatter band of the steel 34CrMo4 according to


DIN EN 10083.
Beside the alloying composition, the progression of hardness is also dependent on
the hardening medium, which shows a specific cooling rate based on its quenching
intensity. In practice, the Jominy test is widely used as a reliable predelivery check
for materials, because of its simple execution and its small scatter. However, in
high alloyed steels, which are harder, the Jominy test has a limited application since
no change in hardness takes place over the length of the sample (100 mm).
Therefore, to check the hardenability of, for example, air-hardened materials, the
standard sample is inserted in a larger ingot and the whole is heated and then cooled
in either air or another hardening medium, whereby significantly lower cooling rates
are achieved.
The results of the Jominy test make it possible to predict the hardening behavior of
real components, but this should be combined with experience. As an example,
quenching intensity and the associated heat withdrawal are temperature-dependent.
Furthermore, they are influenced by the shape, size, and surface quality of the
component, as well as by transformed areas with different thermal conductivities.
Therefore, in practice, the test results are only used as an indication.
5.1.4 Calculating the hardenability

Since hardenability is mainly a function of the chemical composition, intensive


efforts have been made to replace the Jominy test by hardenability calculations,
which are based on chemical composition. Equation 5.4 is a general equation,
applicable at all distances from the face and is dependent on the alloying elements
Jx = b0 + bc * %C + bsi * %Si + bMn * %Mn + ... (5.4)

255
5 Technical Heat Treatments

where Jx = hardness dependent on the distance from the face in HRC, x = distance
from the face in mm, b0 = constant, and b = regression coefficient of the alloying
elements at a distance x.
Table 5.2 shows the regression coefficients from Equation 5.4 for a chromium-
alloyed quenched and tempered steel, as a function of the distance from the face.
However, they are only shown for distances up to 9 mm. It has become common
practice to use hardenability calculations to produce high-grade structural steels
with a specific hardness, both for “hardenability driven” melting and melt
evaluation. A workgroup from the VDEh in Germany has worked since 1980 on a
cooperative analysis of current experimental values using a linear multiple
regression process. From this, generally accepted equations, equivalent to
Equation 5.4, were developed for C-, Cr-, CrMo-, MoCr- und MnCr- case hardened
and quenched and tempered steels. The results of the analysis showed that the
scatter of the calculated hardenability curves using these equations was not larger
than that attained experimentally. Therefore, the calculated end-quenched hardness
is being increasingly used as a characteristic for melts by both producers and
consumers. In the Stahl-Eisen-Prüfblatt 1664 from 1996, the formulas and
regression coefficients for various groups of steels, as dependent on their chemical
composition, are defined in terms of general guidelines for the calculation of
hardenability for the Jominy test. The complete table of coefficients, from which
Table 5.2 was taken, can found in the Stahl-Eisen-Prüfblatt 1664. However, it must
be emphasized that because of the selected process, only limited scientific
statements about the materials can be derived from the coefficients. Due to the
simplification in assuming a linear behavior of the parameters, the validity of such
calculations is limited to those chemical compositions that are defined in the
regression population. Therefore, the coefficients listed in Table 5.2 are only valid
for the range of chemical compositions listed in Table 5.3.
Jx b0 bC bSi bMn bS bCr bAl bCu bN
1.5 29.96 57.91 2.29 3.77 - - - -2.65 83.33
3 26.75 58.66 3.76 2.16 - 2.86 - -2.59 59.87
5 15.24 64.04 10.86 - -41.85 12.29 - - -115.6
7 -7.82 81.10 19.27 4.87 -73.79 21.02 - 4.56 -176.8
9 -27.29 94.70 22.01 10.24 -37.76 24.82 38.31 8.58 -144.1
Table 5.2: Coefficients for alloying elements as dependent on the distance from the
face, for chromium alloyed quenched and tempered steels, up to a
distance of 9 mm, after Caspari et al. and SEP 1664.

256
5.1 Hardening

C Si Mn S Cr Mo Al Cu N
Min. 0.22 0.02 0.59 0.003 0.80 0.01 0.012 0.02 0.006
Max. 0.47 0.36 0.59 0.005 1.24 0.09 0.062 0.32 0.015
Table 5.3: Limits for the chemical composition of melts represented in Table 5.2,
after Caspari et al. and SEP 1664.
5.1.5 Case hardening

Processes for the surface heat treatment or general surface strengthening are mainly
applied to components, the properties of which require a separation between the
function of the base material (e.g. toughness of structure components) and that of
the surface layer (e.g. wear and corrosion resistance). A series of surface
strengthening processes exists corresponding to the variety of the components and
materials, which can be classified as mechanical, thermal, thermochemical and
thermomechanical process according to the operation principles. The particular
selection is based on the stress profile of the component, the largely material-related
process target value (e.g. operating depth) and the economic efficiency (Figure
5.20).

Figure 5.20: Overview of heat treatments for surface modification


Case hardening, a thermochemical diffusion treatment that consists of carburizing
with subsequent hardening and tempering is widely used in the industrial. It is
particularly applied to the forged, cyclically loaded components, for example gear

257
5 Technical Heat Treatments

components, in order to increase the strength resistance capacity and fatigue life.
For an economic manufacture of dimensionally accurate components a good
machinability is essential, which is available in steels with a carbon content of 0.1 –
0.3 mass-%. Steels that meet the requirements include case hardening steel (DIN
EN 10084) or the machining steel (DIN EN 10087). Due to the low carbon content
these steels are not suited for immediate use because of the low hardness. The
hardness of the martensitic surface layer and martensite transformation induced
compressive stress increase almost linearly with the carbon content as long as no
residual austenite occurs. In order to not only guarantee the machinability but also
the high strength in a hardened condition, the/an enrichment of the carbon content
in the surface layer, which is called carburization, is effected. Carburization,
subsequent hardening and tempering are summarized as case hardening. The
process flow of case hardening and the resulting carbon content distribution in a
tooth of a gear are schematically illustrated in Figure 5.21.

Figure 5.21: Temperature – time effect and C-distribution of case hardening.


The compressive stress induced during hardening ensures that because of the
superposition of stress components the externally applied stresses lead to a reduced
stress level in the component (because of the superposition of stress components).
In Figure 5.22, the effect of the generated stress profile is illustrated by an example
of shafts. It becomes evident that the tensile stresses caused by binding or torsion at
the surface are reduced by inherent compressive stresses which result from
martensite transformation. In connection with the increased surface hardness this
effect enables an improved resistance under cyclic loads especially on notched
components. Because of this effect and the economically viable hardening depth,
case hardened components reach the highest fatigue and rolling strengths.

258
5.1 Hardening

Figure 5.22: Interaction of residual stress after case hardening with bending
stresses during loading.
Carburisation is preferably carried out at temperatures above 900°C so that the iron
lattice has a face-centered cubic form, which has the greatest solubility for carbon.
In non-alloyed steels, the maximum solubility extends from about 1.2% carbon at
900°C up to 1.7% at 1050°C. In alloyed steels the solubility is lower since larger
amounts of carbon precipitate out as carbides. For carburising and carbonitriding,
liquid, gaseous, and solid mediums are used, in which the component is held until
the desired depth of hardening is reached.
Carburisation results from the equalisation of the concentration between the
material and the carburising medium, which however, does not occur over the entire
cross-section, but only in the outer surface. The success of carburisation depends
on the flowing steps:
x Supplying the surface of the material with a sufficient amount of a carbon-
donating medium.
x Decomposition of the carbon donor after releasing its carbon atoms and their
adsorption at the surface.
x Diffusion of carbon atoms within the outer layers in the direction of the core.
Powders or granulates of charcoal are used as a solid carburising medium, while
hydrocarbon is used as a gaseous medium. Fused salts consisting of sodium
cyanide serve as the carbon donor and alkali chlorides or carbonates as activators.
The carburized work pieces receive the desired performance properties only after
hardening and tempering. The hardening process, which follows carburization
directly or by interposing process steps (e.g. straightening or machining), can be

259
5 Technical Heat Treatments

performed in different ways. Since the carbon concentration over the cross section
of a carburized work piece decreases from the outer surface to the center, the
transformation behavior changes. Compared with the center part, a lower hardening
temperature and a lower cooling rate for the surface layer are required due to its
higher carbon content. Likewise the start and finish temperatures of martensite
formation are different and drop with the increasing carbon content. After the
carburization process the work piece is usually cooled down to hardening
temperature and then quenched, because the applied carburization temperature is
normally higher than the required hardening temperature for the surface or the
center, as shown in Figure 5.23. The direct hardening is the most economical
method but can be applied only if the austenite grain growth is prevented at high
temperature and the following process procedure for the component is only
polishing. If the fine grain stability of austenite is not assured by carburization, the
coarse austenite grain leads to a coarse microstructure of martensite transformation
and induces negative influences on the performance properties of components. By
single hardening, the work piece goes through a controlled cooling process after
carburization in order to achieve a favorable microstructure on the surface. The
work piece can thus be processed and straightened more easily. For hardening the
work piece will be heated to the desired hardening temperatures of the surface and
the center and then quenched. The austenite-ferrite transformation occurs twice,
which cause a regeneration of austenite grains and eliminate the eventually occurred
coarse grains. For the hardening after isothermal transformation in the pearlite
stage, the work piece will be cooled down to the temperature range of the pearlite
stage (~600°C) after carburization and kept at least until temperature equalization.
After the transformation complete, the work piece is heated again to the desired
hardening temperature and quenched. Also during this process a regeneration of
grains and a formation of microstructure with fine grains take place.

260
5.1 Hardening

Figure 5.23: Procedure processes for hardening according to DIN 17210.


During double hardening at first the center part of the work piece is hardened and
then the surface area. The hardening of the center is achieved by quenching after
carburization from a high temperature. Then a reheating process is carried, during
which the center is partly austenitized and the under hardening exists. The two-time
phase transformation in this hardening process leads to a fine grain microstructure
on surface layer as well. Compared with other processes, double hardening process
requires more processing time and costs. Therefore, in industrial practice double
hardening is rarely applied.
After hardening normally tempering process (150 – 200°C) is carried, during which
the carbon from the supersaturated martensite is partly eliminated. The surface
hardness and compressive residual stress are decreased by tempering. The
toughness, machinability and dimensional stability are improved. However, the
swinging and rolling strength as well as the abrasive wear resistance are slightly
reduced. A tempering temperature more than 200°C causes a significant formation
of carbides and a decrease of toughness.

261
5 Technical Heat Treatments

5.1.6 Properties of hardened materials

Strength and hardness


The goal of hardening is to transform as much of the microstructure as possible into
martensite. The level of the resulting hardness is determined by the amount of
martensite and the content of carbon that was dissolved in austenite. The maximum
possible hardness for steels is 65 HRC. If the microstructure contains ferrite,
pearlite, bainite, or retained austenite alongside martensite, then lower values are to
be expected. In cast irons, the microstructure consists of graphite precipitates in
addition to martensite, so that even after a complete transformation into martensite,
the hardness is still below that of steel. In sintered materials, the pores are
responsible for the low hardness. Hardness is also a measure for the strength of a
hardened component: as hardness increases, so does strength. According to the
data in DIN EN ISO 18265, the tensile strength Rm can be calculated from the
hardness using the relationships:
Rm ~ (3.2 to 3.35) * Vickers Hardness, MPa
(5.5)
Rm ~ (32 to 38) * Rockwell C-Hardness, MPa
In the first formula, the multiplication factor 3.2 is used for a hardness of up to
460 HV. Values between 3.2 and 3.35 are used for the range between 465 HV and
650 HV. In the second formula, the multiplication factor 32 is used for the range
from 31 to 45 HRC, and values between 32 and 38 are used for the range from 46 to
58 HRC. The yield strength can be calculated from the tensile strength, although
the degree of hardness R, which is the ratio between the actual hardness and the
maximum possible hardness, must be taken into consideration:
Rp0.2 ~ (0.42 + 0.5 R) * Rm in MPa (5.6)
where the degree of hardening R = current hardness in HRC/maximum hardness in
HRC.
The range of validity has a maximum hardness of 58 HRC or 650 HV, respectively.
In Figure 5.24, the relationship between hardness and tensile strength is shown.
Also for cyclic loading an empirical relationship between the oscillating loads and
the tensile strength Rm can be found (Equation 5.7):
Vbw ~ (0.38 + 0.29 Z) * Rm in MPa (5.7)
where Vbw = oscillating load, Z = reduction in area, and Rm = tensile strength.

262
5.1 Hardening

Figure 5.24: Relationship between hardness and strength of hardened steels,


according to DIN EN ISO 18265.
The deformability of a material generally decreases as the strength increases, as can
be seen in the relationship between the fracture elongation (A) and the reduction in
area (Z) with strength in Figure 5.25. The reduction in area at fracturing can be
estimated in quenched and tempered steels from the degree of hardening R, the
tensile strength, and the carbon content (Equation 5.8):
Z | 110 - (0.06 - 0.03 R) ˜ Rm + 50 * %C in % (5.8)
Increasing the hardness especially reduces the notch impact energy at high rates of
deformation. Calculating this value is not yet possible.

263
5 Technical Heat Treatments

2400

Reduction in area Z, fracture elongation A in %


Tensile strength Rm and yield strength R e in MPa

Tensile strength
2000

Yield strength
1600
Hardness in HRC 70
60
1200 Hardness
100 50
80 40
800
60 30
Reduction in area
400 40 20
Elongation
20 10
0 0 0 100 200 300 400 500 600 700
Tempering temperature in °C
Figure 5.25: Mechanical properties of the steel C 45 after a heat treatment and
tempering.
Hardening stresses
During quenching stresses arise, which can lead to distortions and cracks in the
component. Figure 5.26 shows the temperature-time curve for the surface and core
of a 35 mm diameter cylinder during water cooling. After cooling for 20 seconds
there is a temperature difference of 350°C between surface and core. Because of
the decrease in volume that occurs in a component with a decreasing temperature, a
temperature gradient develops between the surface and the core, and accordingly, a
difference in volume. This can lead to plastic deformation, such as elongation of
the surface compared to the core. This difference in length is associated with
internal stresses, as well as changes in dimension and shape of the component. The
3-axial stress state during cooling can also lead to cracking. Not only cooling, but
also the transformation into martensite causes stresses and changes in volume. With
the help of the Finite Element Method (FEM), the hardness stresses for simple
component geometries can be calculated. First, longitudinal stresses are examined,
which occur during the quenching of a non-transformed cylinder, for example one
made of an austenitic steel.

264
5.1 Hardening

Figure 5.26: Cooling curve for the surface and core of a cylinder with a diameter of
35 mm, quenched in water; steel 51CrV4 with a M s-temperature of
420°C.
At the beginning of cooling, tensile stresses develop in the surface, and compressive
stresses develop in the core, under which the core is plastically compressed, i.e.
shortened. Due to this deformation, the stresses are not completely removed as the
difference in temperature between the surface and core decreases. As the shortened
core continues to cool, the stress situation is reversed: the surface experiences
compressive stresses, and tensile stresses are present in the core. If a steel
undergoing a martensitic transformation, such as the quenched and tempered steel
51CrV4, is cooled in this manner, then an increase in volume during martensite
formation must also be taken into consideration. As in the first case, at the
beginning of cooling, tensile stresses are found at the surface and compression
stresses at the core, with the respective deformation. If the surface reaches the Ms-
temperature, then the austenite begins to transform to martensite, which is
associated with a very large increase in volume in this area, which in turn, leads to a
rapid reversal of the stresses. In this case, very high tensile stresses act on the core,
which can cause stretching or internal cracking in it. During martensite formation
in the extended core, the stresses change their signs. After cooling, the surface is
under tensile stresses while the core is under compressive stresses. Furthermore,
hardening cracks in the surface can be a result of decarbonisation of the surface
layer, caused by the furnace atmosphere or the composition of the salt bath during
austenisation. In this case, the Ms-temperature of the decarbonized surface is higher
than that of the center, which is richer in carbon, so that the outer layer becomes
martensitic at a much higher temperature during quenching. When the core, whose
increase in volume is larger due to the higher carbon content, then transforms at a

265
5 Technical Heat Treatments

lower temperature, high tensile stresses develop in the brittle martensitic surface,
which can lead to cracking. An exact inspection of the oven atmosphere or the
removal of the decarbonised layer can solve this problem.
Due to the previously described changes in volume, the dimensions and shape of a
component change through hardening. The change in dimensions is mainly the
result of the increase in volume that takes place during martensite formation. This
can be seen in Figure 5.27, which compares the specific volume of various
microstructures, for example of a steel with about 0.8 mass-% carbon, based on
their state at room temperature.

Austenite

Tetragonal martensite

Cubic martensite
(tempered 200°C/6.5h)

Pearlite
(annealed)

0 0.120 0.125 0.130


3
Specific volume in cm /g

Figure 5.27: Specific volume of various microstructures of a steel with 0.8 mass-%
carbon.
The diagram shows that the specific volume of the martensitic microstructure is
about 1% larger than that of the initial pearlitic state. As a result, a change of about
3.3o/oo in the linear dimensions occurs. In steels with a lower carbon content, the
increase in volume is slighter. If the content of martensite is below 100%, then the
increase in volume is also smaller, especially due to the greater amount of retained
austenite in the microstructure.
The effect of an increase in volume on the dimensions of a workpiece depends
entirely on the time at which the austenite-martensite transformation takes place in
the individual parts of the cross-section. For instance, the parts of the cross-section
still containing austenite can be plastically deformed by the stresses that develop
during the transformation of the more rapidly cooling surface layer. Therefore, the
actual change in dimension is lower. This plastic deformation is ultimately the
cause of the dimensional change, which is generally known as “distortion”. The
type of deformation that actually occurs in a workpiece depends mainly on its shape

266
5.1 Hardening

and internal stresses, but the method used to manufacture the base material also
influences the course of deformation.

267
5 Technical Heat Treatments

5.2 Annealing Treatments

5.2.1 Diffusion annealing

According to DIN EN 10052, diffusion annealing is defined as annealing at a very


high temperature (usually between 1050 and 1250°C) with an adequately long
holding time (up to 50 hours). Thereby, local differences in chemical composition
resulting from microsegregations are reduced through diffusion, and microstructure
heterogeneity is eliminated.
The heterogeneity of a microstructure can refer to the shape, size, or arrangement of
the microstructural components. It can be divided into the following groups:
x Zonal heterogeneity,
x Anisotropic heterogeneity,
x Isotropic heterogeneity.
Zonal heterogeneity covers relatively large regions, while the shape of the zones,
for example, regions with a larger percentage of segregated phases or areas with
hardened microstructures, depends on the outer shape of the workpiece
(Figure 5.28a). The formation of anisotropic heterogeneity is connected with the
presence of a preferred orientation, which develops, for example, during plastic
deformation. Included in the category “anisotropic inhomogeneity” are, for
instance, the secondary microstructural banding of ferritic-pearlitic steels (Figure
5.28b) and the banded accumulation of carbides in steels with a high carbon content
(carbide banding). Isotropic heterogeneity is not determined by the preferred
orientations within a material. Clusters of irregularly oriented microstructural
components that are evenly distributed throughout the entire volume of a material
are typical for this type of inhomogeneity (Figure 5.28c).

a 1cm b 50Pm c 50Pm

Figure 5.28: Types of microstructural heterogeneity: a) Regions of segregation in a


screw (Oberhoffer etchant) b) Banded ferritic-pearlitic microstructure
of a steel with 0.25% C (Nital etchant) c) Coarse grain cluster caused
by inhomogeneous grain growth in the soft steel St 24 with 0.03% C
(Nital etchant).

