Download as pdf or txt
Download as pdf or txt
You are on page 1of 168

Engineering Materials

For further volumes:


http://www.springer.com/series/4288
Hans Berns Valentin Gavriljuk

Sascha Riedner

High Interstitial Stainless


Austenitic Steels

123
Hans Berns Sascha Riedner
Ruhr-University Deutsche Edelstahlwerke GmbH
Bochum Kamen
Germany Germany

Valentin Gavriljuk
G.V. Kurdyumov Institute for Metal
Physics
Kiev
Ukraine

ISSN 1612-1317 ISSN 1868-1212 (electronic)


ISBN 978-3-642-33700-0 ISBN 978-3-642-33701-7 (eBook)
DOI 10.1007/978-3-642-33701-7
Springer Heidelberg New York Dordrecht London

Library of Congress Control Number: 2012949081

Ó Springer-Verlag Berlin Heidelberg 2013


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or
information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed. Exempted from this legal reservation are brief
excerpts in connection with reviews or scholarly analysis or material supplied specifically for the
purpose of being entered and executed on a computer system, for exclusive use by the purchaser of
the work. Duplication of this publication or parts thereof is permitted only under the provisions
of the Copyright Law of the Publisher’s location, in its current version, and permission for use must
always be obtained from Springer. Permissions for use may be obtained through RightsLink at the
Copyright Clearance Center. Violations are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

A close cooperation between the Institute of Metal Physics in Kiev, Ukraine, and
the Chair of Materials Technology at the Ruhr University Bochum, Germany, has
been going on now for more than 20 years. During the first decade the joint effort
centered on high nitrogen steels (HNS) and the partners published a book on this
subject in 1999. Already then they had shown that combined alloying of
martensitic stainless steels with carbon + nitrogen considerably improved structure
and properties. These findings were transferred to the hardenable stainless bearing
steel CRONIDURÒ used, e.g., in aviation and to SolNitÒ which allows case
hardening of stainless steel with nitrogen instead of carbon.
The second decade of partnership was dedicated to extending the beneficial
C+N concept to new stainless austenitic grades called high interstitial steels (HIS).
The results compiled in the present book have two major targets. On the scientific
side the structure/ property relation starts at the electron structure and is carried on
to the macroscale explaining the superior performance of HIS. The engineering
aspects cover major steps of industrial manufacture and possible applications.
Compared to similar HNS the new HIS do without costly pressure or powder
metallurgy. Thus the contents are of interest to materials scientists working in
R&D but also to engineers in design, manufacture, and materials selection.
The authors thank Prof. Dr. Bela Shanina (Theoretic Physics), Dr. habil. Yuri
Petrov (Electron Microscopy), Dr. Andrij Tyshchenko (Mössbauer Spectroscopy)
in Kiev and Dr.-Ing. Fabian Schmalt, Dr.-Ing. Lais Mujica-Roncery, and Dipl.
Ing. Nilofar Nabiran in Bochum for their most valuable contributions. Thanks
are also extended to other researchers, students, and technical staff involved in
the development of HIS and to Miriam Rockenbach, Agnes Krolik, and Dipl.
Ing. Fabian Pöhl for preparing the final manuscript. The authors are grateful to
Prof. Dr.-Ing. Werner Theisen, Head of Chair in Bochum, for his continuous
support to the HIS project and to the German research foundation (DFG) for

v
vi Preface

sponsoring part of the work. They feel also indebted to the National Academy of
Science and the Science and Technology Center in Ukraine for financial support.
Last but not least, we thank several companies for melting and processing new
HIS (see Chap. 5).

Summer 2012 Hans Berns


Valentin Gavriljuk
Sascha Riedner
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 High Nitrogen Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 High Interstitial Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Aim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Constitution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Variation of Interstitial Content . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Effect of C/N Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 Variation of Substitutional Content . . . . . . . . . . . . . . . . . . . . . 11
2.4.1 Chromium and Manganese . . . . . . . . . . . . . . . . . . . . . . 11
2.4.2 Molybdenum and Copper. . . . . . . . . . . . . . . . . . . . . . . 14
2.4.3 Tramp Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Selection of Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3 Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.1 As-Quenched . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.1 Electron Structure: Calculated and Measured . . . . . . . . . 22
3.1.2 Atomic Distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.1.3 Chemical Nanoscale Homogeneity . . . . . . . . . . . . . . . . 36
3.2 Structural Change by Loading. . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2.1 Tensile Straining. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2.2 Effect of Subzero Temperature . . . . . . . . . . . . . . . . . . . 59
3.2.3 Effect of Strain Rate . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.2.4 Effect of Cyclic Loading . . . . . . . . . . . . . . . . . . . . . . . 68
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

vii
viii Contents

4 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.1 Mechanical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.1.1 Tensile Properties at Room Temperature . . . . . . . . . . . . 85
4.1.2 Tensile Properties at Subzero Temperatures . . . . . . . . . . 88
4.1.3 Tensile Properties at Elevated Temperatures. . . . . . . . . . 88
4.1.4 Creep Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.1.5 Hardness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.1.6 Notch Impact Toughness . . . . . . . . . . . . . . . . . . . . . . . 93
4.1.7 Rotating Bending Fatigue. . . . . . . . . . . . . . . . . . . . . . . 94
4.2 Wear Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.2.1 Abrasive Wear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.2.2 Impact Wear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.2.3 Wear by Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.3 Corrosion Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.3.1 Submersion Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.3.2 Current Density/Potential Tests. . . . . . . . . . . . . . . . . . . 102
4.3.3 Tests on Intercrystalline Corrosion . . . . . . . . . . . . . . . . 106
4.4 Magnetic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

5 Manufacture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.1 Melting and Casting. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
5.1.1 Ingots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
5.1.2 Centrifugal Castings . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.1.3 Sand Castings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.1.4 Refractories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.2 Hot Working. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.3 Heat Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.3.1 Solution Annealing . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.3.2 Interrupted Quenching . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.3.3 Continuous Quenching. . . . . . . . . . . . . . . . . . . . . . . . . 120
5.3.4 Aging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5.4 Cold Drawing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5.5 Welding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
5.6 Machining . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

6 Assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.1 From Structure to Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.1.1 Mechanical Properties . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.1.2 Wear Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
6.1.3 Corrosion Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . 130
6.1.4 Nonmagnetic State . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Contents ix

6.2 From Manufacture to Application . . . . . . . . . . . . . . . . . . . . . . 131


6.2.1 Constitution and Hot Manufacture. . . . . . . . . . . . . . . . . 132
6.2.2 Workhardening and Cold Manufacture . . . . . . . . . . . . . 133
6.2.3 Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
6.3 Pros and Cons of HIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
6.3.1 Pros. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
6.3.2 Cons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