268
5.2 Annealing Treatments

Microscopic segregations (microsegregations) lead to the previously mentioned


anisotropic and isotropic inhomogeneities. The cause for this phenomenon is due to
the development of concentration differences during industrial solidification
processes, i.e. solidification under a nonequilibrium state. The difference in
concentration exists because an equilibrium state at the liquid/solid transition is not
reached. As a result, there is always a difference in concentration between the
primarily formed dendrites and the interdendritic spaces. The orientation of the
segregated regions in the microstructure is dependent on the dendritic morphology.
The tendency towards microscopic segregation during solidification is influenced
by the type and amount of alloying and tramp elements present.
In a material that is deformed in a preferred direction, for example, through rolling
or forging, the microscopic segregated regions are mostly elongated lengthwise,
depending on their yield strength. For this reason, crystal segregations generally
have a banded orientation. During cooling from the deformation temperature or
during isothermal annealing into the austenite-ferrite or ferrite-carbide two-phase
region, alloying elements can influence the transformation temperature as well as
the type and orientation of carbon segregation. This can lead to the formation of
secondary banding. The banded structure is characterised by the distance between
the bands and the maximum and minimum local concentrations. Both parameters
define the concentration gradient as perpendicular to the bands. The concentration
profile is influenced by deformation and diffusion annealing.
The complete elimination of such phenomena is only possible if the material is
heated to such a high temperature that the concentrations of alloying and tramp
elements are equal. This is based on diffusion, meaning the migration of dissolved
atoms over distances larger than the lattice constant. The time available for
diffusion, along with the rate of diffusion, which is dependent on the temperature
and concentration gradients, are decisive for the degree of equilibrium obtained by
diffusion annealing. Extremely long annealing times are necessary under
unfavorable conditions, such as small diffusion coefficients, high degrees of initial
segregation, large primary band distances, large segregation coefficients, or the
lowest possible residual segregation.
As the difference in concentration disappears, so does the heterogenous character of
the secondary microstructure. As a result, the material properties, especially
toughness, improve in the transverse direction and are adjusted to the longitudinal
properties (Figure 5.29).
As a result of high temperatures and long holding times, strong scaling,
decarburisation, and the formation of coarse grains can occur. The first two defects
cause a loss of material, which can only be avoided through annealing in a
protective atmosphere. The formation of coarse grains leads to a degradation of the

269
5 Technical Heat Treatments

mechanical properties and must be eliminated by a subsequent heat treatment, such


as normalising.

Figure 5.29: Influence of diffusion annealing (homogenisation, 3h, 1310°C) on the


formation of the microstructure and the mechanical properties of the
hot-forming steel X38CrMoV51.

270
5.2 Annealing Treatments

5.2.2 Coarse grain annealing

Coarse grain annealing is defined as the long term annealing of hypoeutectoid steels
at temperatures above Ac3 with a sufficient holding time and a cooling adequate to
obtain a coarse ferritic-pearlitic microstructure, which improves the machinability.
Steels with a carbon content less than 0.4%, such as case hardened steels, which are
relatively soft and tend to “smear” during machining, are considered for coarse
grain annealing. In this case the term “smearing” is used to describe for one, the
cuttings, and for another, the separation of these cuttings from the workpiece. As a
result, cutting performance is reduced as well as the surface quality of the material.
The automatic manufacturing of workpieces has led to a low smearing tendency,
short-breaking chips, and low tool wear. Altogether these features are associated
with shorter downtimes.
The primary goal of coarse grain annealing is to produce a microstructure favorable
for machining. At a temperature between 1050 and 1300°C and with a holding time
between one to four hours, depending on the material and geometry, a coarse
austenite grain is formed. During the subsequent cooling, this leads to a coarse
ferritic-pearlitic microstructure with poor toughness. When choosing an annealing
temperature, it must be considered that steels of a similar composition can have
distinct differences in their tendency towards the formation of coarse grains. In
comparison to non-alloyed steels, the microalloyed fine grained steels show
discontinuous grain growth. After an initial hindrance in grain growth at low
temperatures, a rapid increase in grain size is obtained only when the range of the
dissolution temperature for the fine precipitated carbides and carbonitrides is
reached. Figure 5.30 displays the grain sizes for the steel 16MnCr5, microalloyed
with Nb, at various annealing temperatures. At low annealing temperatures, a
normal distribution of grain-size classes results in a mostly linear plot. However, at
higher annealing temperatures, two distinct grain-size classes develop and when
graphed, are characterised by a sigmoidal (s-shaped) curve. As the precipitates
begin to dissolve, grain growth initiates at around 1100°C. The presence of uniform
coarse grains throughout the microstructure can only be achieved at annealing
temperatures above 1150°C.
The subsequent cooling from austenite is of great importance for steels that show a
JoD transformation. A suggestion for low carbon steels is rapid cooling to 620-
670°C with a subsequent isothermal transformation to the pearlite range. As a
result, the percentage of proeutectoid ferrite segregated on the grain boundaries is
reduced, and a larger percentage of coarse lamellar pearlite is formed. When
machining this type of microstructure, shearing mainly occurs in the soft ferritic
veins, so that deformability is nearly exhausted before it reaches the blade.
Consequently, chip separation and short brittleness are increased, and tool wear
decreases. Simultaneously, because of the lowered ferrite content, compressing the

271
5 Technical Heat Treatments

cuttings and the tendency towards smearing are reduced, which are manifested in an
improved surface quality (Figure 5.31).
Grain size (ASTM)
14 12 10 8 6 4 2 0
99.87
950°C
97.72
Cumulative frequency in %

1000°C
84.13 1050°C

1100°C
50.00
1150°C
15.87

2.28

0.13
1E-06 1E-05 0.0001 0.001 0.01 0.1 1
Grain surface area in Pm 2
Figure 5.30: Influence of annealing temperature on austenite grain size of a Nb-
alloyed steel 16MnCr5.

272
5.2 Annealing Treatments

Coarse grain annealed Normalized

Forging
Coarse grain annealing Forging
15 min at 920°C Normalizing
Treatment 15 min at 1150°C 15 min at 920°C
30 min at 620-650°C air cooled
air cooled Complete machining
Complete machining Case hardening
Case hardening

Microstructure ( 100 Pm )

Abrasion marker width on


the cutter for v=30m/min,
s=1.6mm/cutting rotations
during cutting
after 49 workpieces 0.18 mm 0.43 mm
during indenting
after 98 workpieces 0.15 mm 0.20 mm

Figure 5.31: Influence of various annealings on the microstructure formation,


surface quality, and tool wear during machining of the steel 16MnCr5.
In contrast, for quenched and tempered steels, a higher percentage of ferrite, which
possesses a higher tensile strength, is desired. In addition to the ferrite, a coarse
lamellar formation of pearlite is also desired. For this reason, slow cooling, with the
goal of forming a ferrite mesh on the austenite grain boundaries, is advantageous.
A spheroidisation of cementite, which would worsen the machinability in steels
with carbon content less than 0.4% must be avoided.
There is a greater tendency that banding structure will form in the secondary
microstructure during slow cooling into pearlite. The large difference in strength
between the ferrite and pearlite bands leads to a poor surface quality, especially
when machining in the direction of the bands. After machining, the coarse grain
annealed microstructure should be refined with appropriate measures, such as
normalising.
5.2.3 Normalising

Normalising describes a heat treatment that consists of austenitising and subsequent


cooling in still air. The annealing of hypoeutectoid steels takes place at
temperatures between 30 and 50°C above the Ac3 point. For hypereutectoid steels,

273
5 Technical Heat Treatments

this happens between 30 and 50°C above Ac1 in order to obtain an even distribution
of ferrite-pearlite or pearlite-carbide after cooling (Figure 5.32). The workpiece is
held at the high annealing temperature until it is completely heated through. After
this, it is cooled in still air. Since the D→J transformation is passed through twice,
at the end, the steel has been converted into a uniform, fine-grained state. This
microstructure state can be both an intermediate stage for further thermal
treatments, like before soft annealing or hardening, as well as an end-use state of the
workpiece for construction steels and cast steels. All changes to the microstructure
and properties caused by hardening, quenching and tempering, superheating,
welding, cold deformation, and hot deformation can be removed through
normalising, as long as these changes did not cause any lasting material damages,
such as seams or hardening cracks. Normalising is recommended in the following
cases:
x Coarse grained or irregular microstructures due to, for example,
microstructures overheated by thermal influences or microstructures with
strong ferrite-pearlite banding,
x Formation of coarse grains after critical deformation,
x Solidified microstructures, for example, in weld seams or casted steels,
x Steels made brittle through ageing by dissolving very fine segregations,
x Aluminum-killed, fine grained steels; in order to increase resistance to ageing
by binding nitrogen as aluminiumnitrides,
x Construction steels with insufficient toughness or a yield point that is too
low, in case quenching and tempering is not an option.

274
5.2 Annealing Treatments

Figure 5.32 shows an example of the microstructure changes caused by proper


normalising. The result of the annealing processes and the associated
improvements in quality is equally dependent on the rate of heating, the
austenitising temperature, the holding time, and the cooling rate. Heating and
cooling should generally take place as rapidly as possible. As a result of this
increased superheating or supercooling, the smallest possible austenite grain size
and a fine grained pearlitic microstructure are created.

Figure 5.32: Changes in the microstructure through normalising; elimination of


microstructural banding (left) and elimination of casted microstructure
(right).
The austenitising temperature should rise just above the transformation temperature
Ac3 (generally 30 to 50°C higher). The consequence of greatly surpassing this
temperature is that a coarse grained austenitic microstructure forms, especially in
hypoeutectoid steels. Assuming equal cooling conditions, this leads to a coarse
secondary microstructure after transformation.
Normalising of hypereutectoid steels is a special case with regard to choosing the
annealing temperature. In hypereutectoid steels, the annealing temperature is
chosen based on the desired manipulation of the microstructure. As a rule, the
annealing temperature is set above A1 if the pearlitic microstructure is only to be
refined. Higher temperatures are not generally recommended, due to the danger that
a cementite network might form during improper cooling.

275
5 Technical Heat Treatments

The holding time at the austenitising temperature conforms with the thickness of the
material. Generally, one minute per millimeter of thickness is sufficient. Although
with an increasing content of alloying elements, the holding time can be somewhat
extended, considering the dissolution of carbides.
Cooling from the annealing temperature requires special attention, since the end
state reached by normalising is significantly influenced by the cooling rate within
the range of phase transformation. The fineness of the pearlitic microstructure
increases with increased supercooling of austenite, or in other cases, as the cooling
rate increases.
Generally, cooling in still air is sufficient. With large cross-sections, cooling with
compressed air or under a water shower may be necessary under certain
circumstances. However, the danger exists that unacceptably large stresses might
develop within the workpiece. Therefore, after transformation is complete (as of
about 650°C), the workpiece is cooled very slowly, or a repeated heating to remove
stresses is recommended.
Special care is necessary when cooling alloyed steels. Cooling at a rate that is too
slow leads to a strong ferrite-pearlite banding or the spheroidisation of cementite
lamellae, which leads to a deterioration in quality. Cooling too rapidly can cause an
undesired bainitic or martensitic transformation.
From the schematic transformation diagram of the steel 15CrNi6 (Figure 5.33), it
can be seen that the formation of bainite can only be avoided through slow,
continuous cooling (curve III). In doing so, the temperature range in which ferrite
bands form is passed through. Consequently, a new annealing principle can be
deduced. It consists of rapid cooling from austenite, a subsequent pearlite
transformation held at a constant temperature, and finally, cooling as desired (curve
II).

276
5.2 Annealing Treatments

Area of the banded structure


1000
3 K/min
30
Temperature in °C

300
P
3000
A F
(II)
500 B

M
400 320 250 170 HV
(I) (III)
0
-2 -1 2 3
10 10 1 10 10 10
Time in minute
(I) (II) (III)

0.1mm 0.1mm 0.1mm

Ferrite - pearlite Ferrite - pearlite Ferrite - pearlite


with bainite banded
Figure 5.33: Influence of various cooling strategies on the microstructural
formation of the steel 15CrNi6.
5.2.4 Soft annealing

Soft annealing is a heat treatment to reduce the hardness of a material to a given


value. It is commonly used to improve the formability of quenched and tempered
steels.
The annealing takes place at a temperature just below or above Ac1, or while
oscillating at the Ac1 temperature. This is followed by slow cooling, in order to
attain a state that is appropriately soft for further processing. The optimum soft
annealed state is characterised by a homogeneous distribution of fine, spherical

277
5 Technical Heat Treatments

carbides in a ferritic matrix. The hardness should be below 207 HB 30. In steels with
carbon content higher than 0.4%, soft annealing is also known as spheroidising due
to the change in carbide morphology
After forging or normalising, the microstructure of carbon-rich steels (C > 0.3%),
such as quenched and tempered steels, spring steels, rolled steels, and tool steels, is
pearlitic, which accordingly has a poor machinability and a poor cold formability.
The lamellar carbides and the potential presence of a coarse carbide network on the
grain boundaries increase the wear on tools, because the carbides must be cut during
machining, and due to their flow hindering effect, they decrease the cold formability
of the material during compression or precision cutting. In contrast, spherical
carbides are either pushed aside or torn out by cutting tools, and therefore, hinder
the flow of the ferritic matrix during cold deformation much less. Figure 5.34
shows the relationship between the microstructure and the resulting material
properties. The relative machinability, as shown in the left-hand graph, gives
information about the cutting rate applicable for a given material. If it is possible to
work with a steel using a high cutting speed, and the tool wear and handling time
remain low, then the relative machinability is judged to be good. A high fracture
elongation value A and a low yielding point Re, as are found after complete carbide
spheroidisation, are advantageous for a good cold formability.

Figure 5.34: Machinability of non-alloyed and low-alloyed steels as a function of


carbon content and microstructure (left) and change in mechanical
properties of steel C75 with increasing spheroidisation (right).
As the soft annealed state is also the starting point for the subsequent hardening of
this type of steel, further requirements arise for the microstructure as determined
mostly by the desired qualities of the product. In order to attain the maximum

278
5.2 Annealing Treatments

hardness and a guaranteed hardening depth, an evenly distributed fine carbide grain
is required. These homogeneous grains ensure hardening with minimal distortion.
An increase in carbide spheroidisation decreases the hardness, but it becomes more
difficult to dissolve carbides. Thus, a compromise must be reached between the
best machinability and good hardenability, which is reached at an average carbide
grain size of about 0.003 mm.
Hypoeutectoid steels are held between 680 and 710°C for a few days in order to
remove the pearlitic structure. Figure 5.35 shows various temperature-time graphs
for the soft annealing of steel.

Figure 5.35: Temperature-time curves during soft annealing of steel.


The ability of D-iron to dissolve carbon increases considerably, compared to room
temperature, until the Ac1 temperature (from 10-6 to 0.02% C). The spheroidisation
process can be divided into two steps. First, the cementite lamellae are separated
into irregularly shaped cementite bodies. Second, these grains approximate a
spherical shape, through which the smaller grains disappear at the expense of the
larger ones. The driving force behind both steps is the minimisation of interfacial
energy.
While carrying out this procedure, the highest temperature of the furnace must stay
below Ac1b. Since the annealing time lasts several hours, parts of the component,
whose temperatures are above Ac1b, will be austenitised so that they undergo a
pearlitic transformation during subsequent cooling in air. These areas are not soft
annealed.
Soft annealing of steels with a carbon content below 0.4% generally leads to poorer
machinability. On the other hand, cold formability is greatly improved through
spheroidisation.
In hypereutectoid steels, the material properties become worse when there is a
cementite network on the grain boundaries surrounding the pearlite grains. Above
Ac1e, the cementite network partially dissolves, and the pearlitic structure
completely dissolves. During subsequent, slow cooling below Ac1b, the

279
5 Technical Heat Treatments

precipitating carbides nucleate on the residual carbides, which leads to a


spheroidisation of the cementite. This process is similar to the isothermal
transformation of austenite into pearlite. However, the damaging carbide network
on the grain boundaries is difficult to eliminate through this type of heat treatment.
In order to avoid long annealing times and the eventual coarsening of the grains,
cyclic annealing is used for these types of steels (Figure 5.36). During cyclic
annealing, the temperature repeatedly alternates from above the Ac1e temperature to
below the Ac1b temperature, which results in a gradual spheroidisation of both the
carbide network and the lamellar carbides.
A further possibility for soft annealing is hardening and high temperature
tempering, i.e. annealing directly below the Ac1b temperature. During this heat
treatment, the most uniform distribution of small carbides is reached. However, this
type of soft annealing is much more expensive than the previously mentioned
processes so that this heat treatment can only be economically justified in a few
cases. In addition, the thorough hardening of some steels can lead to cracking.
1100
Accm A A
1000

A+C A+C
900
Ac1e
800
Ac1b F+C+A F+C+A
Temperature in °C

700 C C
F+C P F+C F+C P F+C
600

500

B B
400

300
Ms Ms
200
M M
100

0
2 2 3 4 2 2 3 4
1 10 10 10 10 10 10 1 10 10 10 10 10 10

Time in s
Figure 5.36: Scheme of soft annealing cycles of hypereutectoid steels through
cycle annealing; represented are only two austenitisations and two
transformations to ferrite and carbide (F+C).
Figure 5.37 shows the influence of a soft annealing treatment on the microstructure
of the steel C100. It can be seen how the heat treatment changes the cementite
structure from a lamellar to a globular form. The latter form is advantageous during
machining.

280
5.2 Annealing Treatments

10Pm 10Pm

Figure 5.37: Influence of soft annealing on the microstructure of the steel C100;
initial state (left) and after soft annealing (right).
5.2.5 Recrystallisation annealing

Recrystalisation annealing is a heat treatment, which has the goal of obtaining grain
renewal in a cold deformed workpiece through nucleation and nucleus growth
without inducing a phase change.
During cold forming, for example, through flanging, deep drawing, cold rolling,
folding, or cold bending, the crystal lattice is greatly distorted. The strength greatly
increases (work hardening), and the remaining strain decreases. The physical
causes for this phenomenon are dislocations, whose retention and multiplication
during plastic formation result in work hardening of the component. During this
heat treatment (recrystallisation annealing), strength decreases and deformability
increases. In metal physics, this corresponds to a rearrangement and elimination of
dislocations. After a certain degree of deformation, recrystallisation annealing can
be used to restore the original properties of the material. In doing so, the crystal
lattice restructures itself. Through successive deformation and annealing, a high
degree of deformation can be achieved.
The progression of recrystallisation annealing can be seen in a hardness-temperature
curve, as shown in Figure 5.38. After tempering of a cold rolled steel at various
temperatures for equal periods of time, it can be seen that hardness decreases with
increasing annealing temperature.
If the microstructure is examined with an optical microscope after annealing,
practically no changes in the characteristic deformed microstructure can be found at
annealing temperatures to the left of the steep negative slope in the curve. The
slight changes in hardness that develop are categorised as submicroscopic
processes. These processes are called recovery. In contrast, within the temperature
range of the sharp decline in the hardness plot and the following plateau,
microstructural changes are seen in the form of grain renewal. This process is
known as recrystallisation. Figure 5.38 also shows transmission electron
microscopy pictures of the microstructure of a 50% cold formed austenitic steel
X10CrNiMoTiB15-15 for each characteristic region of the curve.

281
5 Technical Heat Treatments

Above the specific recrystallisation temperature (TR | 0.4 Tm) for each metal and
alloy, the crystal lattice is rebuilt due to the added thermal energy. With the
application of heat above the recrystallisation threshold, the grain boundaries
provide the energy needed for recrystallisation. The fraction of recrystallised
microstructure is represented by the schematic curve in Figure 5.39 as dependent
on the annealing time at a constant annealing temperature. The curve in this figure
is to be viewed in combination with the hardness curve in Figure 5.38. At a
constant annealing temperature, recrystallisation begins with the sharp decrease in
hardness at time t0, and the fraction of recrystallised microstructure grows
accordingly. When the final hardness is reached after recrystallisation is complete,
the entire microstructure has been renewed, and the curve in Figure 5.39 reaches a
plateau. This plateau corresponds to a 100% recrystallised microstructure.

800°C / 2h
Hardness

700°C / 2h 900°C / 2h

b) 1P m

a) 1Pm 1Pm
c)

Temperature
Figure 5.38: Influence of the annealing temperature on the hardness and
microstructure of a 50% cold formed austenitic steel
X10CrNiMoTiB15-15 at room temperature, with TEM photos: non-
recrystallised microstructure (left); beginning grain renewal (center);
recrystallised microstructure (right).

282
5.2 Annealing Treatments

Figure 5.39: Fraction of recrystallised microstructure as a function of annealing


time (isothermal recrystallisation curve).
The quantitative description of the curve is given by the Johnson-Mehl-Avrami-
Kolmogorov (JMAK) equation (Equation 5.9):
­ § t ·n ½
W (t ) 1  exp ® ¨ ¸ ¾ (5.9)
¯ ©W ¹ ¿
where W(t) = fraction recrystallised, t = annealing time in s, W = recrystallisation
time in s, and n = time exponent.
The recrystallisation time W is connected to the activation energy through the
following equation (Equation 5.10):
­ Q ½
W W 0 ˜ exp® ¾ (5.10)
¯R ˜T ¿
where W0 = time constant in s, Q = activation energy in kJ/mol, R = general gas
constant (8.3143 J mol-1 K-1), and T = annealing temperature in K.
Under simplified conditions with known nucleation and growth rates, the fraction of
recrystallised microstructure W, the recrystallisation time W, and the time exponent n
can be derived.
The recrystallisation temperature TR is primarily dependent on the previous
deformation and the alloying content. The temperature decreases with an increasing
degree of deformation and a decreasing content of alloying elements. In non-
alloyed steels, TR is generally between 450 and 600°C, and for middle- to high-
alloyed steels the temperature range is between 600 and 800°C. The activation
energy for recrystallisation is about 160 kJ/mol for an aluminum killed steel.

283
5 Technical Heat Treatments

The recrystallised grain structure becomes finer as the degree of cold forming
increases. The recrystallisation diagram in Figure 5.40 shows the relationship
between the recrystallised grain size, the degree of deformation and the temperature.
Recrystallisation takes place only when a critical degree of deformation (ususally
between 5 and 20%) is exceeded. At a lower degree of deformation, an undesired
coarsening of the grains can occur due to the low number of recrystallisation nuclei.
As a result, recrystallisation annealing should be avoided, if possible. A coarse
grained microstructure leads to a poor surface quality of the workpiece (orange
peel) during deep drawing and decreases the plastic deformability of the material.
The degree of cold rolling typically applied industrially is between 50 and 85%.

Figure 5.40: Isochronous recrystallisation diagram of electrolytic iron.


In order to accelerate recrystallisation, annealing should be done at approximately
50°C above the recrystallisation threshold, in the temperature range between 620–
700°C for a batch annealing furnace, or 700–850°C for a continuous furnace. In
contrast to normalising, during recrystallisation annealing, grain renewal results
from nucleation and grain growth without phase transformation.
Recovery annealing at a temperature below the recrystallisation temperature can be
used to further reduce strengthening after low cold deformations. Recovery begins
at a temperature around 300°C and solely leads to a rearrangement of the distorted
lattice and not to recrystallisation.

284
5.2 Annealing Treatments

Recrystallisation annealing is mainly used between the individual steps of cold


rolling and cold drawing of sheets, strips, and wires. For the production of cold
strips, two methods of recrystallisation annealing have been adopted industrially.
The first is batch annealing, or the annealing of coiled steel strips under a protective
hood and atmosphere. The other process is continuous annealing, or the annealing
of a strip moving at a high rate in a long furnace. These two processes differ
significantly in the manner in which they proceed.
The main differences between the conventional batch annealing and continuous
annealing are shown in Figure 5.41 with their associated time-temperature cycles.
The different annealing temperatures are chosen to allow for a complete
recrystallisation within the available time. Temperatures above 700°C are to be
avoided during batch annealing because of the danger of “adhesion”, the diffusion-
welding of successive turns in a coil. During the cooling phase, the protective hood
can be removed at around 150°C, and the cooling can continue at an accelerated rate
in circulating air. A temperature of less than 40°C is aimed for, before further
handling of the strip with, for example, skin pass rolling. The differences between
the processes become more obvious by comparing the most important annealing and
cooling parameters, given in Table 5.4.

Figure 5.41: Annealing cycle scheme of batch and continuous annealing.

285
5 Technical Heat Treatments

Annealing and cooling Batch annealing Continuous annealing


parameters

Heating rate 0.01 K/s 10 K/s


Recrystallisation range 560 – 620°C 610 – 670°C
Annealing temperature 650 – 700°C 700 – 850°C
5
Holding time 10 s < 102 s
Cooling rate 0.001 K/s 10 – 1000 K/s
Table 5.4: Characteristic annealing and cooling parameters for batch and
continuous annealing.
The differences in the process parameters for the recrystallisation annealing of a
cold rolled strip have a lasting effect on the structure and mechanical properties of
the annealed strip. Thus, an adjustment of the hot strip rolling to the previous
processing steps during steel production is necessary. The characteristic
metallurgic processes that take place during each of the procedures are listed in
Figure 5.42 for example of an aluminum killed, deep drawn steel. Due to different
heating and cooling conditions as well as different annealing temperatures during
batch and continuous annealing, the temperature range in which each metallurgic
process is effective shifts from that given in Figure 5.42.
The binding of nitrogen as aluminum nitride and the segregation of carbon as
cementite proceeds differently in each process. The content of dissolved nitrogen is
essentially set already during the coiling of the hot strip. Either a high or a low
coiling temperature is chosen, depending on the steel grade and the subsequent
annealing process. Nitrogen is dissolved at low coiling temperatures (<600°C), and
at high coiling temperatures (>700°C), nitrogen bonds to form aluminum nitride.
If a steel is annealed using the batch type process, AlN precipitations that arise
during heating serve as a guiding phase for the attainment of a distinct deep drawn
texture. Through the precipitation of AlN at the grain boundaries of the elongated
ferrite grains, a rapid growth of the ferrite grains during the slow heating period is
suppressed. The elongated “pancake” microstructure, with a distinct {111}-texture,
develops, which is characteristic of batch annealing. Thus, the coiling temperature
must remain low, so that nitrogen stays in solution after hot rolling.
In contrast to batch annealing, a high coiling temperature is desired for the hot strip
during the production of continuously annealed steel grades. This delays the
formation of AlN precipitates that are impedimentary for the texture development
during continuous annealing. Figure 5.43 shows an example of the microstructural
development of a cold rolled IF steel during continuous annealing.

286
5.2 Annealing Treatments

In both annealing processes, cementite dissolves during heating and holding at the
annealing temperature. However, during batch annealing, dissolved carbon can
fully precipitate out again during the slow cooling phase, and as a result, no ageing
potential exists after annealing. Carbon that dissolved during continuous annealing
is present in super-saturated solution after quick cooling. Therefore, an overageing
step is added during the annealing cycle (350-450°C, 2-4 min) so that carbides can
precipitate out as cementite, which minimises the danger of ageing. Depending on
the chosen cooling strategy, the temperature and duration of the overageing
treatment vary.

287
5 Technical Heat Treatments

Figure 5.42: Metallurgic processes during batch and continuous annealing in an


aluminium killed deep-drawn steel with dissolved nitrogen.

288
5.2 Annealing Treatments

Rolling
direction

50 Pm 50 Pm

Figure 5.43: Microstructure of a ferritic steel after cold rolling, degree of


deformation 70% (left) and after recrystallisation annealing, pancake
microstructure (right).
5.2.6 Stress-relief annealing

Stress-relief annealing is a heat treatment that consists of heating and holding at an


adequately high temperature (below the lowest transformation point Ac1), and a
subsequent cooling step suitable for the purpose of removing internal stresses
without significantly changing the microstructure.
Internal stresses can be caused by uneven heating or cooling, through differing
thermal expansion during, for example, welding, soldering, cooling, and solidifying
of casted pieces, or during cold forming processes, such as bending, hammering,
straightening or roughing. During further handling or while in use, these internal
stresses can lead to a distortion of the workpiece or even crack formation and brittle
failure.
Stresses in a workpiece can only be eliminated when they cause a plastic
deformation in the micro-range (dislocation movement). However, this requires
that the yield point of the material is reduced to below the stress value. The lower
the yield point falls below the level of stress, the greater the magnitude of plastic
deformation becomes, and with this, the greater the possibility of eliminating
stresses. In most materials, the strength and yield points naturally decrease with
increasing temperature. Because of this, stress-relief annealing always includes
heating to an appropriately high temperature.

289
5 Technical Heat Treatments

The higher the annealing temperature is chosen to be, the more rapid the relaxation,
and the lower the residual stresses become. Thus, the temperature should be as high
as possible, whereby the upper limit is defined by surface oxidation or possible,
undesired microstructural changes, such as recrystallisation in cold formed parts of
components, tempering of the hardened microstructure, spheroidisation of
cementite, precipitation or embrittelment. In non- and low-alloyed steels, the most
favorable annealing temperature is between 450 and 650°C, with a holding time that
is dependent on the construction component, which can range from several minutes
to several hours.
Stress-relief annealing does not make sense at temperatures below 400°C, since
high levels of residual strain remain in spite of long holding times. Of special
importance for the success of stress-relief annealing, beside heating through and
holding at an optimum annealing temperature, are heating and cooling. Too rapid
or uneven cooling from the annealing temperature introduces new stresses into the
material, which can lead to distortion or cracking, especially in components with a
complex shape.
Complex welded component designs can often show critical stresses that increase
the tendency towards brittle fracturing. Because of this, materials with high
operational demands and low operational temperatures must be thermally relaxed.
5.2.7 Combined annealing processes

The limited applicability of normalising for case hardened and quenched and
tempered steels has led to the necessary development of further annealing processes
that have found their way into the standards of the respective steels. An example is
DIN EN 10083 for quenched and tempered steels, or DIN EN 10084 for case
hardened steels. The following processes are recommended to improve the
machinability through heat treatments, as dependent on the chemical composition of
the steel and the dimensions of the workpiece:
1. Heat treatment to reach a certain tensile strength: the component is cooled
from a temperature between 850 and 950°C and, if required, annealed at
around 500 to 650°C,

2. Heat treatment to attain a ferritic-pearlitic microstructure: the component is


cooled in a controlled manner from a temperature between 900 and 1000°C,

3. Heat treatment to improve the machinability: the component is annealed at a


high temperature, usually above the A3 point, and subsequently cooled, in
special cases according to a certain temperature-time diagram, so that the
microstructure does not exceed a certain hardness after cooling.

290
5.2 Annealing Treatments

5.2.8 Wire patenting

Patenting is a heat treatment for wires or strips that consists of austenitisation and
subsequent cooling in a manner appropriate for obtaining a microstructure that is
suitable for drawing or drawing with roller dies.
Different end products of wires are usually produced through non-cutting cold
forming. Hence, the most important demand on the pre-product “wire” is good
deformability. A microstructure that shows a good deformability and a large
ductility, as well as high strength, is very fine lamellar pearlite, also called sorbite.
The heat treatment for producing this fine lamellar pearlite is called patenting. For
these steels, the thinner the lamellae, the smaller the interlamellar spacing.
Furthermore, the lower the fraction of proeutectoid ferrite in the microstructure,
then the better the deformability and thus the higher the strength. It follows from
this that steels with a carbon content close to the eutectoid composition are the most
suitable to produce a good sorbitic microstructure. These steels generally do not
show proeutectoid ferrite precipitation. As the carbon content increases,
deformability decreases. Moreover, the tendency towards carbon segregation also
increases (banding). Therefore, unalloyed steels with a carbon content between
0.45% and 0.65% are used for rolled wires. Alloyed steels are generally not
patented. Patenting consists of austenitisation, quenching and a subsequent
isothermal pearlite transformation (Figure 5.44).

291
5 Technical Heat Treatments

Figure 5.44: Isothermal TTT diagram with temperature curve of a patenting


treatment.
The processes of austenitisation and pearlite transformation can be summarised as
follows:
x Austenitising: the wire is heated to a temperature above A3. High homogeneity
of the austenitic phase is necessary to obtain good patenting properties.
x Pearlite transformation: the austenitic wire is quenched, and held isothermally or
continuously cooled so that a transformation to lower pearlite occurs. As a
result, proeutectoid ferrite precipitation is limited to a minimum.
Continuous processing is used for patenting when a wire has a small diameter or is
very long. For technical processing reasons, batch patenting, in which the wire is
bunched into a coil and subsequently dipped, can only be used on wires with a large
diameter (over approximately 12 mm).

292
5.3 Description of Austenite Transformation for Technical Applications

5.3 Description of Austenite Transformation for Technical Applications

Technical heat treatment as a combined process of heating (austenite formation) and


cooling (austenite transformation) can be explained under simplified conditions in
special nonequilibrium diagrams. Equilibrium diagrams are only valid for infinitely
slow heating and cooling rates. Under the technical conditions of finite heating and
cooling rates, the transformation temperatures shift. Therefore, to show the kinetics
of austenite formation, time-temperature-austenitisation TTA diagrams are used.
The kinetics of the transformation of supercooled austenite is described using time-
temperature-transformation (TTT) diagrams and continuous cooling transformation
(CCT) diagrams.
Determination of nonequilibrium diagrams
It is generally possible to construct TTA and TTT diagrams by means of
metallographic analyses of quenched samples or by measuring the changes in
physical properties. Changes in physical properties can be determined through
differential thermal analysis, measurement of magnetic or electrical properties, or
through measurements using a dilatometer, which is the most prevalently used
method. The dilatometer measures the change in length ('L), which is dependent
on temperature and time (Figure 5.45). According to the Stahl-Eisen-Prüfblatt
1681 (SEP 1681), test evaluation is either isothermal (T=const.) or continuous.

Figure 5.45: Change in length during isothermal and continuous transformations


for the determination of TTA and TTT diagrams, after SEP 1681.
How to evaluate an experiment involving continuous cooling is shown in
Figure 5.46. The beginning of a transformation range is signified by the deviation
from a straight line in the graph of change in length over time. If the curve
continues into another straight line, then the transformation in that field is finished.
If different microstructure components form directly after one another, then a
turning point in the curve signifies the boundary between transformation ranges.

293
5 Technical Heat Treatments

Finally, the temperature is graphed over the logarithmic time axis in a


nonequilibrium diagram. This makes it possible to investigate the transformation
behavior of the steel from short to long periods of time. The accuracy of
nonequilibrium diagrams varies for temperature by ±10K and for time by ±10%.
Variations due to different chemical compositions and initial states can further
increase the range of distribution. All diagrams are valid, in the strictest sense, only
for the chemical compositions observed and the conditions given for each specific
case (austenitising temperature and austenitising time).

Figure 5.46: Example of how to evaluate experiments using continuous cooling,


after SEP 1680.
In every nonequilibrium diagram, a multitude of transformation points are given
along with the specifications for the material and experimental procedures, in
accordance with the Stahl-Eisen-Prüfblatt 1680 (SEP 1680). Holding points during
heating are noted with the symbol Ac (French: “arrêt au chauffage”). The following
temperatures can typically be found in a TTT diagram. The meaning of their
abbreviations (see also Figure 5.47) are given below (according to SEP 1680):
Ac1 Temperature at which austenite formation begins during heating (temperature
of the DJ+ carbides three-phase equilibrium).

294
5.3 Description of Austenite Transformation for Technical Applications

Ac1b For non-alloyed and alloyed steels: temperature at which the formation of
austenite begins during heating (beginning of the DJ + carbides three-phase
field).
Ac1e Temperature at which the first transformation during heating ends, i.e.
moving from three-phase region (DJ + carbides) to the DJ or J + carbides
two-phase area.
Ac3 Temperature at which the transformation of ferrite into austenite ends during
heating.
Accm Temperature at which the dissolution of cementite in austenite ends during
heating of hypereutectoid steels.
Acc Temperature at which the dissolution of carbides in austenite ends during
heating of alloyed steels.
Ms Temperature at which the transformation of austenite to martensite begins
during cooling (Martensite-Start).
Mf Temperature at which the transformation of austenite to martensite nears
completion during cooling (Martensite-Finish).
If the transformation temperatures are given for cooling, then they are abbreviated
as Ar; the letter r refers to a holding point during cooling (French: “arrêt au
refroidissement”).

295
5 Technical Heat Treatments

5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams)

TTA diagrams provide information concerning the course of austenite formation


dependent on temperature and time. They describe the dissolution of carbides, as
well as the homogenisation and changes in grain size of austenite. The formation of
solid solutions is one prerequisite to adjust a supercooled state, i.e. simulating an
industrial cooling. This is accomplished by austenitisation, in which a homogenous
austenitic crystal (J-solid solution) is generally desired. From this crystal the
desired microstructure can be adjusted, according to the cooling conditions. The
near-equilibrium state of austenite in non-alloyed, and in a broader sense, in low-
alloyed steels is described by the iron-carbon-alloying element (Fe-C-M) diagram
(Figure 5.47).
In this case, the alloying elements M can be, among others, small additions of the
carbide-stabilising elements chromium, manganese, and molybdenum. Determining
an equilibrium diagram, which in turn makes it possible to determine the
transformation temperatures Ac, is usually carried out at an “(in)finitely slow”
heating rate of 3 K/min. This ensures that the transformation temperatures are
precise enough to match the equilibrium temperatures for technical purposes.
The initial state for austenitisation is ferrite, which can only dissolve a few
thousandths of a percent of carbon, and carbide M3C, with a mass content of 6.9%
carbon, as well as the alloying element M. If such a steel is heated to a temperature
where austenite is stable, then carbon is homogeneously distributed throughout
austenite. The carbides, which are present beside ferrite at low temperatures below
the A3 temperature, must be dissolved in the J-solid solution. The distribution of
carbon and the formation of austenite are time-dependent due to the limited
diffusion rate of carbon in iron.
Technical austenitisation progresses mostly isothermally, meaning that the
component is heated to a desired temperature and then held at that temperature.
During continuous austenitisation, the component is rapidly heated to a desired high
temperature and then immediately cooled. Such temperature-time behaviors result
from, for example, flame hardening and induction hardening. In both cases, the
course of austenite formation can be described in respective diagrams.
The relationship between a continuously recorded TTA diagram and an equilibrium
diagram is established by placing the middle axis in a three dimensional diagram
perpendicular to the phase diagram. The points of transformation Ac3, Aclb, and
Acle for infinitely slow heating coincide with those of the nonequilibrium diagram
(Figure 5.48).
The near-equilibrium diagram can be seen as an extreme case of a TTA diagram for
an infinitely slow heating. A respective representation can be created for an

296
5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams)

isothermal TTA diagram. The equilibrium diagram then corresponds to an extreme


case for austenitisation with an infinite holding time.
1100

1000

J
Temperature in °C

900
Ac 3 Accm
J+M3C
800
DJ Ac1e

700 Ac1b
D DJ+M3C

600
D+M3C
500
0 0.2 0.4 0.6 0.9 1.0 1.2 1.4
Carbon content in mass %
Figure 5.47: Section through an iron-carbon-alloying metal M three-component
system, with low M content, in order to define the Ac points.

297
5 Technical Heat Treatments

1.0
0.8
0.6 %C
1000 100 10 1 °C/s 0.4
920
Ac3 0.2
900 0
920
880
J 900
860
880
J
840
Temperature in °C

860
Ac1e
820
DJ 840
800 Ac1b
820
780 DJ
800
760
780
740
D+M3C 760
720
740
700
D+M3C 720
680 2 3
0.1 1 10 10 10

8
700
Time in s
680

Figure 5.48: Phase diagram as an extreme case of the TTA diagram for continuous
heating with an “infinitely” slow heating rate.
An isothermal TTA diagram is determined by heating thin samples to a specified
austenitising temperature, quenching them after different austenitisation times, and
then metallographically analysing them (Figure 5.49).
In order to determine continuous TTA diagrams, samples are heated at different
rates, then immediately quenched after having reached a certain austenitising
temperature, and finally evaluated metallographically (Figure 5.49). In both cases, it
is possible to determine the transformation points by measuring the change in
physical properties during the austenite transformation. The beginning and end of
the phase transformations are designated with Aclb, Acle, Acl, and Ac3. Lines
connecting similar transformation points show the chronology of austenite
formation as a function of the heating rate. They are summarised in a T-log-t
diagram.