Appendix A: Tables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

Appendix B: Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

About the Authors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
Chapter 1
Introduction

Standard stainless austenitic CrNi (Mo) steels are used in a wide range of
applications because of their high corrosion resistance and ductility. Their low
interstitial content enhances weldability but lowers the yield strength. In contrast
to sheet material, castings and forgings do not as much rely on weldability but
would profit from more strength. In these cases solid solution strengthening by
interstitial atoms of carbon and nitrogen is a promising way of raising strength
without losing too much ductility. However, the solubility of interstitial elements
depends on the concentration of substitutional elements as e.g. chromium, nickel,
manganese and molybdenum. Of these, Cr and Mo are required for corrosion
resistance, Ni and Mn for austenite stability. Therefore a proper balance of sub-
stitutional elements has to be established to enhance the concentration of inter-
stitials and thus the strength of austenitic steels. In principle one may use C or N or
C ? N but C alone did not meet the requirements.

1.1 High Nitrogen Steels

Chromium reduces the activity of carbon in steel thus increasing its solubility.
However, the maximum solubility at the border to carbide precipitation, i.e. the
range of homogeneous austenite, is reduced. Nickel enhances the activity of
carbon and lowers its solubility in the lattice. The effect of chromium and nickel
on interstitial nitrogen is similar to that on interstitial carbon, except for a higher
solubility of the former. In contrast, volatile nitrogen escapes from the melt as N2
gas. This situation is expressed by a solubility of N [ C in austenite but N \\ C
in the melt and comprises the basic dilemma of stainless high nitrogen steels
(HNS) [1]: Nitrogen offers more solubility in austenite and therefore more strength
but requires special manufacturing routes to introduce high contents to the steel.
Of these melting and solidification under N2 pressure (pressure metallurgy) and

H. Berns et al., High Interstitial Stainless Austenitic Steels, 1


Engineering Materials, DOI: 10.1007/978-3-642-33701-7_1,
Ó Springer-Verlag Berlin Heidelberg 2013
2 1 Introduction

solid state nitriding of steel powder followed by hot compaction (powder metal-
lurgy) have been industrially exploited. But compared to standard ingot metallurgy
at normal pressure of air, pressure and powder metallurgy considerably increase
the costs.
Instead of special manufacturing processes suitable alloying may be used to
enhance the nitrogen solubility of the melt. It aims at an exchange of nickel by
manganese, because the latter raises the nitrogen solubility in contrast to the
former. About twice as much manganese as nickel is required to fend off d-ferrite
which is expressed e.g. by the Schaeffler diagram [2]. Steel Cr18Mn18N0.55
(in mass %) is e.g. used for retaining rings on electric generator shafts, steel
Mn23Cr21Ni2N0.85 for drill collars. The 0.2 % proof strength of these two HNS
grades, molten at normal pressure of air, was raised to about 400 and 600 MPa,
respectively, which amounts to two or three times the level encountered in stan-
dard steel Cr18Ni10. The increase from 0.55 to 0.85 mass % N in these two HNS
is brought about by a higher content of the substitutional elements chromium and
manganese which spur the nitrogen solubility. However, the fracture elongation in
tensile tests is lowered and the ductile to brittle transition temperature (DBTT) in
notch impact tests is raised pointing to some embrittlement.

1.2 High Interstitial Steels

To avoid this deficiency it was proposed to saturate a lean austenitic CrMn steel
with interstitial nitrogen at normal pressure and add interstitial carbon to further
raise the strength. This concept of high interstitial steels (HIS) with C ? N was
presented in 2002 [3] and steel Mn17Cr15N0.43C0.39 showed the following
mechanical properties: proof strength Rp0.2 = 494 MPa, true fracture strength
R = 2635 MPa, elongation A = 78 % [4, 5]. This remarkable combination of
strength and ductility was explained by an increase in the concentration ne of free
electrons in austenite via combined alloying with C ? N, compared to alloying
with C or N alone. A high ne enhances the ductile metallic character of interatomic
bonding. From previous work on the austenite of martensitic stainless steel it was
already known [1, 6] that—in the order of alloying with C, N or C ? N—short
range atomic ordering is promoted which stabilizes the austenitic phase and raises
the interstitial solubility. The significance of the C/N ratio for the precipitation of
M2N nitrides or M23C6 carbides and for welding of austenitic CrMn steels was
mentioned in [3].
The stepwise addition of up to 0.4 mass % C to steels with (mass %) 19Mn,
17Cr, 3Ni and 0.4–0.55 N raised strength and ductility but hardly the DBTT [7–9].
The corrosion resistance was improved as well, in that the passivation and re-
passivation was enhanced and the resistance to pitting corrosion strengthened.
These beneficial effects assigned to carbon may have been based on C ? N,
1.2 High Interstitial Steels 3

though. It was pointed out that the temperature of beginning precipitation


increased with the carbon content leading to embrittlement at lower temperatures.
The strengthening part of this precipitation was used in steels for exhaust valves,
as e.g. Cr21Mn9Ni4C0.53N0.42 which came into use already in 1952 [10]. They
are probably the earliest HIS, but were not meant to dwell on homogenous
austenite.
A most recent development started from high manganese TWIP steels which
rely on twinning induced plasticity [11]. Chromium was added to provide a
moderate corrosion resistance and C ? N to enhance the strength. Steel
Mn25Cr12C0.32N0.45 for instance arrived at Rp 0.2 = 443 MPa, R = 1635 MPa
and A = 99.7 % [12, 13]. This unique combination of strength and high ductility
resulted in a specific fracture energy Ws = 751 J/cm3 which is probably the
highest ever measured at room temperature [14]. At a stacking fault energy (SFE)
of 31 mJ/cm3 twinning is the major strengthening mechanism. The carbide/nitride
precipitation was studied in great detail by experiment and simulation as well as
the implications for welding [13].