298
5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams)

Heating (TTA) Cooling (TTT)

T T
T1
Isothermal

T2 T1
T2

t t

T T
T1

T1 T2 T3

T1 T2 T3
Continuous

t t

T T
T1

T1 T2 T3

T1 T2 T3
log t log t

Figure 5.49: Schematic of heating and cooling processes for isothermal and
continuous transformations to determine TTA and TTT diagrams.
5.4.1 Austenitisation with isothermal heating

TTA diagrams that are determined through isothermal heating (Figure 5.50) are
interpreted by looking at a specific temperature and observing what happens at
different times for this specific temperature, e.g. at what time austenite formed.
After the respective isothermal temperature is reached, the counting of time begins.
Additionally, iso-hardness lines are shown, which were determined from samples
that were cooled so rapidly after a specified holding time, that they underwent a
martensitic transformation. The diagram shows, when heating to 800°C with a
heating rate of 130 K/s, that the formation of austenite begins after a holding time of
0.3s, because the Aclb temperature is exceeded. By doing so, austenite grows

299
5 Technical Heat Treatments

between the cementite lamellae into pearlite. After the complete transformation of
ferrite, only undissolved carbides remain in austenite.
720 740 760HV
1000
700 760HV

J J
Ac 3 Inhomogeneous
Homogeneous
Temperature in °C

900

Ac1e
DJ
Ac1b
800
DJ+M3C

D+M3C
700 2 3 4 5
0.1 1 10 10 10 10 10
Time in s

Figure 5.50: Schematic TTA diagram for isothermal austenitisation of the steel
C45; vA = 130 K/s.
Many technical steels still show carbides along with ferrite and austenite above the
Ac1b temperature, whereas the dissolution of carbides ends above Ac 3. This is
especially true for steels with ferrite and pearlite in their initial structure. Above
Ac1b, ferrite, austenite, and carbides are present, although the carbides no longer
show the form of lamellar pearlite. After 9 seconds, all of the carbides are in
solution so that from a metallographic point-of-view, the microstructure only
consists of ferrite and austenite. After 1600 seconds and surpassing the Ac3
temperature, the microstructure is completely transformed into austenite. After
quenching to room temperature, the steel has a hardness of 700 HV10. With
increasing holding time, the hardness increases to a maximum of 760 HV10 after
about 10500 seconds. Since the final martensitic hardness of this steel is only
determined by the carbon content, the various hardness values imply the difference
in carbon distribution in the austenite. This means for the example just discussed
that after nine seconds, although above Ac 3, austenite is present without visible
carbides. Metallographically, however, an inhomogeneous distribution of carbon
can be seen in the apparently “pure” austenite. Carbon enrichment can be found in
the spots where carbides had previously been present.
This leads to an especially transformation-prone austenite during subsequent
cooling. For example, a formation of soft spots can appear after hardening, because
local carbon enrichments reduce the martensite start-temperature to below room

300
5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams)

temperature. As a result, a microstructure forms that consists of hard martensite and


non-transformed, soft retained austenite (carbon enriched austenite).
Consequent is the demand for a sufficiently homogeneous austenite that is
independent of the austenitisation time and has the largest possible hardness. Apart
from this, austenitisation must be selected so that as the austenitisation times
increase, no undesired coarsening of the grains occurs. For this reason, the grain-
size lines are often graphed instead of the hardness lines (Figure 5.51).
ASTM 12 11 10 9 8 6 4 0 L(Pm)
1000
320
80
40
J J
Ac 3 20
Inhomogeneous Homogeneous
Temperature in °C

900
14

Ac1e
10
DJ
Ac1b 7
800
DJ+M3C 5

DM3C
700
0.1 1 10 10 2 103 104 10 5
Time in s

Figure 5.51: TTA diagram for isothermal austenitisation of the steel C45
(vA = 130 K/s). Lines of the same austenite grain size are additionally
graphed.
After austenitising at 820°C and holding for 105 seconds, fine austenite grains with
an ASTM grain size of 11 (approximately 7Pm) are obtained. In contrast, at
1000°C and 6.104s, the ASTM grain size 0 (approximately 320Pm) has already been
attained. On the other hand, after holding for the same time at 790°C, a
homogeneous austenite with the ASTM size 12 (approximately 5Pm) is just
reached. Thus at high austenitising temperatures and short holding times, as well as
at low temperatures and for long holding times, a homogeneous, fine-grained
austenitic microstructure can be achieved.
In contrast to the previously described hypoeutectoid steels, hypereutectoid steels
are generally austenitised in the austenite and carbide two-phase region, in order to
obtain a fraction of undissolved carbides alongside austenite. In this way, an
abrasion-reducing effect is obtained through the undissolved carbides, which is of
importance for steels that are predominately used as tool steels. Furthermore, a
decrease in hardness for homogeneous austenite is prevented.

301
5 Technical Heat Treatments

It can be seen from the isothermal TTA austenite grain-growth diagram of the
bearing steel 100Cr6 (Figure 5.52), that barely any grain growth appears in the
austenite-carbide mixed region. A coursing of the grains occurs only after the Accm
temperature has been exceeded due to the complete dissolution of the carbides.
After an austenitisation time of 103 seconds, the steel has its highest hardness at a
temperature of around 845°C. Higher or lower temperatures lead to lower hardness
values. This hardness maximum is typical for hypereutectoid steels.
The increase in hardness is connected to the increasing carbon content in
martensite, due to the dissolved carbides. At a high content of carbon in the
austenitic solid solution, the hardness decreases again due to the formation of
retained austenite. The hypoeutectoid and hypereutectoid steels differ
fundamentally from one another in this behavior.
For the technical application of these correlations, it is important to know that many
of the diagrams were determined using very small samples. In contrast to this,
actual workpieces do not show these short heating durations during technical
application, due to the penetration heating that is necessary. Therefore, in technical
applications, austenitisation in the first phase corresponds to continuous heating.
5.4.2 Austenitisation with continuous heating

TTA diagrams for steels heated continuously and at a constant heating rate are used
to show the effects on transformation processes of rapid heating due to, for
example, induction hardening, welding or short term surface layer hardening.
The transformation of a ferritic-pearlitic microstructure to austenite occurs after
nucleation, according to the law of nucleus growth. The nuclei develop when the
transformation temperature is exceeded, after a period of incubation at the grain
boundary or phase boundary. The initial microstructure is the primary determining
factor for the number of nuclei. The number of nuclei increases with increased
superheating or with an increased rate of heating. The continued formation of
austenite results from the growth of the nuclei into the surrounding ferrite, although
the carbon necessary for the transformation has to diffuse to the growth front.
Thereby, the carbon content is dependent on the transformation temperature, i.e. the
carbon content of the newly formed austenite decreases with an increasing heating
rate.

302
5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams)

Figure 5.52: Isothermal TTA diagram for the bearing steel 100Cr6, according to
Atlas zur Wärmebehandlung der Stähle.
The transformation behavior in a continuous TTA diagram is always considered for
each heating rate along a sketched line of the respective individual rate
(Figure 5.53).

303
5 Technical Heat Treatments

Figure 5.53: Continuous Time-Temperature-Austenitisation diagram for the steel


G15CrNi6, according to Atlas zur Wärmebehandlung der Stähle.

304
5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams)

It can be seen from the figure that carbide dissolution is greatly delayed in
comparison to equilibrium conditions. If, for example, a heating rate of 1 K/s is
considered, ferrite and pearlite are present until a temperature of 730°C. Ferrite
increasingly transforms into austenite until 750°C and is fully transformed at around
830°C (Ac3 temperature). At the same time, all carbides are in solution, but are not
yet evenly distributed throughout austenite (“inhomogeneous austenite”). Carbides
are homogeneously distributed only above 855°C (“homogeneous austenite”). A
heating rate of 1000 K/s causes a delay in transformation. Austenite formation
begins at 780°C, and ferrite is then completely transformed to austenite at 880°C.
However, the carbides go into solution so slowly at all heating rates that a
distinction between Aclb and Ac1e is no longer possible. Beginning at a critical
heating rate (in this case, beginning at around 1000 K/s), neither further
superheating, nor either the beginning or end of austenite transformation is possible.
The Ac1 and Ac3 temperatures are horizontal.
During superheating above Ac3, the entire carbon content is no longer required for
an (“inhomogeneous”) austenite formation. Residual carbides and irregularities in
the carbon content remain in the “inhomogeneous” austenite. With increasing
superheating, austenite formation occurs more rapidly, since both nucleation and
diffusion-dependent growth require less time. Therefore, a larger field of
“inhomogeneous” austenite forms with an increasing heating rate.
The influence of grain growth and martensite hardness can be depicted in a
continuous TTA diagram (Figure 5.54). An ideal rate can be determined
accordingly for heating, in order to acquire a sufficiently large number of
undissolved carbides, and with that to achieve the optimum quenching hardness.
An ideal initial hardness of 900 HVl and a grain size of 9 to 11 can be reached with
a heating rate of 100 K/s and an austenitising temperature of around 980°C.
Exceeding this temperature leads to superheating, which means that the carbon
content increases in the lattice, the martensite needles coarsen after quenching, and
the fraction of retained austenite increases. A lower austenitising temperature or a
shorter austenitisation period lead to an inhomogeneous microstructure. Thus, in
contrast to hypoeutectoid steels, for hypereutectoid steels a short or long
austenitisation period produces different austenitic states, associated with different
properties. In hypoeutectoid steels, hardly any difference is noticable in the
homogeneity above Ac3, since the carbide dissolution is already complete once the
Ac3 temperature is exceeded, and homogenisation proceeds relatively quickly.

305
5 Technical Heat Treatments

Figure 5.54: Continuous TTA quenching-hardness diagram for the bearing steel
100Cr6, according to Atlas zur Wärmebehandlung der Stähle.

306
5.4 Time-Temperature-Austenitisation Diagrams (TTA diagrams)

5.4.3 Influence of austenitisation

The macroscopic differences in the chemical composition of two melts of one steel
result in different types of austenitisation behavior. The characteristic differences
between individual steels in a TTA diagram are, however, usually greater than the
tolerance given in the chemical analysis. The alloying of carbide stabilising
elements delays carbide dissolution, through which the Ac 1 temperature is raised up
to the Ac3 temperature in extreme cases.
Equations for calculating the transformation temperature TT have been developed
using regression analysis of the Ac temperatures measured for many steels, and
have the following form:
TT a ˜ C  n0  n1 ˜ M1  n2 ˜ M 2  ...  ni ˜ M i (5.11)
Mi and C represent the respective mass-content of alloying elements and carbon,
while a and ni are constants. These constants are only applicable for the specific
analysis range of the steels from which they were originally determined.
The initial microstructure state and especially the size and distribution of carbides
influence the austenitisation behavior. Large carbides require more time to dissolve
in the already formed austenite than small carbides, which were formed during the
tempering of austenite. Since carbon, which is predominantly bound in carbides,
must diffuse into austenite, time is necessary. Therefore, a carbide dissolution of
larger carbides in the initial state requires longer holding times during isothermal
austenitisation, or analogously, higher temperatures during continuous heating.
If an initial ferritic-pearlitic microstructure is compared to an initial microstructure
of martensite and if heated at a rate of 30 K/s, then the martensitic initial state is
homogeneously austenitised at 830°C, and the ferritic-pearlitic microstructure only
at 930°C (Figure 5.55). When heating below the Ac1 temperature (tempering), fine,
evenly distributed carbides precipitate from martensite, which dissolve significantly
more rapidly during austenitisation than the strongly carbon-enriched carbides of
pearlite.

307
5 Technical Heat Treatments

Figure 5.55: Influence of the initial state on austenite formation during continuous
heating for the steel C53 (Martensite as initial microstructure).

308
5.5 Transformation Diagrams

5.5 Transformation Diagrams

Information regarding the possibilities for heat treatments of steels can be found in
transformation diagrams, which describe the transformation behavior during
cooling. Information about the microstructure formation and hardness to be
expected are also possible, as long as the experimental and actual processing
conditions coincide with equal hardenabilities.
The comparability of the diagrams is ensured by SEP 1680, in which, among others,
specifications for sample selection and preparation, experimental procedure and
evaluation, and the presentation of test results can be found.
To determine a TTT diagram with isothermal transformation, samples are quickly,
“infinitely” cooled after austenitisation to the transformation temperature TT < Ar1.
After various holding times and temperature homogenisation at the transformation
temperature, the amount of transformed microstructure can be metallographically
determined. If the isothermal transformation behavior is determined by means of
dilatometric tests, then a sigmoidal curve appears (Figure 5.45). The procedure is
based on the change in specific volume in the cubic system through phase
transformation, and the resulting change in length. For CCT diagrams, samples are
either linearly or exponentially cooled from the austenitising temperature. In this
case, continuous cooling agrees well with Newton’s Law. However, deviations do
arise due to transformation-dependent heat disturbances. This means that the heat
that is set free during transformation continuously interferes with the cooling.
Abrupt cooling, for example, with oil or water, also deviates from the Newtonian
course.
With the help of dilatometric measurements, the transformation boundaries can also
be determined for various continuous coolings (Figure 5.46).
In order to develop a complete transformation diagram, 9 to 12 cooling curves are
necessary. The transformation points are plotted in a T/log t diagram. By
connecting the points measured at the beginning and the end of austenite
transformation for each of the respective microstructural products, the fields for the
various microstructural types are created. In doing so, the beginning of the
proeutectoid ferrite or carbide precipitation, pearlite, bainite, and martensite
formation, and the end of the pearlite and bainite transformations are determined.
5.5.1 Isothermal transformation

An isothermal TTT diagram is interpreted along its isothermal lines. The


transformation of the previously austenitised steels occurs during cooling through
nucleation and nucleus growth processes. Therefore, it is a time-dependent process.
At each temperature, the transformation rate is calculated as the product between
nucleation rate and nucleus growth. This relationship can be gathered from the left

309
5 Technical Heat Treatments

hand diagram in Figure 5.56. The maximal transformation rate corresponds,


according to the right hand diagram in Figure 5.56, to a minimum in the time
needed for transformation, which behaves reciprocally to the transformation rate.
This C-shaped curve is typical for many transformation processes. Therefore, all
diagrams show the characteristic “nose-shaped” or “C-shaped” transformation
boundary.

Equilibrium transformation temperature Equilibrium transformation temperature

Growth
rate

Temperature
Temperature

Time for 50%


transformation
Transformation
rate
Nucleation
rate
Time
Figure 5.56: Derivation of growth rate from the nucleation rate and the
transformation rate (left); sketch of the „C-shaped“ curve of a typical
transformation process (right).
Thus, nucleation is significantly driven by supercooling (oversaturation), which can
be understood as the driving force for transformation. Nucleus growth is primarily
driven by diffusion. The diffusion coefficient D is temperature-dependent, and by
graphing it logarithmically over the reciprocal of the temperature, the relationship
can be represented linearly in an Arrhenius graph (Equation 5.12).
ln D = ln D0 - Q/RT (5.12)
where D = diffusion coefficient, D0 = diffusion coefficient at t=0, Q = activation
energy, R = general gas constant, and T = temperature.
The closer the transformation temperature is to the equilibrium temperature, the
more the solubility in the homogenous austenite crystal increases, and the smaller
the supercooling (oversaturation) becomes, and with it the driving force for
nucleation. The critical nucleation energy becomes very high and infinitely large at
the equilibrium temperature TE. In contrast, the nucleation energy that must be
supplied decreases with increasing supercooling, decreasing transformation
temperature, and increasing tranformation energy.
If a material is “infinitely” cooled to the transformation temperature TT=Tl<TE from
homogeneous austenite and held isothermally, then the driving force for
transformation is very small (Figure 5.57).

310
5.5 Transformation Diagrams

Figure 5.57: Transformation curve for an isothermal test.


The first nuclei to form require a great deal of time. At high temperatures, the
atoms are quite mobile so that the few new nuclei grow very quickly. As a result, a
coarse grained transformation phase develops, that consists of few grains. At a
lower transformation temperature T2, the driving force for nucleation, and with it
the growth rate of the nuclei, is significantly larger than at the temperature Tl. The
transformation occurs more rapidly, although the atomic mobility (diffusion) is
limited at decreased temperatures. At very low transformation temperatures T 3, the
great increase in driving force and decrease in the critical energy of nucleation are
over-compensated by an inadequate possibility for diffusion. This leads to a delay
in the transformation of austenite. If the normalised volume that is transformed
during the diffusion-controlled transformation for ferrite, pearlite, and bainite
formation (from 0% to 100% or from 0 to 1) is graphed over time or the logarithm
of time, then the curve is sigmoidal (Figure 5.58). The characteristic curve that is
valid for one respective transformation temperature T = T T shows three regions:
x A beginning region, I, in which no transformation has yet occurred
(transformed amount W=0); this region is also identified as an incubation
period. It represents the period of time necessary for the formation of nuclei
that are capable of growing under experimental conditions. According to
convention, the end of the incubation phase is the point in time when 1% of
the product phase has developed.

311
5 Technical Heat Treatments

x The middle curve section, II, describes a great increase of the product phase.
Along with the process of nucleation, rapid grain growth occurs.
x As of the inflection point in the curve, the transformation rate dx/dt in area III
becomes increasingly smaller, and the curve asymptotically approaches the
value 1 (100%). The end point of the transformation is determined by the
point at which 99% of the product-microstructure has been formed. The
inflection point and the following decay of the reaction are
metallographically interpreted as such, that the product-phase regions
increasingly constrict one another’s growth through “collision”, until the
entire initial microstructure is finally consumed. The initial point (1%) and
end point (99%) of transformation at the temperature T represent two
distinctive points in this type of microstructural volume-time graph.

0.99

I II III

W
Inflection point

0.01
log t
Figure 5.58: Sketch of the JMAK function W(t) = 1 - exp (-btn) with the three
characteristic phase transformation regions.
Additional data on the beginning and end of transformation is given through the
construction of further microstructural volume-time graphs for different
transformation temperatures. All data is entered into a temperature/logarithm-
transformation time diagram, and the times for the beginning and end of each
transformation are connected with one another (Figure 5.57). In this diagram, the
characteristics of the C-curve are made clear:
x Long incubation periods at high transformation temperatures (good diffusion,
low number of nuclei),
x Long incubation periods at low transformation temperatures (hindered
diffusion, large number of nuclei),
x Minimal incubation periods at middle transformation temperatures (relatively
good diffusion, relatively high number of nuclei).

312
5.5 Transformation Diagrams

The mathematical description of the sigmoidal curve for the normalised product
microstructure W, as dependent on the reaction time, can be described using the
Johnson-Mehl-Equation. It is valid for the growth of spherical pearlite, and is
derived under the assumption of the validity of the “extended product-phase
volume” or the “extended product-phase boundary surface”. This concept requires
that all fields can grow freely, without moving phase boundaries colliding, and that
continuous nucleation can take place throughout the entire volume of the sample,
including the fields that are already transformed.
§ 1 ·
W 1  exp¨  ˜ S ˜ K ˜ G 3 ˜ t 4 ¸ (5.13)
© 3 ¹
where W = volume fraction transformed, K = rate of nucleation in nuclei/volume *
time, G = growth rate in distance/time, and t = time.
A more general form to describe the sigmoidal curve of the product microstructure
is the JMAK Equation (Johnson-Mehl-Avrami-Kolmogorov), which can be
generally applied for all diffusion-controlled phase transformations:
­ § t ·n ½
W (t ) 1  exp ® ¨ ¸ ¾ (5.14)
¯ ©W ¹ ¿
where W = precipitated volume fraction, τ = temperature-dependent time constant, t
= time, and n = temperature-dependent exponent.

One method to mathematically describe the lattice shearing mechanism of


martensite transformation is that from Koistinen and Marburger:
W 1  exp  D ˜ M S  T (5.15)
where W = fraction transformed, Ms = martensite start temperature, T = temperature,
and D material constant.
If all of the transformation regions are graphed in one diagram, then overlaps occur
at some of the transformation temperatures (Figure 5.59). In order to improve the
overview of the phase diagrams, and to reduce the experimental complexity,
simplifications were agreed upon, that take into consideration the different
influences on transformation.
To reduce measurement expenses, the end of the proeutectoid formation of ferrite,
carbide, bainite or martensite is not entered into the TTT diagram, since this
demands great experimental effort.
Moreover, the beginning of a microstructural transformation is first given, when
1 vol.-% of this type of microstructure, based on the total volume, has been formed.
This boundary cannot be used for the beginning of the proeutectoid carbide
precipitation, because the recalescence cannot be sufficiently detected. In this case,

313
5 Technical Heat Treatments

the starting point is chosen, after a metallographic examination reveals the first
visible precipitations. After the beginning of the pearlite formation, no other
transformations begin. This explains the omitted lines for 1 vol.-% of proeutectoid
precipitates and 1 vol.-% bainite, as long as they lie sequentially later than the line
for 1 vol.-% pearlite, at a certain transformation temperature. The end of the
diffusion-controlled transformation is given at 99 vol.-% transformed austenite.
The beginning of the martensite transformation is characterised by the martensite
start temperature (Ms). The Ms-temperature is usually determined by the
measurement of physical parameters, which are especially influenced by changes in
the lattice structure. Taking into consideration the above mentioned limitations, a
schematic, isothermal TTT diagram can be established. Also plotted are the
relevant equilibrium temperatures for proeutectoid ferrite formation (TEV), pearlite
and bainite formation (TEP, TEB), and martensite formation (TEM), although the
equilibrium temperature of bainite TEB generally cannot be determined
experimentally (Figure 5.60).

314
5.5 Transformation Diagrams

T EV
A+V
TEP
A A+V

A+P V+P
TEB
Temperature

A+B B

Ms
20
TEM
60
80% M

log t
Figure 5.59: Isothermal transformation diagram of a steel with 0.45% C, without
taking into consideration the interactive influences of transformations
(A = austenite, V = proeutectoid precipitations, P = pearlite,
B = bainite, M = martensite, Ms = start of the martensite formation,
and TE = equilibrium temperature).

315
5 Technical Heat Treatments

Figure 5.60: Isothermal transformation diagram of a steel with 0.45% C, under


consideration of the reciprocal influences on transformations.
Figure 5.61 shows an isothermal TTT diagram of the quenched and tempered steel
34Cr4 for an austenitisation at 850°C for 5 minutes. The beginning of ferrite
formation asymptotically approaches the Ac 3 temperature. The beginning and the
end of pearlite formation asymptotically approach the Ac 1 temperature. The
equilibrium temperatures are thus only reached after an infinitely long holding
period. If the steel is quenched from the austenitising temperature of 850°C to the
transformation temperature of 670°C, then the austenite transformation begins after
5 seconds with a proeutectoid ferrite precipitation on the grain boundaries of the
austenite. Pearlite formation begins after 18 seconds, without the certainty that the
ferrite formation has come to an end. After about 130 seconds, the austenite
transformation is complete, and at room temperature, the microstructure consists of
70 vol.-% pearlite and 30 vol.-% ferrite, with a hardness of 240 HV10. Since
austenite is transformed into the thermodynamically stable ferrite and the metastable
pearlite after a period of 130 seconds, no further transformations can occur after that
point. If a sample of this steel is cooled to 300°C, 50 vol.-% martensite forms
abruptly. The amount of martensite does not change over time. After 6 seconds the

316
5.5 Transformation Diagrams

bainite formation begins and ends after 220 seconds. The austenite transforms into
50 vol.-% martensite and 50 vol.-% bainite.

Figure 5.61: Isothermal Time-Temperature-Transformation diagram for the steel


34Cr4, according to Atlas zur Wärmebehandlung der Stähle.
During nonequilibrium cooling of hypoeutectoid steels, which show a ferritic-
pearlitic microstructure in equilibrium, a lower percentage of proeutectoid ferrite is
to be found throughout the microstructure as supercooling and cooling rate increase.
In this manner, for example, a steel whose microstructure consists near-equilibrium
of 80% pearlite and 20% ferrite, transforms to 100% pearlite by increasing the
supercooling. An example of the practical application of this is a patenting
treatment. This behavior can be explained by the fact that the carbon rejected on the
ferrite/austenite phase boundary during the formation of proeutectoid ferrite can
only diffuse with increasing difficulty into the metastable ferrite as supercooling
increases, and thus piles up on the phase boundary. Directly after the critical carbon
content is reached, the formation of ferrite stops because the conditions for the
nucleation of pearlite are fulfilled, although the eutectoid concentration in the
remaining austenite has not yet been reached. Thus, an iron-rich pearlite is created

317
5 Technical Heat Treatments

at the cost of proeutectoid ferrite. The ratio between the width of the ferrite
lamellae and the width of the cementite lamellae increases as the cooling rate and
supercooling increase (isothermal process). Altogether, the lamellar spacing of
pearlite becomes smaller as the temperature decreases and cooling rate increases,
due to a greater supercooling. The lamellae themselves become finer, and the size
of the individual pearlite grains becomes smaller. As a result, the strength and
hardness of pearlite increases. The upper bainite shows a coarser ferrite-carbide
distribution than the lower bainite, which develops at lower temperatures. As
supercooling increases, the bainitic microstructure also becomes finer, and the
strength increases. The martensitic microstructure reaches the maximal hardness
and strength values of all steels.
If the hypereutectoid bearing steel 100Cr6 is quenched to 600°C after
austenitisation at 1050°C for 15 minutes in the homogeneous austenite, then
proeutectoid carbides begin to precipitate out as a closed film on the austenite grain
boundaries after 15 seconds, due to the favorable situation for nucleation (Figure
5.62). Pearlite formation begins after 25 seconds and is complete after 110 seconds.
A film on the grain boundary made of hard and brittle carbides forms in a purely
pearlitic microstructure, which has a hardness of 534 HV10. During heat
treatments, hypereutectoid steels are usually austenitised in the austenite and
carbide two-phase region. After austenitisation at 860°C for 15 minutes, the
microstructure consists of undissolved carbides and austenite. Instead of the one
phase region A (austenite), a two-phase region A+C is graphed. Since the dissolved
carbon nucleates on the undissolved carbides during subsequent transformation, the
beginning of carbide precipitation can only be measured with great difficulty.
Therefore, in this case, the designation of proeutectoid carbide precipitation is
omitted.
One difference to hypoeutectoid steels is the position of the martensite start
temperature, which affects the final hardness. After all carbon has gone into
solution at an austenitising temperature of 1050°C, the Ms temperature is lower than
after an austenitising temperature of 860°C. After austenitisation in the
homogeneous austenite, the martensite formation is not complete at room
temperature (50 vol.-% martensite). The hardness is significantly lower at room
temperature, due to an increased retained austenite content, than after austenitisation
in the austenite and carbide two-phase region.

318
5.5 Transformation Diagrams

Figure 5.62: Isothermal TTT diagram of the steel 100Cr6, according to Atlas zur
Wärmebehandlung der Stähle.

319
5 Technical Heat Treatments

5.5.2 Continuous transformation

The transformation mechanisms during isothermal and continuous temperature-time


progression are identical. However, the percentage of microstructural
transformation depends upon how long samples remain in a certain temperature
range, according to continuous cooling. The testing method to determine CCT
diagrams has already been described.
According to the approach for isothermal diagrams, the lines for 1%, in relation to
the total volume of the microstructure formed, and for 99% transformed austenite
are graphed in the diagram (Figure 5.63).

Figure 5.63: CCT diagram for the steel 34Cr4, according to Atlas zur
Wärmebehandlung der Stähle.
The lines showing the end of ferrite formation are not determined. The same is true
for the end of pearlite and bainite formation, if at lower temperatures the

320
5.5 Transformation Diagrams

transformation ends with bainite or martensite formation. The completion of


martensite formation can often only be approximated. In many cases the percentage
of retained austenite present at room temperature is given to designate the
transformation state for shorter cooling periods. A CCT diagram can only be
interpreted along the cooling curve. The course of the transformations as well as
the microstructural ratios present at room temperature and the final hardness values
are only valid for one specific cooling curve.
Often in the course of continuous cooling during pearlite formation, not only a
delay in cooling, but also a rise in temperature can be observed. This effect, also
known as recalescence or heat tone, results from the heat of transformation released
and causes the sample to heat up.
After a sample (the exponential cooling curve designated with an X in Figure 5.63)
is austenitised at 850°C for 8 minutes, the transformation of austenite to ferrite
begins at 690°C. Pearlite formation begins at 650°C and finishes at 600°C. After
passing through a transformationally inert region, bainite transformation begins at
550°C, and martensite formation at 300°C. For completion, each of the
microstructural content ratios determined at room temperature is given on the
cooling curve, although, it was not determined when the given microstructural
contents came to exist. Finally, the hardness value (29 HRC), also measured at
room temperature, is given. For hypoeutectoid steels, the Ms-temperature always
appears to decrease as the cooling rate decreases. This phenomenon is related to the
precipitation of proeutectoid ferrite and the ferrite within upper bainite. In both
cases, the carbon content of the surrounding austenite increases, due to the low
carbon solubility of proeutectoid and bainitic ferrite. This causes the Ms-
temperature to sink.
Exponential cooling curves are often identified by the t 8/5-time, which indicates the
duration of cooling from 800 to 500°C or from Ac3 to 500°C. The cooling
parameter O = t8/5/100 can be used instead of the t8/5-cooling time. This makes a
direct comparison of the graphs of transformation from different austenitising
temperatures possible by superimposing the graphs. It can be defined for any
cooling curve. Hence, the curve X corresponds to a cooling time of t 8/5 = 150s.
In Figure 5.64 continuous cooling phase transformation diagrams for the
hypereutectoid steel 100Cr6 are given for two different austenitising conditions.
After austenitisation in the homogeneous austenite at 1050°C for 15 minutes
followed by continuous cooling (cooling curve X) from Ac1e to 500K in a t8/5-time
of 72 seconds, austenite transforms into martensite after the precipitation of
proeutectoid carbides (Figure 5.64). Because of the high content of dissolved
carbon in austenite, the Mf-temperature is far below room temperature so that a high
percentage of non-transformed retained austenite (abbreviated in the figure with
RA) remains. The proeutectoid carbide precipitation that occurs before the

321
5 Technical Heat Treatments

martensite formation leads to a carbon depletion in austenite and therefore, to a rise


in the Ms-temperature.

322
5.5 Transformation Diagrams

Figure 5.64: CCT diagrams of the steel 100Cr6, according to Atlas zur
Wärmebehandlung der Stähle.

323
5 Technical Heat Treatments

In contrast, the transformation diagram for an austenitisation at 860° in the two-


phase region of austenite and carbide only shows a pure martensitic phase
transformation of austenite below a t 8/5-cooling time of 15 seconds. The rise in the
Ms-temperature to 250°C as compared to 130°C for higher austenitising
temperatures is a result of the lower content of dissolved carbon in the austenite.
Furthermore, it becomes clear that a preliminary bainite transformation causes an
enrichment of carbon in austenite, and therefore, decreases the Ms-temperature. The
lowest content of retained austenite results from the highest Ms-temperature.
5.5.3 Other possibilities for the representation of transformation behavior

The change in the microstructure phase distribution is often graphed in a CMT


(cooling time-microstructure-transformation) diagram. This graph is a function of
the cooling rate, which can be represented by the cooling time between 800°C and
500°C or between Ac3 and 500°C (Figure 5.65). This kind of graph makes it
possible to characterise the microstructure at room temperature with respect to its
components.
To further simplify the transformation behavior, a “critical cooling time” is often
given. The upper critical cooling rate Km marks the longest cooling time and the
slowest cooling rate, through which only martensite forms. The lower critical
cooling rate represents the shortest holding time and the most rapid cooling rate,
through which 1% martensite is formed. Generally, the lower critical cooling rate
is identical to the cooling time Kp, which leads to a complete transformation into
pearlite. Kf is the lowest duration of cooling in order for hypoeutectoid steels to
form 1% proeutectoid ferrite. Analogous to the hypereutectoid steels, Kk represents
the formation of 1% proeutectoid carbide.
In CTT (cooling time-temperature-transformation) diagrams, the transformation
temperature is graphed over the t8/5-time. Here, the graph is no longer interpreted
along the cooling curve. Instead, the development of the microstructure is read
perpendicular to the x-axis. As an example, the CTT diagram of the steel S355GT
is given in Figure 5.66. Further special graphs of transformation behavior are the
DTTT (deformation TTT) and WTTT (weld TTT) diagrams, which do not differ in
their appearance from the common TTT diagrams. The difference lies in the
previous history of the material. In DTTT diagrams, the austenite transformation
precedes the deformation in the austenite. This results in a change in the
transformation behavior. Similarly, DCTT (deformation CTT) and DCMT
(deformation-cooling time-microstructure-transformation) diagrams can also be
constructed. The type of pre-deformation must be shown in the diagram when a
preceding deformation has occurred. WTTT diagrams are specifically important for
welding technicians. Here, a welding-adjusted temperature cycle precedes cooling
from austenite (rapid heating to a high temperature, short holding time in austenite).

324
5.5 Transformation Diagrams

900

Ac 3
800

38 Ac 1
34
Temperature in °C

700 62
X 8 20 66
A 4 82
600 F P
50
5

500
B

400
Ms
16
37 78
300 40
M

200

100
HV 10 720 560 440 330 350 285 235 225
0
1 10 10 2 10 3 104 10 5

100 1000
Microstructural fraction

Hardness in HV 10
M P
B 800
80
in Vol.-%

60 600
HV

40 400
F

20 200
Km KIf K`p Kp
0 0
1 10 10 2 103 10 4 10 5
Cooling time from Ac3 to 500°C in s
Figure 5.65: CMT diagram for continuous cooling of the steel 41Cr4. Initial
microstructure: 25 Vol.-% Ferrite + 75 Vol.-% Pearlite,
Austenitisation: 840°C, 15 min (top); microstructural fractions and
hardness dependent on the cooling time from Ac3 to 500°C (Km = 7.5s;
Kf = 43s, Kp’= 80s; Kp = 280s) (bottom).

325
5 Technical Heat Treatments

1000

Ac3
900

800
Ac1e
Ac1b
700
F 80 80
75 80
50 20
20
600 P 20
Temperature in °C

20

500
B
45

90
400
M 10

300

200

100
414 244 191 178 159 151 146
0
0 2 3 4 1 2 3 4 2 2 3 4 3 2 3 4 4
10 10 10 10 10
t8/5 - time in s
Figure 5.66: CTT diagram of the steel S355GT, according to SEP 1680.
5.5.4 Influences on the transformation behavior

The austenite transformation and its kinetics during cooling are decisively
influenced by the process parameters and the preliminary processing steps.
Significant influencing factors are:
x Chemical composition,
x Austenitisation conditions, and
x Deformation before cooling.
The influence of the initial state on the transformation behavior is summarised in
Figure 5.67. The precipitated MeX particles, which lead to more diffusion nuclei,
accelerate the diffusion-controlled phase transformations. However, a high
austenitising temperature (TJ) causes a reduction in nuclei through increased grain
growth, leading to a delay in the transformation. Pre-deformation, in turn, increases
the number of sites possible for nucleation and consequently, usually accelerates the
phase transformation.

326
5.5 Transformation Diagrams

Figure 5.67: Influence of the initial state or a previous austenite deformation on the
JoDtransformation (schematic).
Figure 5.68 shows the influence of carbon on the transformation behavior.
Hypoeutectoid steels show a decreasing M s-temperature at relatively low cooling
rates. This is due to the low solubility of carbon in ferrite (max. 0.02%). During
the proeutectoid ferrite formation, carbon diffuses into the retained austenite and is
concentrated there, which decreases the Ms-temperature. Hypereutectoid steels
show a rise in the Ms-temperature after cementite has precipitated out and, through
this, the austenite is depleted of carbon. Carbon shifts the diffusion-controlled
transformation so that it requires more time. However, a carbon content above
approximately 0.9% (in non-alloyed steels) leads to an accelerated transformation
behavior. This is due to the proeutectoid cementite grains, which serve as
nucleation sites for pearlite formation. The bainite and pearlite transformation is
greatly delayed up to the eutectoid content, as the carbon content increases. In
hypereutectoid steels, the proeutectoid carbide formation that occurs with increasing
carbon concentrations leads to shorter transformation times. The influence of
alloying elements on the transformation behavior is shown schematically in Figure
5.69. Manganese, nickel, molybdenum, chromium, and vanadium delay the
transformation to pearlite and cause a distinct separation of pearlite from bainite.

327
5 Technical Heat Treatments

The Ms-temperature is also decreased. On the other hand, bainite formation is


accelerated by carbon, chromium, manganese, nickel, and vanadium.

T
Ac 3
Ac 1
F
A P

B
Ms
Hypoeutectoid
steel

lg t

Ac 3=Ac1

P
A
B
Ms Eutectoid
steel

lg t

T Ac cm

Ac 1

A P

B Hypereutectoid
Ms steel

lg t

Figure 5.68: Influence of austenitising temperature and chemical composition on


the transformation behavior of steels.
Because of these influences on transformation, TTT diagrams are, strictly speaking,
only valid for the individual melt investigated. Therefore, in segregated steels,
there are localised differences in the transformation behavior. The segregated zones
show extremely different distributions of alloying elements, and accordingly, act as
different charges with regard to their transformation behavior.

328
5.5 Transformation Diagrams

Figure 5.69: Influence of alloying elements and boundary conditions of heat


treatments on the transformation behavior, according to the Stahl-
Informations-Zentrum Merkblatt 450.
The transformation behavior is also influenced by the austenitisation conditions
(temperature and holding time). Increasing the austenitising temperature or time
results in a coarsening of the grains. Because of the lack of nucleation spots, this
leads to a delayed austenite transformation (Figure 5.70). Transformation into
pearlite, especially, starts at the grain boundaries with low supercooling and is,
therefore, severely decelerated by coarse grains. The Ms-temperature and
martensite hardness are largely independent of the austenitisation conditions, if
quenched from homogeneous austenite.
A further factor influencing the phase transformation is a deformation occurring
before or during the phase transformation. The number of nucleation sites available
for the subsequent transformation is increased through a previous deformation. If
the final deformation temperature is higher than the recrystallisation temperature,
then recrystallisation occurs either during or after deformation. As a result, the
austenite grains are distinctly smaller in comparison to the grain size before hot
rolling. However, the defect density is greatly reduced through recrystallisation. If
the final deformation takes place below the recrystallisation temperature, then a
non-recrystallised, deformed austenite with an increased defect density is present
before transformation.

329
5 Technical Heat Treatments

Figure 5.70: Influence of austenitising temperature on the transformation behavior


of steels.
During diffusion-controlled transformation processes, nucleation usually begins on
the grain boundaries and results in a fine grained austenite, and hence, leads to an

330
5.5 Transformation Diagrams

acceleration of the transformation. Through hot deformation in the region of the


non-recrystallised austenite, the effect of deformation on the microstructure
transformation is magnified. The effective grain boundary area per unit volume is
almost independent of the austenite grain size and is increased mainly by
deformation (Figure 5.71).
Degree of deformation
0 0.2 0.4 0.6 0.8 1.0 1.2
8

7
S V, mm-1
Austenite grain size, G(ASTM)

6 80
70
5 60
50
4
40
3 30

2 20

1 10

0
-1
0 10 20 30 40 50 60 70
Reduction in height in %
Figure 5.71: Change of grain boundary area (SV) through deformation below the
recrystallisation temperature.
Starting from an undeformed state, the effective grain boundary area is increased
through grain elongation (Figure 5.72a). Through the increased defect density on
the grain boundaries of the deformed grains (b), the nucleation rate per unit area of
the grain boundaries increases. In doing so, the nucleation area is also increased
through the formation of deformation bands inside the grains (c). This leads to an
acceleration of the diffusion-controlled ferrite and pearlite transformations, since
nucleation and nucleus growth can progress more quickly due to the introduced
defect sites. The effective grain boundary area per unit volume SV is often used to
describe the austenite grain boundary area and the lattice defects as nucleation sites
for the diffusion-controlled phase transformation. The effective grain boundary
area increases with deformation, although the increase is the result of both grain
elongation and the appearance of deformation bands.

331
5 Technical Heat Treatments

Differing conclusions are made in the literature regarding the Ms-temperature.


Increased Ms-temperatures are a result of the formation of preformed nuclei for
martensite formation. Lower Ms-temperatures are attributed to a hindered
cooperative shear movement through the increased concentration of defect areas in
deformed austenite (Table 5.5).

Figure 5.72: Nucleation sites for diffusion-controlled phase transformations:


a) nucleation in undeformed austenite, b) increased nucleation rate on
grain boundaries and c) additional nucleation on deformation bands.
Deformation parameter Type of metallurgical mechanism Effect on Ms
Carbide pre-precipitation on the
High deformation temperature nucleation spots for martensite Increased

Low deformation temperature, Dislocation configuration, localised


Increased
small degree of deformation alloying element depletion
Low deformation temperature, Presence of bainite, hindrance of
Decreased
high degree of deformation atomic shear movements

Table 5.5: Influence of different deformation parameters on the Ms-temperature.