1.3 Aim

All development starts from the respective state of art. For high strength austenitic
steels it was e.g. compiled in the proceedings of several HNS conferences and a
book on HNS [1]. In addition the previous work on HIS of moderate interstitial
content was taken into account. Starting from this basis our aim is to explore the
potential of HIS and develop advanced grades centering on the following
objectives:
(a) Strive for a really high interstitial content of C ? N, e.g. between 0.8 and 1.1
mass %, to boost the strength of homogeneous austenite.
(b) Balance the effects of C ? N on strength, embrittling precipitation and cor-
rosion, paying special attention to the C/N ratio.
(c) Rely on melting and solidification at normal pressure of air to avoid costly
pressure or powder metallurgy.
(d) Deal with manufacturing on an industrial scale, e.g. with melting, casting,
forging, heat treatment.
(e) Select HIS for special applications.
In all, it is the key aim of this book to come up with manufacturable and
applicable new steels with a unique combination of properties, as strength,
toughness, wear and corrosion resistance as well as non-magnetisability.
4 1 Introduction

Fig. 1.1 Multiscale


approach of investigating
high-interstitial austenitic
steels from the electron
structure to components.
Examples are: ab initio
calculations,
CESR = electron spin
resonance, Mössbauer
spectroscopy,
TEM = transmission
electron microscopy,
LOM = light optical
microscopy, EDX = energy
dispersive X-ray analysis and
EBSD = electron
backscattered diffraction in a
scanning electron microscope
(SEM), tensile tests

1.4 Procedure

To reach this engineering goal the competence of metal physics and materials
technology is combined.
The development starts with a survey on the constitution of the Fe–Cr–Mn–C–
N alloy system to come up with suitable HIS grades. This task is carried out by
thermodynamic simulations which not only give the type, amount and composition
of phases in dependence of the alloy composition, temperature and nitrogen
pressure, but also allow to predict process parameters for melting, casting, forging
and heat treatment.
Next, the structure of selected HIS grades is studied in depth, starting from the
electron structure of austenite and the atomic nano-scale distribution of the ele-
ments involved. Transmission electron microscopy (TEM) is e.g. employed to
investigate structural changes provoked by mechanical loading. Scanning electron
microscopy (SEM) and light optical microscopy (LOM) help to analyse e.g. the
progress of fracture, wear and corrosion. High carbon, high nitrogen or low
interstitial austenitic reference steels are included for comparison. The attempt is
made to describe the structure of HIS from the electron scale up to the scale of
components (Fig. 1.1).
Key properties of selected HIS grades are measured and traced back to struc-
tural features of austenite as e.g. concentration of free electrons, short range atomic
order, stacking fault energy, transformation or twinning induced plasticity (TRIP
and TWIP), grain boundary precipitation. This knowledge on properties and
structural background is to provide a sound basis for applications.
References 5

References

1. Gavriljuk VG, Berns H (1999) High Nitrogen steel. Springer, Berlin


2. Schaeffler AL (1949) Constitutional diagram for stainless steel weld metal. Metal progress
56:680–680B
3. Shanina BD, Gavriljuk VG, Berns H, Schmalt F (2002) Concept of a new high-strength
austenitic stainless steel. Steel Res 73:105–113
4. Schmalt F, Berns H, Gavriljuk VG (2004) Mechanical properties of a stainless austenitic
CrMnCN steel, Steel Grips 2. Suppl High Nitrogen Steels 2004:437–446
5. Schmalt F (2004) Nutzung der Löslichkeit von C ? N in nichtrostenden Stählen, doctoral
thesis Ruhr University Bochum, see also Fortschr. Ber. (2005) VDI 5-702, VDI Verlag,
Düsseldorf
6. Gavriljuk VG, Berns H (1999) Precipitates in tempered stainless martensitic steels alloyed
with nitrogen, carbon or both, Trans Tech Publ. Zürich Mat Sci Forum 318–320:71–80
7. Bernauer J, Speidel MO (2003) Effects of carbon in high-nitrogen corrosion-resistant
austenitic steels. In: Speidel MO, Kowanda C, Diener M (eds) Proceeding HNS 2003, vdf
Hochschulverlag AG, ETH Zürich, pp 159–168
8. Bernauer J, Saller G, Speidel MOl (2004) Combined influence of carbon and nitrogen on the
mechanical and corrosion properties of Cr-Mn steel grades, Steel Grips 2, Suppl High
Nitrogen Steels, pp 529–537
9. Bernauer J (2004) Einfluss von Kohlenstoff als Legierungselement in stickstofflegierten
Chrom—Mangan Stählen, doctoral thesis ETH Zürich, No. 15457
10. Müller R, Weintz R (1998) Ventilwerkstoffe für Verbrennungsmotoren, Materialwiss.
u. Werkstofftechn 29:97–130
11. Bonaziz O, Allain S, Scott CD, Cugy P, Barbier D (2011) High manganese austenitic
twinning induced plasticity steel: a review of the microstructure properties relationships,
Current Opinions in Sol. State Mat Sci 15:141–168
12. Mujica Roncery L, Weber S, Theisen W (2010) Development of Mn-Cr-(C-N) corrosion
resistant twinning induced plasticity steel: Thermodynamic and diffusion calculations,
production, and characterization. Metall Mat Trans 41A(10):2471–2479
13. Mujica Roncery L (2010) Development of high-strength corrosion-resistant austenitic TWIP
Steels with C ? N, doctoral thesis, Ruhr University Bochum
14. Berns H, Gavriljuk VG (2007) Steel of highest fracture energy, Key engineering materials,
vols 345–346. Trans Tech Publications, pp 421–424
Chapter 2
Constitution

2.1 General Remarks

The constitution of high interstitial steels describes the state of atomic order in
thermodynamic equilibrium. It depends on the three variables of state: concen-
tration of alloying elements in iron, temperature and pressure. A region that is in
the same state of order is known as a phase. Nitrogen is a volatile element and
therefore the gas phase and especially the partial pressure of N2 has to be taken
into account. The liquid phase appears during melting, solidification and welding.
Of the solid state phases austenite, d-ferrite, carbides, nitrides and sigma phase are
to be expected. To cope with such a complex alloy system the commercial soft-
ware program THERMO-CALCTM, version R with TCFE4 database [1] was used
to calculate the constitution of multi-component HIS.
It depends on experimental data and theoretical models covering a wide range
of steel compositions by minimizing the Gibb’s free energy [2]. The program
provides phase diagrams as isothermal or isoplethal sections through a system.
Also the mole, mass or volume fraction of phases in a given steel may be plotted
over the temperature. In addition the chemical composition of each phase is
available. In the high temperature range from solidification to solution annealing,
which are of practical importance, less deviation from the calculated equilibrium is
to be expected than at lower temperatures. However during quenching the kinetics
of precipitation are of interest only. It is shown in the next chapter that the atoms in
homogeneous austenite are not necessarily distributed evenly, but that microseg-
regation (mm range) or short range atomic decomposition (clustering, nm range)
may cause a chemical inhomogeneity which is not covered by the calculations.
Tramp elements are not considered to reduce the computing time except for
manufactured grades.
Limited experimental verification was in good agreement with the simulation.
Therefore the calculated phase diagrams are taken as a reasonable guideline to
reveal tendencies.