332
5.6 Further Readings

5.6 Further Readings

Benninghoff, H.:
Wärmebehandlung der Bau- und Werkzeugstähle
3. Auflage, BAZ Buchverlag, Basel, 1978

Béranger, G.; Henry, G.; Sanz, G.:


The Book of Steel
Intercept, Andover, 1996

Berns, H.; Siekmann, G. u. Wiesenecker, I.:


Gefüge und Bruch vergüteter Federstähle – Mikrogefüge und
Bruchmorphologie
Draht 35 (1984) Nr. 5, S. 247/52

Bleck, W.; Meyer, L.; Kaspar, R.:


Thermomechanisches Walzen von Warmband aus modernen
mikrolegierten Baustählen
Stahl und Eisen 111 (1991) Nr. 4, S. 51/56

Buchmayr, B.; Hirschmanner, F.; Aigmüller, G.:


Mikrostrukturelle Optimierung des Warmwalzens

Caspari, R.; Gulden, H.; Krieger, K.; Lepper, D.; Lübben, A. u.a.:
Errechnung der Härtbarkeit im Stirnabschreckversuch bei Einsatz- und
Vergütungsstählen
Stahl (1992) Nr. 2, S. 31/35

Cuddy, L. J:
Fundamentals of the controlled rolling process
The Metallurgical Society of AIME, Warrendale (PA),1986, S. 235/243

Cuddy, L.J
Thermomechanical Processing of Microalloyed Austenite
The Metallurgical Society of AIME, Warrendale (PA), 1982, S. 129/140

333
5 Technical Heat Treatments

Deardo, A.J.; Gray, J.M.; Meyer, L.:


Fundamental Metallurgy of Niobium in Steel
Niobium, H. Stuart, The Metallurgical Society of AIME, 1989

DIN EN 10052
Begriffe der Wärmebehandlung von Eisenwerkstoffen
Januar 1994

DIN EN 10083-1
Vergütungsstähle, Teil 1: Technische Lieferbedingungen für Edelstähle
August 2003

DIN EN 10084
Einsatzstähle, Technische Lieferbedingungen
Juni 1998

DIN EN ISO 18265


Metallische Werkstoffe – Umwertung von Härtewerten
Februar 2004

DIN EN ISO 642


Stahl – Stirnabschreckversuch (Jominy- Versuch)
Januar 2000

DIN EN 10020
Begriffsbestimmungen für die Einteilung von Stählen
Juli 2000

Eckstein, H.J.:
Technologie der Wärmebehandlung von Stahl
2. Auflage, VEB Deutscher Verlag für Grundstoffindustrie, Leipzig, 1988

Gladman, T.:
The Physical Metallurgy of Microalloyed Steels

334
5.6 Further Readings

The Institute of Materials, London, 2002, ISBN: 1902653815

Grosch, J.:
Grundlagen der technischen Wärmebehandlung von Stahl
Werkstofftechnische Verlagsgesellschaft, Karlsruhe, 1981

Gottstein, G.:
Physikalische Grundlagen der Materialkunde
3. Auflage, Springer Verlag, Berlin, Heidelberg, New York, Tokyo, 2007

Gray, J.M.; DeArdo, A.J.:


Austenite conditioning alternatives for microalloyed steel products
HSLA Steels: Metallurgy and Applications
ASM Int., Metals Park, OH, 1986, S. 83/96

Grossmann, M.A.:
Elements of hardenability
Stahl und Eisen 111 (1991) Nr. 7, S. 103/10

Haasen, P.:
Physikalische Metallkunde
3. Auflage, Springer Verlag, Berlin, Heidelberg, New York, 1994

Haberling, E. u. Rasche, K.:


Technologische Eigenschaften vergütbarer Warmarbeitsstähle
Thyssen Edelst. Techn. Berichte 9 (1983) Nr. 2, S. 111/20

Heisterkamp, F. u. Meyer, L.:


Mechanische Eigenschaften perlitarmer Baustähle
Thyssen Technische Berichte 44-65 (1971)

Hensger, K.-E.:
Grundlagen der HTMB von Stahl
Neue Hütte 22 (1977) Nr. 9, S. 490/97

335
6 Thermomechanical Treatment

6 Thermomechanical Treatment
The mechanical-technological properties (strength, toughness, corrosion resistance)
of metallic materials are determined by the type and constitution of their
microstructure. The previously described heat treatments, and their resulting
microstructural changes often lead to only a one-sided improvement in the
mechanical-technological properties. Generally, an increase in strength means a
decrease in toughness or a deterioration in corrosion resistance. Moreover, some of
the conventional heat treatments, such as soft annealing or diffusion annealing, are
very time consuming and require a lot of energy, and therefore, expensive.
The process of thermomechanical treatment (TMT) can be used in order to find the
most cost-effective way to a desired microstructure, whose mechanical-
technological values should be the same, if not better, as after a traditional heat
treatment. Practically every hot deformation can be understood as TMT. Generally,
however, a process is only considered a TMT when the temperature and
deformation conditions are consciously manipulated to give the final product
material properties that are not attainable through conventional manufacturing.
With the help of TMT, many known metallurgical mechanisms to improve strength
can be used simultaneously (Figure 6.1). In contrast to the conventional production
methods, toughness can also be increased. The improvements made in the
microstructure and the properties of the steel after applying a TMT can be attributed
to the interaction between the deformation and alloy-dependent structure of the
austenitic structure. The improvements can also be attributed to the transformation
process during subsequent cooling. Microalloying elements (MAE) play a special
role here.
700
Dislocation hardening (transformation hardening)
by TMT+Nb, Ti
600
Yield strength in MPa

Age hardening by TMT+Nb, Ti


500

400
Grain refining by TMT+Nb, Ti
300
Solid solution hardening
200

100 Base strength, soft steel

0
Figure 6.1: Contribution of different strength-improving mechanisms to increase
the yield point of a thermomechanically treated steel.

336
6.1 Terminology

6.1 Terminology
By examining hot rolling, three typical, controlled hot deformation processes can be
explained by metallurgical principles. During conventional rolling, the final
deformation is at a relatively high temperature in austenite in order to acquire the
necessary reduction in thickness with a small amount of force. The disadvantage is
that relatively coarse austenite grains could form. Three temperature ranges are to
be taken into account when considering the effectiveness of normalising and
thermomechanical deformation (Figure 6.2), where TNR (TNR: Temperature no
recrystallisation, recrystallisation stop temperature) is the temperature limit, below
which no further recrystallisation takes place:
x The range in which austenite recrystallises (T>TNR),
x The range in which austenite no longer recrystallises (TNR>T>A3), and
x The range of the D→J transformation (A3>T>A1).
The final rolling temperature is just above the recrystallisation stop temperature of
austenite for normalised rolling. Microalloying is not always necessary, but it is
used very often for different reasons. A fine, polygonal austenitic structure forms,
which leads, after phase transformation is complete, to a ferritic microstructure that
has the same quality as a normalised microstructure.
Principally, the large number of thermomechanical treatments can be divided into
three groups. The groups are divided according to the point in time when hot
deformation takes place with respect to the phase transformation, whereby the final
deformation occurs in either the J, (J+D), or D-region.
The shift in the transformation to lower temperatures through manganese, and the
enlargement of the non-recrystallised austenite region through microalloying
represent important conditions for the effective application of TMT. During
thermomechanical rolling, the final rolling temperature is reduced to a point where
no further austenite recrystallisation occurs. The number of dislocations, with
numerous deformation bands, accumulated in austenite because of this, lead to a
fine grained D-microstructure during transformation.
If a high yield strength is especially desired, the last rolling passes can even take
place in the (JD) two-phase region, through which the already transformed ferrite
is directly strengthened through deformation. The level of the coiling temperature
during hot strip production, and an accelerated cooling belong to the most important
parameters influencing the microstructure during TMT.
Advances in the field of TMT are not possible without respective development in
the alloying concepts. Thereby, microalloying plays a dominating role. There is
hardly a procedure within TMT in which the actions of microalloying elements can
be disregarded.

337
6 Thermomechanical Treatment

Figure 6.2: Classification of hot deformation processes and schematic indication


of microstructure formation and effects of alloying.

338
6.2 Role of microalloying elements

6.2 Role of microalloying elements


A microalloying element is an element that is added in small amounts in order to
improve the mechanical-technological properties of steels. The concentration of
these elements is usually less than 0.1%. During the development of modern steels,
these elements have already been of great importance for a long period of time.
Alloying elements, such as Cr, Mo, and Mn, cause changes in the microstructure
through their influence on the matrix and thus, the material properties. In contrast to
alloying elements, microalloying elements have an effect mainly through the
precipitation of a second phase. Furthermore, in dissolved form, they can also help
in changing the properties. Ca or other rare earth elements are included to control
non-metallic inclusions. Al is the classic microalloying element for grain
refinement. In addition to these elements, the carbide and nitride forming elements
Ti, Nb, and V are of great interest because they enhance strength and toughness.
Since the last elements mentioned play an important role in TMT, their mechanisms
will be described in detail. A section of the periodic table of elements is given in
Figure 6.3 to help characterise the various elements.
The metals that melt at a high temperature (Groups IVa to VIa) have the potential to
form carbides and nitrides. This potential increases from the top right to the lower
left in the periodic table, although, the tendency towards nitride formation
predominates. The reason for this is a decrease in free enthalpy in the direction of
the arrow shown. As a result, Ti forms more stable precipitates than Nb and V.
Furthermore, chemical compounds containing nitrogen are more stable than those
with carbon.

Figure 6.3: Tendency of metals to form oxides, sulfides, carbides, and nitrides
(order according to the periodic table of elements).

339
6 Thermomechanical Treatment

Moreover, the elements in Group IVa show the tendency towards oxide and sulfide
formation, which is much more developed than the tendency towards carbide and
nitride formation. Since the precipitated particles that should effectively influence
the microstructure of a steel must be precipitated during heat treatment or hot
deformation, it is absolutely necessary that they are brought into solution before the
heat treatment or hot deformation takes place. An indicator for the possibility of
dissolving an element in the steel matrix is the difference in atomic size between the
respective element and the iron atom, as can be seen in Table 6.1.
The elements Zr and Hf, which have a strong tendency to form carbides and
nitrides, cannot be dissolved in steel, due to the significant difference between the
atomic radii of these elements and Fe. Therefore, these elements are not relevant
for a practical application. Because of these physical and chemical criteria as well
as the low availability of the element Ta, it is clear that only Nb, Ti and V are of
interest for a practical application as microalloying elements for a carbide and
nitride formation.
Figure 6.4 shows the relationship between the optimum enthalpy of formation and
the solubility of micro-allying elements. In order to allow the dissolution of the
alloying elements Nb, V, and Ti, and a desired precipitation behavior, a very
specific solubility is necessary in austenite. While TiN remains stable, even in the
melt, and VC first forms at low temperatures, only NbN, NbC, VN and TiC lead to
the desired effects. This is due to the fact that they are completely dissolved at
temperatures around 1200°C and can subsequently precipitate out in a finely
distributed manner during hot deformation down to finish rolling temperatures of
850°C.
All carbides and nitrides of the microalloying elements Nb, Ti and V have a face
centered cubic (fcc) structure and are completely soluble in each other. As a result,
complex carbide-nitride precipitates can also be formed, such as (Ti,Nb)(C,N).
Metal Atomic Difference to
radius in nm Fe atom in %
Ti 0.147 + 14.8
V 0.136 + 6.2
Cr 0.128 + 0
Zr 0.160 + 25.0
Nb 0.148 + 15.6
Mo 0.140 + 9.4
Hf 0.168 + 31.3
Ta 0.148 + 15.6
W 0.141 + 10.2
Table 6.1: Atomic radii and differences in atomic radii to Fe.

340
6.2 Role of microalloying elements

Figure 6.4: Enthalpies of carbides and nitrides and their effect on precipitation
during hot rolling.

6.2.1 Solubility behavior of microalloying elements


The elements Nb, Ti and V form carbides, nitrides, and carbo-nitrides with the
carbon and nitrogen that are present in steel. In doing so, the three microalloying
elements differ in their solubility behavior because of the differences in their
enthalpies of formation.
The equilibrium conditions for the dissolution and formation of precipitates are
described by means of their solubility products in the form:
log K = log [M][X] = A + B/T (6.1)

where K = the equilibrium constant, [M] and [X] = the concentrations of the
microalloying elements and carbon or nitrogen, A and B = constants, and T =
temperature in Kelvin.
Figure 6.5 shows various solubility products for carbides and nitrides, for both
austenite and ferrite. A region of typical values for the solubility product of [Nb]
and [C] in modern steels is also marked. At temperatures above 1150°C, the
solubility product is larger than the product of the concentrations of Nb and C. As a
result, Nb and C are fully dissolved in austenite. The solubility product decreases
when the temperature decreases. Consequently, in steels within the higher
concentration range, the solubility product is already exceeded at temperatures
between 1150 and 900°C, and NbC precipitates out. The exact solubility
temperature of precipitates is dependent on the concentration of carbon, nitrogen,
and the microalloying elements as well as the size of the precipitates. In steels with
a low content of Nb and C, precipitation first takes place after transformation, since
solubility in ferrite is significantly lower than in austenite, which is noticeable by
the solubility gap between the two-phases.

341
6 Thermomechanical Treatment

Figure 6.5: Solubility products of microalloying elements; for example: NbC in


austenite: log [Nb][C] = 2.26 - 6770/T; NbC in ferrite: log [Nb][C] =
4.62 - 10990/T. Typical values for the solubility product [Nb] [C] in
modern steels are shaded gray.
Titanium forms very stable compounds with both nitrogen and sulfur (TiN,
Ti(C,S)), which only dissolve in the melt. Vanadium-carbides and vanadium-
nitrides, on the other hand, dissolve in austenite beginning at around 1000°C. VC
precipitates out first during or after the J→Dtransformation. The solubility
products of TiC, NbN and NbC are between those of TiN and the compounds with
VC.
Isothermal solubility curves are a common method to describe solubility products.
The concentration of the microalloying elements is graphed over the concentration
of carbon or nitrogen, and the solubility product is shown for a constant
temperature. When the axes are scaled linearly, the isothermal solubility curves
have a hyperbolic shape, and for logarithmic scaling, the curve is linear. The
isothermal solubility curves signify the solubility boundaries; in all products above
this line, the respective alloying elements can no longer be completely dissolved.
Therefore, they precipitate out. A straight line through the origin with a positive
gradient represents the stoichiometric ratio of the compound.
Figure 6.6 shows a schematic solubility diagram. The points A, B, and C in the
left-hand graph signify the chemical compositions of different steels. These were
heat treated at 1300°C, subsequently cooled at a very low rate to 900°C, and then

342
6.2 Role of microalloying elements

rolled. Different types of precipitates result, as shown in the right hand graph.
Under the conditions given, NbC dissolves completely in the steels B and C at
1300°C. During cooling, precipitation proceeds along the line through point B, and
the distance BC is proportional to the volume content of precipitated NbC. With
this, fine precipitates develop, which are distinguishable by their means of
formation. On the one hand, these precipitates developed during cooling, and on
the other hand, these precipitates formed either during or after deformation.
However, the former hardly play a role, since precipitation in non-deformed
austenite is usually extremely slow. The latter are designated deformation-induced
precipitates, and play a large role in the course of hot deformation, which will be
discussed later in more detail.

Figure 6.6: Schematic solubility diagram.


If a steel with the composition A experiences the same treatment as a steel with the
composition B, then two different types of precipitation can be expected. The one
will have relatively large precipitates that do not dissolve during heat treatment at
1300°C, and whose volume content is proportional to the distance AB. The other
will have very fine deformation-induced precipitates with a volume content
proportional to the distance BC. Figure 6.7 shows the isothermal curves of NbC
and TiC, for both the upper and lower temperature ranges of austenite as well as two
isothermal curves of VN in the lower temperature range of austenite.

343
6 Thermomechanical Treatment

NbC TiC VN
Solubility
log [Nb] [C]= - 6770 + 2.26 log [Ti] [C]= - 7000 + 2.75 log [V] [N]= - 8330 + 3.46
product: T T T
0.14
1000°C
Alloying element in %

0.12
0.10
1200°C
0.08
0.06
0.04 1200°C
900°C
0.02
900°C 900°C
00 0.1 0.2 0.3 0 0.1 0.2 0.3 0 0.005 0.010
Carbon content in % Carbon content in % Nitrogen content in %

Figure 6.7: Isothermal solubility curves at various temperatures of NbC, TiC, and
VN in steel.
In a structural steel with 0.08% C, Nb concentrations of up to 0.08% and Ti
concentrations of up to 0.11% can be brought into solution during austenitisation at
1200°C. In the course of a TMT, these elements precipitate out as fine particles
during hot deformation, which delays the recrystallisation of the deformed
austenite. The isothermal solubility curves of VN show that even at a low
temperature, V is fully dissolved in austenite, especially considering the high
affinity of Al for N. In order to provide an overview of the course of precipitation
for the microalloying elements, Figure 6.8 shows an example of a sequence of the
temperature-dependent precipitation in a mico-alloyed steel.

Figure 6.8: Sequence of the temperature-dependent precipitation of Ti and Nb


compounds in a (Ti + Nb)-alloyed steel.

344
6.2 Role of microalloying elements

For most steels, the minimum austenitising temperature is determined by the content
of Nb and C. As can be seen in Figure 6.9, a steel with 0.1% C, 0.08% N and
0.03% Nb has a minimum austenitising temperature of 1150°C. The presence of Ti
leads to the formation of stable TiN particles, whereby nitrogen is bound and thus
unavailable for the formation of Nb(C,N). As a result, only NbC forms, which
dissolves already at low temperatures.
The equilibrium conditions, which are described by the solubility product, are
reached during heat treatment, such as austenitisation before a hot deformation.
During hot deformation, however, equilibrium conditions are not reached. As a
result of this, a somewhat larger amount of microalloying elements is in solution
than is calculated by means of the solubility product. Furthermore, it must be taken
into consideration that the solubility product can be influenced by the presence of
other elements, in so far as these elements change the activity coefficients. In this
way, the solubility product is increased through the addition of, for example, Mn,
and decreased by Si.

Figure 6.9: Solubility of NbC and Nb(C,N) in soft steels.


Solubility products solely make a rough classification of the solubility temperatures
possible and disregard the interaction of the alloying elements with one another.
However, programs such as CHEMSAGE and Thermo-Calc can calculate the
complex thermodynamic equilibrium states for multi-component systems. This is

345
6 Thermomechanical Treatment

done using the CALPHAD (CALculation of PHAse Diagrams) method. This


method is fundamentally based on the determination of free enthalpy minima. For
this, a database is used that contains basic thermodynamic data, such as specific
heat capacities, entropies, reaction enthalpies, and interaction parameters for
individual components and compounds that are possible within the system. The
equilibrium state is derived by calculating the Gibb’s free energy of the entire
system and varying within the system the type of phases formed, as well as their
composition, so that the Gibb’s energy reaches a minimum. A simplified
calculation for diffusion-controlled procedures can be made by coupling
thermodynamic and kinetic data entries, as done in the DICTRA program.

6.2.2 Precipitation kinetics


The precipitation of particles takes place by nucleation, nucleus growth, and grain
coarsening. These processes cannot be viewed in isolation, since they proceed
concurrently but at different rates, and are mutually dependent upon one another.
While some particles are just forming, others are growing and ageing already due to
more energetically favorable conditions. A critical energy must be reached in order
for nucleation to take place. The various factors that determine the magnitude of
this critical energy for nucleation are described by thermodynamics. The rate at
which the critical energy is successfully overcome is described by kinetics.
During and after deformation, microalloying elements begin to precipitate out. In
doing so, precipitation progresses most rapidly in the middle temperature range. At
high temperatures, while the diffusion rate for nucleus growth is very large, the
oversaturation, and with that, the driving force for precipitation is low. Therefore,
nucleation progresses very slowly. At low temperatures, nucleation progresses
more quickly, however nucleus growth is limited by a low rate of diffusion. The
precipitation curves in a temperature-time diagram show a characteristic C-shape,
due to the interaction of nucleation and nucleus growth (Figure 6.10). Through
deformation, the density of spots for nucleation is distinctly raised, and with that,
the kinetics of deformation-induced precipitation is considerably accelerated.

346
6.2 Role of microalloying elements

1100
C Dutta & Sellars
N Weiss & Jonas
Nb
Temperature in °C
1000
increasing pre-deformation

900

M = 0.6 0.14 0.05 0

800
10-1 100 101 102 103 104
Time in s
Figure 6.10: Precipitation diagram of Nb(C,N) in non-deformed austenite or during
deformation.

6.2.3 Mechanisms of microalloying elements


The influence of microalloying elements on the microstructure development in
steels is dependent on the dissolution and precipitation temperatures. The
precipitation temperature has a significant influence on precipitation size, and with
that, on the mechanism of precipitation, as shown in Figure 6.11.

Figure 6.11: Influence of precipitation temperature on particle size, important


mechanisms.

347
6 Thermomechanical Treatment

The microalloying elements can influence the processes of metal physics during hot
deformation procedures in both dissolved and precipitated form, contributing to an
increase in strength and toughness, whereby the influence of the precipitated
particles is greater. The increase in strength and toughness are reached
proportionally through grain refinement and hardening. Only grain refinement,
however, leads to a simultaneous increase in both strength and toughness. The
mechanisms behind this are:
x Restriction of austenite grain growth,
x Restriction of recrystallisation during hot deformation,
x Influence of the transformation behavior, and
x Formation of precipitates.
An overview of the mechanisms of the various microalloying elements concerning
the hot rolling process is given in Figure 6.12.

348
6.2 Role of microalloying elements

Figure 6.12: Mechanisms of dissolved and precipitated niobium, vanadium, and


titanium in steel.

349
6 Thermomechanical Treatment

6.2.4 Influence of microalloying elements on austenite grain growth


During heating before hot deformation, the average austenite grain size increases as
the temperature increases (Figure 6.13). Microalloying elements effectively hinder
the grain growth process, as long as they are present in the form of particles up to a
certain size. This effect can be overcome through coarsening or by the dissolution
of particles in the temperature ranges associated with the particular microalloying
elements. As a result, the austenite grain size greatly increases. This is shown in
the shaded area in Figure 6.13.
While non-microalloyed C-Mn steels experience steady grain growth with
increasing temperature, microalloyed steels show a sharp increase in grain size
when the dissolution temperature particular to the respective microalloying
elements is reached. Since vanadium already goes into solution at a temperature of
around 1000°C, vanadium precipitates do not participate in the hindrance of grain
growth at austenitising temperatures around 1200°C. Therefore, an effective
control can only be accomplished with niobium, and especially titanium, whose
stable compounds with nitrogen and sulfur only dissolve in the melt.
400

300
Grain size in μm

200

100
C-Mn
V Al Nb Ti

0
800 900 1000 1100 1200 1300
Temperature in °C
Figure 6.13: Austenite grain growth in steels with different microalloying
elements.

6.2.5 Influence of microalloying elements on softening behavior


In contrast to unalloyed steels, in which softening takes place very rapidly during
hot deformation, in microalloyed steels, the incubation time begins before
recrystallisation, and the critical degree of deformation is increased. This happens,
on the one hand, through the “solute drag effect”, in which the movement of
dislocations, grain boundaries, and subgrain boundaries is slowed by dragging
along the dissolved alloying atoms. On the other hand, an even more potent effect

350
6.2 Role of microalloying elements

is the deformation induced precipitation of very fine particles from the matrix, as
shown in Figure 6.14. Because of these finely distributed precipitates, dislocations
and grain boundaries are frozen, and thus, the nucleation of recrystallisation is
greatly hindered. Figure 6.14 shows the curve of static softening between two roller
stands at 1000 and 900°C for a C-Mn-steel, as well as two steels microalloyed with
Nb that differ solely in their carbon content. Additionally, the fraction of
precipitated Nb is given for the microalloyed steel with the higher carbon content.
It is demonstrated that the delay in softening after a pause between two deformation
passes is approximately two seconds greater for the steel with a higher carbon
content but equal Nb concentration, because the Nb can precipitate out as NbC.
Due to the significantly higher content of Nb precipitates at 900°C, softening is
considerably delayed at low temperatures.

Figure 6.14: Influence of dissolved and precipitated Nb at two deformation


temperatures on the delay of recrystallisation in austenite; graphed is

351
6 Thermomechanical Treatment

the decrease in strength between the first and second deformation pass
as a function of pause duration.
The nucleation of particles and the diffusion of microalloying elements are
facilitated through the delay in softening, and thus, the precipitation process is
significantly accelerated in comparison to the non-deformed state. This relationship
is clarified in Figure 6.15.
Above the temperature Tcr, recrystallisation ends before the precipitation process,
and no interactions between the two processes are evident. Below T cr,
recrystallisation can no longer progress to completion, and precipitation will be
accelerated when the deformed solid solution is sufficiently oversaturated.
Simultaneously, the anchoring of dislocations, grain boundaries, and subgrain
boundaries begins through the particles that are formed, which delay
recrystallisation. However, coarse precipitates that are not dissolved during
austenitisation can serve as nucleation sites, which actually accelerate
recrystallisation.

Figure 6.15: Interactions between recrystallisation and precipitation.


Nb has proven itself to be the most effective microalloying element to delay
softening, as shown in Figure 6.16. Even low concentrations of Nb lead to a large

352
6.2 Role of microalloying elements

increase in the TNR-temperature, and with this, to a significant restriction of


recrystallisation, already at relatively high temperatures.

1100
Nb
1050
for complete recrystallisation in °C Steel composition:
0.07% C, 1.4% Mn
1000
Minimum temperature

deformation per pass


10-15%
950

Ti
900
Al V
850

800

750
0 0.05 0.10 0.15 0.20 0.25
Dissolved content in %
Figure 6.16: Delay in recrystallisation in austenite through microalloying elements

6.2.6 Influence of microalloying elements on transformation behavior


Transformation behavior is also influenced by microalloying elements. As with
nearly all dissolved alloying elements, dissolved microalloying elements lead to a
delay in diffusion-controlled processes, and thus to a decrease in the
J→Dtransformation temperature. This delay becomes greater as the difference in
atomic size between the alloying elements and Fe increases. In this respect, Nb is
the most effective microalloying element.
Precipitated microalloying elements have the reverse effect. Transformation is
accelerated through coarse precipitates and the associated decreased concentrations
of C, N and microalloying elements in the matrix. This is demonstrated in Figure
6.17 per example of different austenitising temperatures.
After an austenitisation at 900°C, with an increasing Nb concentration, undissolved
NbC particles lead to an accelerated ferrite transformation; the transformation into
ferrite begins already at high cooling rates. This can be seen in a TTT diagram as a
shift to shorter times of the “ferrite nose”. In contrast, at 1250°C, approximately
0.06% Nb is dissolved, and the transformation is delayed so that ferrite does not

353
6 Thermomechanical Treatment

form until cooling at a lower rate. The opposite effect for both dissolved and
precipitated microalloying elements becomes apparent at 1100°C. At this
temperature, as much as 0.03% Nb can be dissolved, which leads to a delay in
transformation. At higher Nb concentrations, the precipitates do not completely
dissolve and can greatly accelerate the transformation.

Figure 6.17: Influence of Nb on transformation behavior at various austenitising


temperatures.

6.2.7 Precipitation hardening


Small precipitates, which are formed at low temperatures, contribute significantly to
an increase in strength, whereby the amount of precipitation, as well as the size of
the particles plays an important role (Figure 6.18).
Precipitations with a size of 1 to 2 nm, which form during or after the
J→Dtransformation in the ferritic phase, are very effective for precipitation
hardening. As the transformation temperature increases, or the cooling rate
decreases, the size of the precipitates increases. The precipitates become incoherent
in the steel matrix because they formed before the J→D transformation. These
coarse particles are less effective in increasing strength, due to their size and

354
6.2 Role of microalloying elements

because of their incoherent nature. Undissolved precipitates, such as TiN, and


precipitates that form during hot deformation or heat treatment that are relatively
coarse, hardly have an influence on the increase in strength. These precipitations
mostly have an effect on grain refinement, which in turn, positively affects the
toughness properties.
Although the elements Nb, Ti, and V, as well as their precipitates are principally
very similar in their behavior, their influences on toughness, and the increase in
strength differ greatly, as is shown in Figure 6.19 and Figure 6.20. The acid
soluble content refers to very fine precipitates.

Figure 6.18: Increase in yield strength 'VS through Nb as a function of Nb


concentration and precipitation size.

355
6 Thermomechanical Treatment

Figure 6.19: Increase in yield strength through precipitation hardening for a steel
with 0.01 to 0.5% carbon (austenitisation at 1300°C and subsequent
isothermal heat treatment at 600°C for maximal increase in yield
strength).

Figure 6.20: Influence of precipitation and grain refinement on the transition


temperature and yield strength of microalloyed steels.

356
6.2 Role of microalloying elements

For instance, the addition of 0.03% Nb to steel leads to an increase in strength of


approximately 150 MPa and to an improvement in the transition temperature of
about 30°C, due to precipitation hardening and grain refinement. In contrast, the
addition of 0.03% V causes an increase in strength of 50 MPa and an increase of 5K
in the transition temperature, since the fraction of grain refinement is very small.
Precipitation and grain refinement have the effect of increasing strength, with
respect to yield strength. While precipitation is disadvantageous on toughness,
grain refinement always has a beneficial effect toughness.
To sum up what has been discussed so far, Ti, V, and Nb very effectively influence
the microstructure during hot rolling, both as dissolved elements and in the form of
carbides and nitrides. Since the solubility products and physical properties of the
microalloying elements differ, their mechanisms are also very different.
Ti forms very stable nitrides, which effectively delay austenite grain growth in a
pusher type furnace. Furthermore, Ti has a positive effect on the toughness through
binding N. It also has a positive influence on the effectiveness of Nb, since N is no
longer available to form Nb(C,N). Instead, very fine NbC precipitates are formed.
Nb has proved itself to be a very effective element concerning grain refinement. On
the one hand, it regulates the austenite grains during heating, and on the other hand,
it effectively delays recrystallisation during thermomechanical rolling and causes a
decrease in the JoDtransformation temperature. The grain refinement achieved
through this cannot be obtained through any other form of heat treatment.
Furthermore, Nb is very effective with respect to precipitation hardening by forming
extremely fine niobcarbides, which lead to a large increase in strength.
Vanadium hardly forms any precipitates in austenite, and therefore, is available for
precipitation hardening during or after the JoD transformation. This is especially
beneficial when heat treatments are applied.

357
6 Thermomechanical Treatment

6.3 Factors influencing the TMT of microalloyed steels


Hot deformation is associated with a series of metallurgical processes that
significantly determine microstructural properties before, during, and after
deformation. An overview of the most important processes that occur during the
production of hot strips is given in Figure 6.21. A wide spectrum of processes
occur in the microstructure, which often temporarily overlap and influence one
another during all procedures starting with austenitisation in a pusher type furnace,
through an entire series of rolling passes, to subsequent cooling in the cooling zone,
and finally, during coiling. These mechanisms can be controlled through a
multitude of variable parameters during or after hot deformation:
x Austenitising temperature,
x Deformation temperature,
x Final deformation temperature,
x Degree of total deformation,
x Division of total deformation into the individual deformation steps (rolling
schedule),
x Deformation rate,
x Final deformation rate,
x Pauses between deformation steps, and
x Cooling rate.
The features and interactions of these processes are discussed in the following
sections.

6.3.1 Austenitisation
The goal of TMT is to obtain a fine transformation microstructure. The prerequisite
for this is an appropriate, homogeneous initial state before transformation, obtained
through austenitisation. Additionally, the purposeful development of the austenitic
structure throughout the process is of great importance. The microstructural
processes in austenite are:
x Grain coarsening of austenite during austenitisation,
x Grain refinement of austenite through repeated recrystallisation during
deformation, and
x Grain elongation in austenite without recrystallisation through deformation.
During austenitisation, the temperature of the pusher type furnace must not be too
low. This ensures that the necessary homogenisation of the austenite is obtained

358
6.3 Factors influencing the TMT of microalloyed steels

and the rolling force can be kept low during the following deformation. If the slab
temperature is too high, however, it can lead to coarse austenite grains through
intense grain growth, and with this, to poor final properties. Other disadvantages of
a high austenitising temperature are descaling and a high energy consumption with
associated high costs.
A fine austenitic microstructure is a fundamental prerequisite for a fine ferrite grain.
Therefore, the temperature must be high enough to produce homogeneous austenite
and to dissolve enough microalloying elements in austenite, which will be sufficient
for the subsequent TMT. The dissolution of microalloying elements at the
beginning of the treatment is an important prerequisite in order for them to function
throughout the TMT process. The purpose of microalloying is to sufficiently
dissolve the less stable particles, while the more stable particles retain their
inhibiting effect on grain growth during heating.
The rate of grain growth increases with increasing austenitising temperature.
However, grain growth can only occur if no growth inhibiting obstacles are present.
Fine precipitates on the austenite grain boundaries are one example of an obstacle.
Grain growth can occur when the particles finally coagulate or even dissolve at high
temperatures. This relationship is shown in Figure 6.22. Grain growth is distinctly
suppressed at an austenitising temperature of 950°C. After holding at 1050°C for
6-12 hours, the precipitates dissolve, causing a significant increase in grain size,
from approximately ASTM 8 to ASTM 0. At 1150°C grain growth starts after only
a few minutes.

6.3.2 Deformation
Besides the influence of the austenite grain size, the ferrite grain size is also
affected by the recrystallisation and transformation behavior. The most important
goal of hot deformation is to obtain a predefined austenitic state before the
JoDtransformation.
During a multi-stage deformation, which often extends over a broad temperature
range, the austenite grain refinement takes place through dynamic or static
recrystallisation. The composition of the steel, the temperature, and the deformation
parameters each determine one or the other recrystallisation mechanism.
The metallurgical phenomena that occur during hot rolling of wide strips under the
circumstance of thermomechanical treatment are summarised in Figure 6.21. The
hot rolling of wide strips can be divided into five steps: heating (in a pusher type
furnace), roughening, finish rolling, rapid cooling (in the cooling zone), and slow
cooling (in the coil). During these steps, softening and strengthening processes
occur, sometimes simultaneously.

359
6 Thermomechanical Treatment

Figure 6.21: Metallurgical phenomena during hot rolling of wide strips in the
different steps of a thermomechanical treatment.

Figure 6.22: Influence of austenitisation temperature and annealing time on


austenite grain size.
A hot rolling pass (Figure 6.23) exemplifies the transition from a coarse to a fine
microstructure through dynamic recrystallisation, i.e. recrystallisation during

360
6.3 Factors influencing the TMT of microalloyed steels

deformation, and static recrystallisation, i.e. recrystallisation after deformation.


Dynamic recrystallisation, which dominates at high temperatures, occurs during
rolling as soon as a critical degree of deformation Mcrit is reached. This can occur
multiple times during one pass. After deformation is complete, all of the grains
present have been recrystallised, and some have even been rehardened so that
complete softening does not occur until after deformation. Static recrystallisation
does not begin until after an incubation period following deformation, i.e. between
the individual rolling passes and after final rolling. The duration of static
recrystallisation is dependent on temperature and can last from a few seconds to
several hours.

Figure 6.23: Course of austenite recrystallisation during hot rolling.


The microalloying elements V, Nb, and Ti increase the critical degree of
deformation for dynamic recrystallisation (Figure 6.24), as well as the incubation
time allowing for the start of static recrystallisation (Figure 6.25). Under the
conditions of technological deformation processes, dynamic recrystallisation occurs
only at high temperatures, for example at temperatures >1000°C for a Ti-alloyed
steel, if the degree of deformation per pass is d0.05, as shown in Figure 6.24.
However, an effective delay in recrystallisation is mainly caused by the fraction of
deformation-induced precipitation. The fine carbo-nitrides that result block the
movement of dislocations and prevent softening by this means. In this way, the
deformed austenite grains provide a very high number of nuclei for the J→D
transformation, through their large grain surface area, the large dislocation density,
and the coarse, undissolved precipitates. Therefore, a very fine D microstructure is
present after transformation. This microstructure shows a distinct texture (grain
orientation) in the deformation direction of the austenite. The grain refining, and
therefore toughness improving effect increases with the final degree of deformation,
as shown in Figure 6.26.

361
6 Thermomechanical Treatment

1.0
End Ti-steel
Nb-steel
0.8
Degree of deformation M

Start
0.6

0.4
Ti-Stahl
0.2

0
900 950 1000 1050 1100 1150 1200 1250
Deformation temperature in °C
Figure 6.24: Influence of transformation temperature and degree of deformation on
the dynamic recrystallisation of Nb and Ti alloyed steels.
1 min 10 min 1h 10 h
1050
Temperature in °C

1000
0.21% Ti

950

unalloyed 0.06% Nb
900
0.06% Ti
850 0 1 2 4 5
10 10 10 103 10 10
Time in s
Figure 6.25: Static recrystallisation: holding time for 90% softening of an
unalloyed steel as well as for Nb- and Ti-alloyed steels with 0.05% C
and 0.6% Mn.

362
6.3 Factors influencing the TMT of microalloyed steels

Figure 6.26: Relationship between degree of deformation in the final deformation


pass, ferrite grain size, and transition temperature of microalloyed
steels.

6.3.3 Cooling
The influence of microalloying elements on phase transformation depends on the
alloying content as well as the austenitisation temperature and the state of the
precipitates. Low austenitising temperatures lead to an accelerated ferrite formation
because on the one hand, the carbon content of the matrix is lowered through
undissolved carbo-nitrides and on the other hand, the precipitates themselves are
nucleation sites for the J→D transformation. A higher level of microalloying

363
6 Thermomechanical Treatment

elements also leads to an accelerated transformation by raising the solubility


temperature. At high austenitising temperatures, the level of carbon and
microalloying elements in the matrix is increased, and the nucleus density is
reduced after dissolution of the carbo-nitrides. Therefore, the formation of ferrite is
delayed and a transformation into bainite or martensite can occur (Figure 6.27).
Through the addition of microalloying elements, which increases strength and
toughness, the carbon content of construction steels can be greatly reduced. The
microalloyed steels show such a short transformation time that danger of increasing
the hardness practically does not exist.

Figure 6.27: Cooling time for a 50% transformation in pearlite of a Ti-


microalloyed steel with 0.08% C and 0.8% Mn after different
austenitisation conditions.

6.3.4 Influence of deformation on the J→D transformation


For most transformation processes, nucleation begins at the austenite grain
boundaries, through which the austenite grain refinement gains special importance.
During TM rolling, the critical nucleation frequency for D-grain refinement is
increased by three main mechanisms:
x Increasing the effective grain boundary area SV per unit volume by elongating
the austenite grains,
x Increasing the nucleation rate on the defect-rich grain boundaries of deformed
austenite, and

364
6.3 Factors influencing the TMT of microalloyed steels

x Additional potential nucleation sites through deformation banding.


Figure 6.28 shows how an increase in the cooling rate leads to finer ferrite grains.
Additionally, the influence of deformation before transformation in austenite on the
ferrite grain size is clearly recognisable.

Figure 6.28: Influence of cooling rate and austenite state on ferrite grain size of hot
strips.

6.3.5 Coiling temperature


The amount of precipitates rapidly decreases with decreasing coiling temperature,
which is caused by the decreasing diffusion rate. At the same time, however, a
large increase in strength can be seen. This can be explained by the increased
nucleation that occurs through more intense super cooling and the finely dispersed
precipitates associated with it. At temperatures around 600°C, the increase in
strength reaches a maximum (Figure 6.29).

365
6 Thermomechanical Treatment

Smaller ferrite grain size


650
Yield strength in MPa e ss
ed
pr
600 su
VC

fin
rVe
C
550

500
450 500 550 600 650 700 750
Coiling temperature in °C
Figure 6.29: Influence of coiling temperature on yield strength of a Vanadium
microalloyed steel.
At these temperatures, the J→D phase transformation is already complete. By
varying the coiling temperature, maximal hardening can be reached as a
compromise between growth rate (particle size) and nucleus density (number of
particles). If the coiling temperature is too high, fewer precipitates form that grow
quickly and cause a marginal increase in strength. If the ageing temperature is too
low, the precipitate nucleus density is large and particle growth is suppressed so
that the full hardening effect is not reached. In microalloyed steels, grain
refinement as well as hardening increase the strength, while toughness is only
improved through grain refinement. Therefore, it is of importance how much each
mechanism contributes to the final characteristics.