H. Berns et al., High Interstitial Stainless Austenitic Steels, 7


Engineering Materials, DOI: 10.1007/978-3-642-33701-7_2,
Ó Springer-Verlag Berlin Heidelberg 2013
8 2 Constitution

Fig. 2.1 Isoplethal phase diagram of Fe-18Cr-18Mn in dependence of (a) Carbon content.
b Nitrogen content. c C ? N content at constant C/N = 0.6. d C content at constant N = 0.6
mass %, i.e. increasing C/N. Shaded area = homogeneous austenite

2.2 Variation of Interstitial Content

To start with, the alloy system Fe-18Cr-18Mn is studied in respect to additions of


C, N or C ? N, because the commercial grade Cr18Mn18N0.55 has been suc-
cessfully manufactured and applied. In fact, nitrogen opens up a wide field of
homogeneous austenite (shaded area in Fig. 2.1b) while carbon provides austenite
only in combination with carbide M23C6 and/or ferrite (Fig. 2.1a). This is to say
that a steel alloyed with 18 mass % Cr to make it stainless and with the same
content of Mn to stabilize austenite cannot be strengthened by interstitial carbon
without precipitation of carbides which consume chromium and impair toughness.
In contrast the high solubility of nitrogen in austenite during solution annealing
(Fig. 2.1b) offers intensive strengthening within the target range of 0.8–1.1
mass % N without detrimental precipitation of M2N nitrides. However, the N2
isobars reveal that a partial pressure pN2 = 0.8 bar nitrogen in air is not sufficient
2.2 Variation of Interstitial Content 9

to dissolve 0.8–1.1 mass % N in the melt. At lower contents solidification passes


through a regime of ferrite whose solubility for nitrogen is much lower than that of
austenite. The resulting degassing may cause boiling or foaming of the solidifying
melt or pores in the solid material. Pressure metallurgy is a means of raising the
nitrogen content in the melt to a level which assures a fully austenitic solidifica-
tion. It is important to note that pN2 governs the uptake of nitrogen to the equi-
librium content, but that the formation of bubbles at the beginning of degassing
depends on the total pressure including the ferrostatic part of the melt which
somewhat eases the problem in respect to the calculations.
The concept of alloying with C ? N is now applied to the basic composition of
steel Cr18Mn18N0.55 to avoid pressure metallurgy (Fig. 2.1c). The C ? N con-
tent is plotted along the abscissa at a selected C/N ratio in mass %. The atomic
ratio is smaller by [1-(12/14)] 100 & 14 %. The shaded phase field of homo-
geneous austenite is sufficiently large to dissolve 0.8–1.1 mass % C ? N at a
solution anneal temperature TSA of e.g. 1100 °C. As in Fig. 2.1b the L-F-A triple
point remains just below 0.9 mass % of interstitials but the corresponding pN2 is
lowered from 1.5 to 0.8 bar and a fully austenitic solidification is to be expected
(Fig. 2.1c). This means that the partial replacement of N by C is as effective in
fending off d-ferrite but comes with less volatility of the remaining nitrogen. The
solidus temperature TS is lowered by carbon, but the temperature of beginning
precipitation TP is hardly changed within the interstitial range of interest. The type
of precipitate turns from M2N to M23C6 though, which is followed by M2N at
lower temperatures.
The addition of carbon to commercial steel Cr18Mn18N0.55 will reduce the
content of d-ferrite during solidification and allow of more nitrogen in the melt.
Therefore Fig. 2.1d starts from 0.6 mass % N and directly demonstrates the effect
of carbon implying an increase of the C/N ratio. The L-F-A triple point is located
at about 0.25 mass % C and the corresponding temperatures TS and TP are about
1320 and 950 °C respectively.
At higher carbon contents the range of homogeneous austenite between TS and
TP is narrowed and with it the interval for hot working and solution annealing. At
the example of the commercial grade Fig. 2.1c and d clearly demonstrate that the
C/N concept is suited to increase the interstitial content by melting at normal
pressure of air, but that there is a limit to the optimal carbon content.

2.3 Effect of C/N Ratio

An isothermal section through the Fe-18Cr-18Mn-C–N system at TSA = 1100 °C


outlines the shaded target area of homogeneous austenite (Fig. 2.2). It is encased
by ferrite to the left, M2N to the right and M23C6 above. The carbon content of
austenite grows from (mass %) 0.3 at 0.2 N to almost 0.8 at 1.1 N. This underlines
the beneficial effect of jointly alloying C ? N in respect to the intended increase of
the interstitial content in austenite. It may be raised further by a higher TSA,
10 2 Constitution

Fig. 2.2 Isothermal phase


diagram at 1100 °C of
Fe-18Cr-18Mn-C–N. The
dotted lines represent
different C/N ratios. Shaded
area = homogeneous
austenite

Fig. 2.3 Isoplethal phase


diagram of Fe-18Cr-18Mn in
dependence of the C/N ratio
at C ? N = 1 mass %. At
(C/N)op the shaded phase
field of homogeneous
austenite extends to the
lowest temperature Top

because the phase field of austenite is expanded at the expense of precipitates, but
gives way to ferrite on the left.
The dotted lines in Fig. 2.2 represent C/N ratios. At C/N = 1 the respective line
cuts the austenitic phase field at the low interstitial end, while at 0.7 it touches the
point of highest interstitial solubility. At 0.5 it runs in parallel to the A/A ? M23C6
boarder. The results of this isothermal plot suggest C/N \ 0.7. In an isoplethal
section through the Fe-18Cr-18Mn system the C/N ratio is varied by plotting C and N
in opposite directions along the abscissa at a constant content of C ? N = 1 mass %
(Fig. 2.3). It is evident that the phase field of homogenous austenite extends to the
lowest temperature Top at (C/N)op = 0.41 at which M2N and M23C6 start to pre-
cipitate simultaneously. The index ‘‘op’’ refers to optimal conditions in respect to
2.3 Effect of C/N Ratio 11

Fig. 2.4 Isothermal phase


diagram at 1100 °C of
Fe–Cr-Mn-0.3C-0.6 N.
Shaded area = phase field
of homogeneous austenite
with C18 mass % Cr,
C ? N = 0.9, C/N = 0.5

retarding the begin of precipitation during quenching from TSA. To the left of (C/N)op
M2N starts to precipitate at TP [ Top and to right this holds true for M23C6. In the
middle part of the C/N range a fully austenitic solidification prevails.

2.4 Variation of Substitutional Content

The basic substitutional elements are chromium and manganese. Molybdenum and
copper are of interest in respect to corrosion resistance. Tramp elements can affect
the constitution.

2.4.1 Chromium and Manganese

The influence of these elements on the constitution is demonstrated by an iso-


thermal section at TSA = 1100 °C and a C ? N content within the target range
(Fig. 2.4). The plot suggests a wide phase field of homogeneous austenite of which
more than half does not apply, if a minimum content of 18 mass % Cr is chosen to
promote corrosion resistance. The lower the manganese content the closer the
alloys come to superheated stainless tool steel with retained austenite and little
ductility. In addition a partially ferritic solidification and a loss of nitrogen is to be
expected. The higher the manganese (and chromium) content, the lower the
concentration of free electrons [3]. Therefore the Mn content in Fig. 2.5a is varied
only moderately in the range of 18 ± 5 mass %. Unexpectedly ferrite is stabilised
by Mn at 1100 °C and M23C6 as well. However, the austenitic phase field grows at
the expense of M2N. It shrinks in respect to M23C6 by a higher C/N ratio
(Fig. 2.5b) and by a lower temperature (Fig. 2.5c) which may entail a shift of the
12 2 Constitution

Fig. 2.5 Isothermal phase diagrams showing the effect of Cr and C ? N content and of (a) Mn
content. b C/N ratio. c Temperature. Shaded phase field = homogeneous austenite, dotted
line = C/N = 0.7 at 1050 °C, (+) = target composition

target point (+) at (mass %) 18 Cr and 1 (C ? N) from the A to the A ? M23C6


phase field. As expected, Cr stabilises ferrite and C ? N are required to obtain
austenite which loses ground to M23C6 and the more so the higher C/N (Fig. 2.5b).
After this view on the constitution at solution anneal temperature TSA in
Figs. 2.4 and 2.5 the situation at higher and lower temperatures is of interest. At
the example of two steels it is demonstrated that an increase from 13 to 23 mass %
Mn lowers the temperature range of solidification but prevents ferrite and raises
the temperature of beginning N2 gas evolution (Fig. 2.6). At temperatures below
TSA manganese enhances the precipitation of M23C6 at the expense of M2N as
already visible in Fig. 2.5a. The promotion of r-phase by Mn is confined to such a
low range of temperature that it is likely to be subdued during quenching. The
2.4 Variation of Substitutional Content 13

Fig. 2.6 Phase fraction of two steels with (mass %) 18Cr and 1 (C ? N) at C/N = 0.6 but
different Mn content in dependence of temperature, shaded area = homogeneous austenite

Fig. 2.7 Phase fraction of steel Cr18Mn18(C ? N)1 at two different C/N ratios in dependence of
temperature, shaded area = homogeneous austenite

jump from 13 to 23 mass % Mn narrows the regime of homogeneous austenite on


either side (Fig. 2.6). Thus 18 mass % Mn seem to be a good compromise between
high TS, low TP and suppression of ferrite as well as gas. The respective steel
Cr18Mn18(C ? N)1 is now analysed as to the influence of the C/N ratio
(Fig. 2.7). To stay below (C/N)op (Fig. 2.3) would mean to give away interstitial
solubility and strength. A mole fraction of 1 would correspond to C/N = 0.857
which according to Fig. 2.3 would require TSA [ 1100 °C. As for the higher Mn
content in Fig. 2.6, the higher C/N ratio narrows the range of homogeneous aus-
tenite on both sides (Fig. 2.7). The results suggest not to exceed these limits of the
C/N range.
14 2 Constitution

Fig. 2.8 Isoplethal phase


diagram of Fe-20Cr-18Mn-
0.6N–C, shaded
area = homogeneous
austenite

Fig. 2.9 Isoplethal phase


diagram of Fe-18Cr-18Mn-
0.6N-0.25C-Mo

In Fig. 2.4 chromium contents below 18 mass % were already excluded


because of corrosion resistance. Contents above this level are likely to improve
this property but promote ferrite and M23C6. This becomes immediately evident if
Fig. 2.8 is compared with Fig. 2.1d. The shaded phase field of austenite with
20 mass % Cr is reduced but still allows steels in the upper interstitial target
range.

2.4.2 Molybdenum and Copper

Alloying stainless steels with molybdenum is a common measure to impede pitting


corrosion [4]. As this element is a carbide former and a ferrite stabilizer, the
question is what content is permitted in austenitic CrMnCN steels. The influence
2.4 Variation of Substitutional Content 15

Fig. 2.10 Isoplethal phase diagrams of Fe-18Cr-18Mn-2Mo in dependence of a C ? N content


at C/N = 0.6. b C/N ratio at C ? N = 1, the pairs of (C/N)op and Top are marked by (+) for
different Mo contents

Fig. 2.11 Isoplethal phase


diagram of Fe-18Cr-18Mn-
0.6N-0.25C-Cu

of Mo on the constitution is depicted in Fig. 2.9 at the example of steel


Cr18Mn18N0.6C0.25. The gas phase is shifted to higher temperatures as Mo
lowers the activity of nitrogen in the melt. Ferrite is stabilised to lower temper-
atures but—up to & 2 mass % Mo—not below &1300 °C which allows hot
working in the range of homogeneous austenite. The temperature TP of beginning
precipitation (M23C6 followed by M2N) is hardly raised and at 2 mass % Mo stays
just below 1000 °C. The temperature of r precipitation is raised, though. The
addition of B2 mass % Mo to the above steel appears to be feasible.
Starting from the previous example, the influence of C ? N and C/N is
investigated next (Fig. 2.10). Compared to Fig. 2.1c the austenitic phase field in
16 2 Constitution

Fig. 2.12 Isoplethal phase diagram of Fe-18Cr-18Mn-2Cu in dependence of a C ? N content at


C/N = 0.6. b C/N ratio at C ? N = 1, the pairs of (C/N)op and Top are marked by (+) for
different Cu contents

Fig. 2.10a is reduced by ferrite to the left and liquid above. In respect to Fig. 2.3
Top and (C/N)op in Fig. 2.10b are changed only moderately by up to
4 mass % Mo.
Stainless steels are alloyed with copper to reduce general corrosion e.g. in non-
oxidising acid solution [4]. In contrast to molybdenum, copper stabilises austenite
and raises the activity of interstitials. This is reflected in Fig. 2.11 indicating a
steep rise of TP. At CuG & 1.8 mass % the evolution of N2 gas ends the range of
homogeneous austenite. At this Cu level the range of 0.8–1.1 mass % C ? N
leads to an austenitic solidification and a reasonable TSA (Fig. 2.12a). The pairs of
(C/N)op and Top at 1 mass % C ? N increase considerably with the Cu content
(Fig. 2.12b). While the addition of 2 mass % Mo or Cu seem to be feasible, joint
alloying of both elements to this level leads to a dramatic shrinkage of the au-
stenitic phase field which would make it difficult to process such a steel.