6.3.6 Summary of the influencing variables


In Figure 6.30 the influencing variables are summarised for each of the individual
process steps of thermomechanical treatment in terms of the metallurgical
mechanisms. It is clear that thermomechanical treatment requires an exact
coordination between the chemical composition of the steel and the temperature-
time-deformation sequence. In all cases, control of the recrystallisation process and
the interaction between microstructural formation and precipitation are
characteristic. The great advantage of thermomechanical treatments is that an
optimised microstructure can be adjusted for a favorable combination between
strength and toughness during the finishing process, without an additional annealing
treatment.
The influence of the process parameters during rolling is shown in Figure 6.31.

366
6.3 Factors influencing the TMT of microalloyed steels

1) Austenitising temperature: Before the start of hot deformation, the material


that will be rolled must be austenitised. In doing so, the temperature in the
reheating furnace should not be too low or too high, in order to keep the
rolling strength at a minimum. If the austenitising temperature is too high,
then this results in coarse austenite grains through strong grain growth, which
leads to a decrease in the yield strength and an increase in the transition
temperature. Which as a result would lead to a final product with poor
mechanical properties.

2) Final rolling temperature: As the final rolling temperature increases so does


the percentage of softening. The number of nuclei for phase transformation is
reduced due to dislocation reduction and due to recrystallisation of the
austenitic microstructure. This results in a coarse microstructure, which leads
to a lower yield strength and a higher transition temperature.

3) Degree of deformation: As the final degree of deformation increases, the


dislocation density and the effective grain boundary area SV also increases.
With this, the number of nuclei for phase renewal increases, which results in
a fine grained microstructure with a high yield strength and a low transition
temperature.

4) Cooling rate: In order to further increase the nucleus density for


transformation, a greater supercooling can be adjusted by a higher cooling
rate. This creates a finer microstructure with improved properties.

5) Coiling temperature: The temperature dependency of hardening has its effect


here. At high temperatures, relatively coarse precipitates lead to only a slight
increase in the yield strength. However, if the temperature is too low, the
diffusion rate is so slow that the formation of precipitates is suppressed.

367
6 Thermomechanical Treatment

Figure 6.30: Schematic overview of property determining processes and their


influencing variables during TM rolling of fine grained construction
steels.

368
6.3 Factors influencing the TMT of microalloyed steels

Figure 6.31: Influence of rolling parameters on mechanical properties.

369
6 Thermomechanical Treatment

6.4 Further Readings

Béranger, G.; Henry, G.; Sanz, G.:


The Book of Steel
Intercept, Andover, 1996

Cahn, R.W.; Haasen, P.; Kramer, E.J.:


Materials Science and Technology: A Comprehensive Treatment
Pickering, F.B.:
Volume 7: Constitution and Properties of Steels
Wiley-VCH, December 1991

DeArdo, A.J.; Gray, J.M.; Meyer, L.:


Fundamental Metallurgy of Niobium in Steel
Niobium, H. Stuart, The Metallurgical Society of AIME, 1989

Gottstein, G.:
Physical Foundations of Materials Science
Springer Verlag, Berlin, Heidelberg, New York, Tokyo, 2004

Gladman, T.:
The Physical Metallurgy of Microalloyed Steels
The Institute of Materials, London, 1997

Krauss, G.:
Steels: Processing, Structure, And Performance
1. Edition, ASM International, August 2005

Lankford, W.T.; Samways, N.L.:


The making shaping and treating of steel
11th ed., United States of Steel, Pittsburgh, Pennsylvania, 1998

370
6.4 Further Readings

Llewellyn, D.T.; Hudd, R.C.:


Steels: Metallurgy and Applications
3rd ed., Butterworth-Heinemann, February 1998

Totten, G. E.:
Steel Heat Treatment: Metallurgy and Technologies
2nd ed., Taylor & Francis, 2006

Verein Deutscher Eisenhüttenleute (Ed.):


Steel
Volume 1: Fundamentals
Volume 2: Applications
Verlag Stahleisen, Düsseldorf
Springer Verlag, Berlin, Heidelberg, New York, Tokyo, 1992/1993

Yamamoto, S.; Ouchi, Ch. u. Osuka, T.:


Thermomechanical Processing of Microalloyed Austenite
The Metallurgical Society of AIME, Warrendale (PA), 1982, pp. 613-639

Cuddy, L. J.:
Fundamentals of the controlled rolling process
The Metallurgical Society of AIME, Warrendale (PA), 1986, pp. 235-243

Cuddy, L.J.:
Thermomechanical Processing of Microalloyed Austenite
The Metallurgical Society of AIME, Warrendale (PA), 1982, pp. 129-140

371
6 Thermomechanical Treatment

Grossmann, M.A.:
Elements of hardenability
Stahl und Eisen 111 (1991) Nr. 7, pp. 103-110

Meyer, L.; Heisterkamp, F. u. Müschenborn, W.:


Microalloying 75, Union Carbide Corp., Washington DC, 1975, pp. 153-
167

Robiller, G.; Meyer, L.:


Work hardening and softening behavior of Ti- and Nb-alloyed steels
during hot deformation
Recrystallization and grain growth of multi-phase and particle containing
materials
1st Riso International Symposium of Metallurgy and Materials Science,
1980, pp. 311-316

DIN EN 10052
Begriffe der Wärmebehandlung von Eisenwerkstoffen
Januar 1994

DIN EN 10083-1
Vergütungsstähle, Teil 1: Technische Lieferbedingungen für Edelstähle
August 2003

DIN EN 10084
Einsatzstähle, Technische Lieferbedingungen
Juni 1998

DIN EN ISO 18265


Metallische Werkstoffe – Umwertung von Härtewerten
Februar 2004

372
6.4 Further Readings

DIN EN ISO 642


Stahl – Stirnabschreckversuch (Jominy- Versuch)
Januar 2000

DIN EN 10020
Begriffsbestimmungen für die Einteilung von Stählen
Juli 2000

Stahl-Eisen-Prüfblätter (SEP) des VDEh 1664:


Ermittlung von Formeln durch multiple Regression zur Berechnung der
Härtbarkeit im Stirnabschreckversuch aus der chemischen Zusammen-
setzung von Stählen
1. Ausgabe, Düsseldorf, 1996

Stahl-Eisen-Prüfblätter (SEP) des VDEh 1680:


Aufstellung von Zeit-Temperatur-Umwandlungsschaubildern für
Eisenlegierungen
3. Ausgabe, Düsseldorf, 1990

Stahl-Eisen-Prüfblätter (SEP) des VDEh 1681:


Richtlinien für Vorbereitung, Durchführung und Auswertung
dilatometrischer Umwandlungsuntersuchungen an Eisenlegierungen
2. Ausgabe, Düsseldorf, 1998

373
7 Index

7 Index
300°C embrittlement, 256 bcc-lattice, 6
acicular ferrite, 188p, 196, 203 blade, 183p, 278
activation energy, 8pp, 137, 209, 290, Bloch walls, 54pp
317 blue brittleness test, 114
Ageing, 224, 229 boundary area, 14, 112, 145, 159, 219,
ageing resistance, 231 338, 371pp
allotropic modification, 72 Burns Curve, 179
allotropic transformation, 73 calcium-aluminate, 122pp
alloy formation, 65p carbide, 10pp, 78pp, 103pp, 128p,
Anelasticity, 47 134, 139, 144, 149, 153, 186, 189,
196, 209, 213p, 221pp, 247p, 255,
Anisotropic heterogeneity, 275 275p, 281, 285pp, 303, 308p,
annealing, 46, 90, 98pp, 182, 209, 312pp, 320, 325, 329pp, 344p
221pp, 239pp, 249, 257, 275pp, carburisation, 266
340, 366, 373
Case hardening, 243, 264
artificial ageing, 208, 228p
cementite precipitation, 212p, 228
austempering, 201, 253
chip separation, 278
austenite, 7pp, 12p, 73p, 80pp, 94,
101pp, 128pp, 147pp, 155pp, cluster formation, 65
243pp, 269pp, 276pp, 287, 300pp coagulation, 208p, 214, 219
Austenite, 101p, 233pp, 300, 312, 355, coarse grain annealing, 239, 278
378 coarsening, 214, 226, 248, 287, 291,
austenite transformation, 102, 169, 308, 336, 351, 355, 364
181, 300, 312, 323 coiling temperature, 207, 293, 342,
axis ratio, 166p, 181 372p
Bain model, 160pp colony, 139p, 148
Bake-hardening effect, 229 Combined annealing, 297
bake-hardening potential, 230p Continuous annealing, 227p, 293
banding, 101pp, 276, 280pp, 298, 372 continuous cooling, 101pp, 169, 202,
batch annealing, 291pp 253, 283, 300p, 316, 327p, 332

bcc lattice, 20p, 25, 31, 48pp, 66, 160 cooling rate, 13, 96pp, 104, 128, 142p,
157, 172p, 227, 240, 245, 248p,
bcc solid solution, 73

374
7 Index

259pp, 282p, 324, 328, 331, 360, diffusionless transformation, 104, 168
372pp dilatometer, 197, 300
cooperative shear movement, 156, 339 distortion, 66, 167p, 173, 212, 217,
critical cooling rate, 249, 331 220, 273, 285, 296p
critical nucleus radius, 214 double reaction, 139
cross-diffusion, 140p DTTT diagrams, 331
crystal defect, 28 eddy current losses, 60
crystal segregation, 90, 99pp Elastic modulus, 5, 41pp
Crystal Structure, 236 electrical conductivity, 6, 37, 54, 61p
crystal structures, 3, 18 end usage properties, 65
crystallographic ordering, 62 end-quench test, 259
CTT diagram, 331p Equilibrium diagrams, 300
Curie temperature, 6, 18, 30, 37, 43, equilibrium solubility, 209p
52pp, 57p eutectic, 6pp, 79pp, 88, 117p
cycle annealing, 287 eutectoid reaction, 139
damping, 47pp, 221, 232 Examination of segregation, 106
deep drawing steel, 45pp fcc-lattice, 6
deformation, 2, 9, 14, 28, 41p, 45pp, ferrite, 7, 13p, 73p, 79pp, 87, 101pp,
58, 67, 77, 100p, 105pp, 114, 128, 131pp, 185pp, 197pp, 212,
118pp, 128p, 160, 165, 174pp, 227, 253, 260, 269, 276pp, 287,
181p, 185, 200, 229pp, 270pp, 281, 293,298pp, 307pp, 312pp, 359, 365,
285, 288, 290p, 296, 331pp 370pp
Degenerated pearlite, 153 ferromagnetism, 51pp, 62
dendritic solidification, 93 fine grained steels, 278, 281
diffusion, 2, 8, 10, 13, 28, 38, 90, 92, foreign atoms, 27, 48p, 61, 65pp, 90
99, 100pp, 128pp, 155pp, 172, 189,
195pp, 201, 208, 212pp, 253, Gas bubble segregation, 95
275pp, 292, 303, 312, 317pp, Gibbs-Thomson Equation, 219
333pp, 351, 357pp, 372pp
Grain boundary, 90
diffusion annealing, 100
grain boundary area, 338
diffusion-controlled, 104, 128, 139,
grain refinement, 201p, 344, 353,
189, 318, 333, 338
361pp, 371pp
diffusion-controlled transformation,
Gravity segregation, 95
131, 138, 196, 321, 334, 337

375
7 Index

habit plane, 159pp, 202 Isotropic heterogeneity, 275


hardenability, 11, 149, 243, 253, JMAK Equation, 12, 142, 320
257pp, 286, 379 Jominy test, 12, 259pp
hardening capacity, 243pp, 260 Koistinen-Marburger, 221
Hardening stresses, 271 Kurdjomov-Sachs relationship, 163
hardness penetration, 243pp, 260 lamellae, 139, 145, 148, 152p, 283,
hardness penetration depth, 243 286, 298, 307, 325
hardness stresses, 271 lamellar spacing, 8, 142, 145pp, 325
heat treatment, 2, 76, 83, 103, 154, latent heat of transformation, 73
184, 205pp, 296pp lath martensite, 9, 129, 170
heating, 5, 12p, 18p, 29, 32, 76, 118, lath size, 200
181, 184, 227, 240, 243p, 253pp,
282p, 293pp, 332, 355, 363pp lattice constant a, 20, 29, 31
heating duration, 240 lattice constant aA, 159p
heating rate, 240, 305p, 309, 312 lattice constant aM, 159
Hexagonal martensite, 181 lattice constant cM, 168
holding time, 100, 241, 248p, 255, lattice defects, 4, 28, 37, 61, 90, 208,
275, 278, 282p, 297, 304pp, 331pp, 214, 218, 338
369 lattice distortion, 25, 66p, 160
Hume-Rothery rule, 27 lattice oscillations, 37, 53, 61, 67
hysteresis, 18, 54, 58p, 182pp lattice-changing deformation, 160,
incubation, 157, 209, 214, 218, 309, 166, 168
318p, 355, 367 lattice-retaining deformation, 160
induction constant, 6, 51 Leidenfrost temperature, 251
intermetallic phase, 28, 69 logarithmic decrement, 6, 48
internal cleanliness, 96, 114pp low temperature carbides, 213
Interstitial, 22pp, 66 lower bainite, 9, 186, 188p, 195, 197,
Invar steel, 35 202, 325
invariant deformation, 160 machinability, 239, 278, 280, 285p,
297
Inverse ingot segregation, 96
macrosegregation, 97, 99
isothermal martensite formation, 182p
magnetic domains, 54
isothermal transformation, 253, 278,
287, 316 magnetic field, 6, 51pp

376
7 Index

magnetic hysteresis loss, 54 normalising, 277, 280pp, 291, 297,


magnetic induction, 5, 51pp 342

magnetic moments, 51pp nucleation, 9p, 101, 134, 139, 142pp,


157, 189, 196p, 208, 212pp, 224,
magnetically harder, 54 249, 288pp, 309, 312, 316pp, 324p,
magnetostriction, 33, 58 333p, 336pp, 351, 356pp, 370pp
Marageing steels, 184 nucleation rate, 144, 216, 339
martensite, 9, 11, 13p, 87, 99p, 129, nucleus growth, 13, 140pp, 153, 157,
155pp, 194p, 202, 206, 243pp, 196, 208, 214, 288, 309, 316, 338,
253pp, 269pp, 302, 307pp, 331, 351
336, 339, 371 Nucleus growth, 218, 317
martensite finish temperature Mf, 169 nucleus growth rate, 144
martensite start temperature Md, 182 orientational relationship, 163, 165
Md30-temperature, 175p overageing, 209, 227pp, 294
melting point, 6, 18, 32, 43, 90, 117p, oxide inclusion, 112
122
packing density, 6, 20, 22
Microalloying elements, 340, 355
pancake microstructure, 296
micro-precipitations, 209
paramagnetism, 51, 62
microstructural banding, 275, 282
patenting treatment, 143, 299, 324
microstructural transformation, 320,
327 pearlite, 8, 14, 81p, 101p, 128pp, 185,
194, 196, 198, 233, 245, 248p, 253,
microstructure, 4, 14, 58, 65, 77, 96, 257, 260, 269, 278, 280pp, 298p,
101pp, 128pp, 135, 139, 142, 147, 307, 312pp, 371
149, 152pp, 165, 169, 179, 181,
183pp, 194, 198pp, 212, 227, 239, pearlite colonies, 147
241, 243pp, 253pp, 269, 273, peritectic, 68, 72, 79p, 88
275pp, 352, 363pp, 373p phase, 3, 9pp, 44, 48, 58, 65, 67pp, 99,
mobility rate, 8, 137p, 144 101, 105p, 112, 118p, 128pp, 153,
Nishiyama-Wassermann relationship, 155pp, 179, 181, 184p, 189, 196pp,
164 208, 211pp, 227, 247, 251, 276,
283, 288, 291pp, 299pp, 316,
nonequilibrium diagram, 301, 303 318pp, 360, 370, 373p, 379
nonequilibrium diagrams, 300p phase boundary, 91, 139pp, 214, 320,
Non-lamellar pearlite, 153 324
non-metallic inclusion, 112 plate martensite, 9, 165, 170pp
Plate martensite, 156

377
7 Index

polymorphic transformation, 18 saturation induction, 5, 53pp, 58


polymorphism, 18 Secondary cementite, 81
precipitate, 28, 69, 79p, 90, 133, 153, secondary segregation, 90
185p, 189, 208, 212pp, 254, 256p, secondary structure, 101
266, 294, 314, 325, 345, 347, 349,
351, 356, 373 segregation, 7, 68, 69, 90pp, 95pp,
255pp, 275p, 293, 298
precipitation, 10, 15, 69, 79, 87, 102p,
106, 128p, 135p, 139, 142, 189, segregation coefficient, 91
196, 203, 208p, 212pp, 255, 293, segregation coefficients, 104, 276
297pp, 316, 320, 323, 325, 328,
Segregation in continuous casting, 96
339, 344pp, 373
Shear modulus, 42p, 45
precipitation driving force, 214, 228
short brittleness, 278
preformed nuclei, 339
smear, 278
Primary cementite, 81
Snoek effect, 49p, 232
processing properties, 65, 239
soft annealing, 239, 281, 285, 287p
quenching, 77, 104, 177, 179, 183,
202, 209, 243pp, 271p, 281, 298, soft spots, 307
305, 307, 312p softening, 355, 357p, 365, 367, 369,
Quenching and tempering, 243 374, 379
quenching medium, 244, 251 solid solution, 9, 66, 67pp, 81, 92,
128, 131, 139, 177, 179, 202, 208,
railway steels, 154, 205
210, 212, 218, 225, 303, 309, 357
range of strength, 201
solid solutions, 57, 60, 80p, 93, 105,
Real Structures, 27 131, 133, 208, 303
recalescence, 152, 320, 328 Solubility, 67, 83, 210p, 346p, 350
recrystallisation, 12, 15, 227, 254, sorbite, 151p, 298
288pp, 336, 338, 342, 349, 353,
spheroidising, 285
355, 357pp, 363pp, 373p
stepped torsion test, 114
recrystallisation annealing, 288, 291
strain, 10, 185, 208p, 227, 230, 243,
recrystallisation diagram, 291
288, 297
regression coefficient, 11, 263
strength of martensite, 177
relative permeability, 6, 51, 52p, 58
strengthening, 147, 177, 200, 202,
remnant polarisation, 54 225, 230, 233, 291, 365
retained austenite, 155, 175, 177, 179,
246, 247p, 269, 308, 325, 329, 331

378
7 Index

stress, 6, 7, 41p, 47pp, 123, 155, 165p, thermomechanical treatment, 202, 340,
174pp, 182, 202, 208, 230, 239p, 365p, 373
243, 271p, 296p time exponent, 12, 290
substitution, 66 tool wear, 278, 280, 285
substitutional solid solutions, 69, 75, torsional pendulum, 48p
164
transformation temperature, 5, 8, 13,
Substitutional solid solutions, 65 18, 63, 72, 144, 146p, 150p, 170,
supercooling, 8p, 18, 128, 138, 142pp, 189pp, 276, 282, 309, 314, 316pp,
168, 195, 216pp, 227, 282p, 317, 321, 323, 331, 359p, 363, 368
324, 336, 374 troostite, 151p
superheating, 18, 94, 96, 98, 181, TTA diagram, 244, 303, 305, 307pp
281p, 309, 312
TTA diagrams, 300pp
superlattice, 65, 69, 182
TTT diagrams, 198, 300, 306, 316,
supersaturated solid solution, 129, 208 327, 331, 335
supersaturation, 69, 208, 212, 214, unit cell, 7, 18, 20pp, 26p, 29, 48,
216, 219, 221p, 228p 65pp, 134, 160, 166
surfboard” method, 114 upper bainite, 9, 129, 186, 188, 189,
t8/5-time, 328, 331 201, 325, 328
temperature-time cycles, 239 very fine lamellar pearlite, 152
tempering, 105, 201p, 239, 243p, V-segregations, 95
253pp, 271, 281, 287p, 297, 314 Widmannstätten ferrite, 202
Tempering, 237, 243, 253, 255, 257p Wire patenting, 298
tempering embrittlement, 255 WTTT diagrams, 331
Tempering resistance, 257 Zonal heterogeneity, 275
tertiary cementite, 81 zone without transformation, 197p
theoretical density, 6, 21, 23p
J-iron, 18, 22, 30, 32p, 75, 119, 168
thermal conductivity, 6, 36p, 62, 240,
J-solid solution, 139
249, 253
thermal expansion coefficient, 5,
30pp, 62

379

You might also like