2.4.3 Tramp Elements

Small quantities of the strong carbide and nitride formers vanadium, niobium and
titanium may result in MX precipitates. In view of the high interstitial content of
HIS these precipitates hardly dissolve at TSA and would represent an additional
phase.
Silicon raises the activity of C and N and thereby promotes precipitation. This is
reflected for Fe-18Cr-18Mn-1(C ? N) by a shift of the pairs (C/N)op and Top from
0.408 and 971 °C at zero Si to 0.486 and 1026 °C at 0.5 mass % Si and further to
2.4 Variation of Substitutional Content 17

0.560 and 1075 °C at 1 mass % Si. To avoid a detrimental increase of Top by up to


100 °C it is recommend to not fully exploit the range of Si B 1 % given e.g. in
EN10088 for austenitic steels, but to keep its content as low as possible.

2.5 Selection of Steels

The phase field of homogeneous austenite is the target area of HIS to be reached
by solution annealing and preserved by quenching. An austenitic solidification is
desirable to transfer nitrogen from the melt to the austenite without degassing. To
keep the concentration of free electrons high and with it the ductile metallic
character of interatomic bonding a reduction of substitutional alloy content,
namely of Cr, Mn, Mo, would be helpful (see Sect. 3.1.3). However, the experi-
ence with HIS Mn17Cr15N0.43C0.39 and general knowledge on stainless steel
speak for 18 mass % Cr. As shown above, 18 mass % Mn are a reasonable match
to avoid ferrite during solidification.
The interstitial content is aimed at C0.8 mass % C ? N to boost strength. In
view of experience with HNS and the calculations above the envisaged upper limit
of 1.1 mass % C ? N is confirmed to avoid problems during hot working and heat
treatment. As to the C/N ratio, one has to start from the soluble content of nitrogen
in alloys with 18 mass % of Cr and Mn each at normal pressure of air. It slightly
depends on the carbon content, but 0.6 mass % N seems to be a fair value to start
with. Combining the C ? N content and the C/N ratio in mass % we arrive at
C ? N = N [(C/N) ? 1]. Inserting N = 0.6 and (C/N)op = 0.41 the C ? N
content is 0.85. Higher interstitial contents have to rely on more carbon which
entails C/N [ (C/N)op and TP [ Top.
Based on these constitutional considerations three steels with 18 mass % of Cr
and Mn each were selected: one at the lower end of the C ? N target range, one at
the upper end and one in the middle. These three new HIS were molten and hot
worked on an industrial scale and designated according to their C ? N content,
times 100, i.e. CN85, CN96 and CN107 (Table 2.1).
To these Mo and Cu were added by remelting a smaller batch. A series of HIS
with 0.65–1.15 mass % C ? N was produced as castings. In accordance with
European standards the designation is preceeded by ‘‘G’’. A few reference steels
are listed of which CrNi represents a standard low interstitial grade, MnC a high
carbon Hadfield steel and CrMnN a high nitrogen steel. The grades MnCr82 and
MnCr70 are reference HIS of lower chromium and interstitial content. For the
investigation of structure (Chap. 3) and properties (Chap. 4) premachined speci-
mens were solution annealed, quenched and machined to final size. Details of HIS
manufacture are given in Chap. 5.
18

Table 2.1 Chemical composition in mass % of the austenitic steels investigated


No. designation C N P S Cr Mn Si Mo Cu Ni C?N C/N
1 CN85 0.26 0.59 0.018 0.001 18.3 18.5 0.26 0.04 – 0.26 0.85 0.44
2 CN96 0.344 0.614 0.021 0.002 18.2 18.9 0.30 0.06 –(a) 0.34 0.96 0.56
3 CN107 0.489 0.578 0.026 \0.001 18.8 18.9 0.43 0.07 –(a) 0.40 1.07 0.85
4 CN94Mo1 0.324 0.620 0.025 \0.001 17.9 19.0 0.23 0.96 – 0.32 0.94 0.52
5 CN103Mo1 0.452 0.582 0.027 \0.001 18.3 18.6 0.42 0.94 – 0.38 1.03 0.78
6 CN96Cu2 0.370 0.594 0.023 \0.001 17.6 19.3 0.18 0.04 1.90 0.28 0.96 0.62
7 GCN65 0.033 0.616 0.017 0.002 19.9 18.0 0.33 0.04 0.398 0.47 0.65 0.05
8 GCN88 0.228 0.654 0.017 0.001 20.2 18.0 0.28 0.04 0.316 0.39 0.88 0.35
9 GCN98 0.400 0.583 0.018 \ 0.001 20.1 18.1 0.14 0.04 0.306 0.40 0.98 0.69
10 GCN115 0.512 0.641 0.018 \0.001 19.9 18.0 0.33 0.06 0.268 0.46 1.15 0.80
11 GCN85 0.256 0.596 0.017 \0.001 18.3 18.4 0.54 0.02 –(a) 0.10 0.85 0.43
12 CrNi 0.004 0.050 0.020 0.022 18.7 1.9 0.57 – – 9.04 0.05 0.08
13 MnC 1.190 0.009 0.090 0.014 0.2 12.1 0.49 – – 0.1 1.20 132
14 CrMnN 0.040 0.880 – – 21.0 23.1 0.30 0.2 – 1.5 0.92 0.05
15 MnCr82 0.387 0.431 0.044 0.007 14.7 17.2 0.48 \0.02 – \0.07 0.82 0.90
16 MnCr77 0.319 0.447 – – 12.0 25.4 – – – – 0.77 0.71
No. 1–6 = Hot worked steel, 7–10 = Centrifugal castings, 11 = Sand casting, a section of which was hot worked ? CN859, 12–16 = Hot worked
reference steels
(a)
V \ 0.07
2 Constitution
References 19

References

1. Software System and Users Guide (2008) Thermo-Calc Software AB SE- 11347 Stockholm
2. Saunders N (1995) Phase diagram calculation for high-temperature structural materials. Phil
Trans Royal Soc London A351:543–561
3. Shanina BD, Gavriljuk VG, Konchitz AA, Kolesnik SP (1998) The influence of substitutional
atoms upon the electron structure of the iron-based transition metal alloys. J Phys Condensed
Matter 10:1825–1838
4. Heimann W, Oppenheim R, Wessling W (1993) Stainless steels. In: Steel, vol 2. Springer,
Berlin, pp 382–422
Chapter 3
Structure

It is generally accepted among metal scientists and engineers that ‘‘structure’’


stands for crystal lattice, lattice defects and their distribution as well as for grain
size. In the solid solutions, the type of solute atoms and their distribution, as well
as precipitates are taken into account.
The aim of this chapter is to show that, in fact, the structure of metals and alloys
starts from localized or free electrons. Under external force and resultant straining,
the atoms are being shifted from their positions, and mechanical response, plastic
deformation or brittle fracture, depend on the character of interatomic bonds. In
comparison with the nuclei, the response of electrons is quicker by many orders of
magnitude. The closed electron shell, so-called ‘‘ion core’’, can be excluded from
the consideration because it does not take part in chemical reactions and, under
straining, can be only slightly deformed, i.e. polarized. Only the external, i.e.
valence electrons are responsible for chemical bonds and deformation behaviour.
According to a modern approach, they reveal an ‘‘itinerant’’ behaviour, which
means that the same valence electrons are sometimes free and sometimes localized
at the atomic sites. However, some permanent part of valence electrons is always
free, and the free/localized electron ratio is the controlling factor of the nature of
metals, particularly in respect to their behaviour under external action. The pre-
vailing localized valence electrons form covalent bonds between the atoms in the
crystal lattice, which causes brittleness because even a slight shift of the atoms
under shear stress in the slip plane leads to the breaking of interatomic bonds. This
is, e.g., the case of the transition metals and alloys of group V and VI in the
periodic table (V, Nb, Ta, Cr, Mo, W). Free electrons are responsible for the
metallic character of interatomic bonds, and the higher their fraction is the more
ductile are metals and alloys.
In relation to phase transformations, valence electrons are responsible for
the height of the energy barrier which has to be overcome by the atoms consti-
tuting a new crystal lattice, either during its nucleation or during their jumps
through the interface between matrix and new phase. Again, the stronger the

H. Berns et al., High Interstitial Stainless Austenitic Steels, 21


Engineering Materials, DOI: 10.1007/978-3-642-33701-7_3,
Ó Springer-Verlag Berlin Heidelberg 2013
22 3 Structure

covalent bonds between the atoms are, the higher this energy barrier is. For this
reason, the transition metals of group V and VI do not reveal any polymorphic
transformations.
Moreover, as will be shown in this chapter, the control of interatomic bonds and
the free/localized electron ratio affects short-range atomic order in multicompo-
nent solid solutions and, for this reason, their thermodynamic stability.
It is a privilege of metallurgists that iron belongs to metals with a rather high
part of free electrons. Alloying it with the elements located to the right of iron in
the periodic table (Ni, Co, Cu) increases the concentration of free electrons thereby
enhancing the metallic character of interatomic bonds and assisting ductility.
Elements to the left of iron (Mn, Cr, Mo, V etc.) act in the opposite direction (see
about details [1]).
The alloyed steel is a complicated engineering material and the knowledge of a
fundamental correlation between the interatomic bonds on the one hand and
thermodynamic stability as well as mechanical and chemical properties on the
other should be useful for their deliberate and successful design. The current
chapter is an attempt to approach this ambitious task.

3.1 As-Quenched

The initial microstructure of homogeneous austenite is achieved by solution


annealing and quenching in water. At first this structure will be discussed followed
by an investigation on the effect of straining presented in Sect. 3.2.

3.1.1 Electron Structure: Calculated and Measured

The proposed concept of alloying the austenitic steels with carbon ? nitrogen is
based on the ab initio calculations of the electron structure where only the type of
the crystal lattice and the charge of the nuclei of constituting atoms are set as
starting points. The interatomic distances are obtained from the minimum of the
calculated cohesive energy and all further properties are calculated without any
assumptions. The construction of the calculated atomic configuration consisting of
32 substitutional and 2 interstitial atoms is shown in Fig. 3.1. The corresponding
chemical composition amounts to (atom%) 58.8 Fe, 23.5 Mn, 11.8 Cr and 5.9 of
interstitials. This composition was chosen for the calculations as a compromise
between the requirements to keep up the permanent translation of calculated
configuration in three orthogonal directions and to minimize the total number of
calculated atoms in order to reduce the calculation time.
The method of full potential linearized augmented plane waves (FLAPW) based
on the density functional theory [2, 3] and the program package Wien2k [4] were
used for the calculations. Of the electron properties which can be obtained using
3.1 As-Quenched 23

Fig. 3.1 Atomic configuration Fe20Mn8Cr4 with C2, N2 or C1N1 interstitial atoms chosen for ab
initio calculations of the electron structure. The chosen symmetry of the configuration allows its
translation along three orthogonal directions

Fig. 3.2 Density of electron states in the valence electron band of the calculated solid solutions.
The change of the DOS in the vicinity of the Fermi level is shown in the upper insert

these calculations, the density of electron states, DOS, at the Fermi level, EF, is
particularly important because it determines the interatomic bonds, namely the
concentration of free electrons and the thermodynamic stability of structures.
According to Fermi statistics for free electrons, their distribution on the energy
scale is so that only electrons at the Fermi level can change their energy, i.e. be
really free and contribute to heat capacity, conductivity etc.
24 3 Structure

Figure 3.2 shows how nitrogen, carbon or carbon ? nitrogen affect the density
of electron states in the upper energy band of the CrMn austenitic steel.
The electron structure of the iron is described as 1s22s22p63s23p63d64s2, where
1–4 are the main quantum numbers which determine the energy of the corre-
sponding electron band, whereas s, p and d characterize the space symmetry of
electrons, namely a spherical symmetry for s-electrons and a leaf symmetry for p-
and d-electrons. The two latter differ in their angular moments. The electron
energy states 1s22s22p63s23p6 form a closed shell, the ion core, which remains
unchanged at any chemical reactions, whereas 3d64 s2 electrons belong to the
valence band and reveal a so-called itinerant behaviour. As mentioned above,
sometimes they are free, sometimes localized. However, any time a definite
constant part of these electrons are free. The highest valence electron energy
corresponds to the Fermi level, of which the width is proportional to kT, where k is
the Boltzmann constant. The Fermi level in Fig. 3.2 is located at the co-ordinate
origin. All energy states located below the Fermi level are occupied. Only the
electrons within the kT interval can change their energy. Just these electrons
determine the electron capacity, conductivity and thermodynamic stability of
phases.
Alloying affects the distribution of valence electrons on their energy states. The
addition of new elements does not just shift the Fermi level along the rigid band.
Each new chemical composition creates its own valence electron energy band.
One can see in Fig. 3.2 that interstitial elements create so-called bound states at
the bottom of the valence electron band of the FeCrMn substitutional solid solu-
tion, shift the electron levels on the energy scale and change their amplitude. As
follows from the insert in the upper left corner of Fig. 3.2, interstitial elements
change the population of states at the Fermi level at which electrons can change
their energy and, therefore, contribute to a change in the properties of materials.
Alloying with carbon decreases the DOS, whereas nitrogen increases it. The partial
substitution of carbon by nitrogen leads to a further increase of DOS.
The most significant consequence of the effect of interstitial elements on the
DOS at the Fermi level presented in Fig. 3.2 is the expected change in the con-
centration of free electrons, i.e. in the character of interatomic bonds. Judging on
the results presented in Fig. 3.2, one can predict that carbon should decrease the
concentration of free electrons, whereas nitrogen and, particularly, car-
bon ? nitrogen should increase it.
A qualitative presentation of this effect is given in Figs. 3.3 and 3.4. A striking
difference in the distribution of valence (free) electrons in the vicinity of carbon
and nitrogen atoms is observed (compare Figs. 3.3a and b), which suggests that
nitrogen atoms migrate through the crystal lattice being surrounded by clouds of
free electrons, whereas carbon atoms are expected to have a shortage of electrons
in their vicinity as compared with free atoms. It is relevant to note that these
theoretical results confirm the old experimental observations by Seith et al. of the
electrotransfer in austenitic steels with carbon [5] or nitrogen [6] according to
which the carbon atoms are positively charged, whereas the nitrogen ones carry a
negative electric charge.
3.1 As-Quenched 25

Fig. 3.3 Spatial distribution of the valence electron density el/a.u.3 along the (110) plane in the
fcc lattice of solid solutions. a Fe20Mn8Cr4C2. b Fe20Mn8Cr4N2 and c Fe20Mn8Cr4C1N1. 1
a.u. = 0.529 Å

Fig. 3.4 Projection of the spatial distribution of the valence electron density on the (100) plane.
The electron density in the interatomic space increases in the sequence of a Fe20Mn8Cr4C2 ?
b Fe20Mn8Cr4N2 ? c Fe20Mn8Cr4C1N1. The plane (100) is chosen in order to demonstrate how
the alloying with C ? N increases the density of free electrons even in the vicinity of the carbon
atoms (compare Figs. a and c). The nitrogen atoms are not present on the (100) plane in the C1N1
composition

The spatial valence electron distribution around C and N atoms in case of


alloying with C ? N (Fig. 3.3c) repeats main features of steels with C or N.
However, the two-dimensional projections of the spatial valence electron dis-
tribution allow to estimate the electron distribution within the interatomic space,
and it is clearly seen from Fig. 3.4 that the electron density in the space between
26 3 Structure

Fig. 3.5 A scheme for CESR


measurements. a Free
electron energy levels split
under an applied magnetic
field H. b Precession of the
electron spin along the field
and in opposite direction (the
ground and upper energy
levels in Fig. 3.5a,
respectively). c Absorbed
microwave energy P spent for
transfer of electrons from the
ground to the upper energy
level. For convenience, it is
presented as a derivative of
the applied field, because this
visualizes the asymmetry of
the signal, a feature of the
signal from free electrons
caused by their migration for
the spin relaxation time

the atomic sites is highest for the steel alloyed with C ? N (Fig 3.4c). And, of
course, it is higher in the N steel in comparison with the C steel.
This suggests an increase in the concentration of free electrons in austenitic
steels with nitrogen or carbon ? nitrogen. It is particularly important that, in case
of alloying with carbon ? nitrogen, the spatial distribution of free electrons is
more homogeneous and their density increases also in the vicinity of carbon atoms
(compare Figs. 3.4a and c).
The occurrence of excessive free electrons around the nitrogen atoms allows to
interpret some unordinary mechanical properties of austenitic nitrogen steels,
namely quasi-cleavage at low temperatures and high strain rates, which will be
discussed in Sect. 3.2.3.1.
Experimental studies confirm the results of theoretical calculations. The con-
centration of free electrons was measured by conduction electron spin resonance,
CESR, [7–9], the essence of which is shown in Fig. 3.5.
3.1 As-Quenched 27

Table 3.1 Effect of C, N and Composition ne, 1022 -3


cm
C ? N (mass %) on the
concentration of free Cr13Mn18N0.7 1.1
electrons ne in steel Cr13Mn18C0.25N0.25 2.9
Cr13Mn18 Cr13Mn18C0.4N0.4 3.7

Table 3.2 Concentration of Steel ne, 1022 cm-3


free electrons ne in new HIS
CN107 2.0
CN96 2.9
CN85 2.8

Under an applied magnetic field (Fig. 3.5a), the degenerated energy level of
free electrons with magnetic quantum number ‘ is split into two energy states
with spin -‘ (spin orientation along the applied field) and +‘ (spin orientation in
the opposite direction). The microwave energy spent for a transfer of electrons
from the ground to the upper energy level, i.e. for overturning the spin into the
direction along the field (Fig. 3.5b), is proportional to the concentration of free
electrons, which is calculated using a reference sample with a known spin con-
centration. Three CESR signals in Fig. 3.5c belong to the austenitic nitrogen steel
Cr13Mn18N0.7 and two steels of the same basic composition where nitrogen is
partly replaced by carbon. Thus, the combined alloying with nitrogen ? carbon
enhances the signal from free electrons, i.e. increases the concentration of free
electrons in the austenitic steel as predicted by theoretical calculations. The
quantitative data of free electron concentration are presented in Table 3.1.
Based on the described results, a concept of alloying steels with car-
bon ? nitrogen was developed in [10–18]. Three austenitic steels with (mass %)
18Cr, 18Mn and (i) 0.489C ? 0.578N (CN107), (ii) 0.344C ? 0.614N (CN96)
and (iii) 0.26C ? 0.59N (CN85), see Table 2.1, were chosen for detailed studies.
The obtained data of the free electron concentration are presented in Table 3.2.
The increase of the chromium content from steel Cr13Mn18C0.4N0.4 to steel
CN85 needed for the improvement of corrosion properties, leads to some decrease
in the concentration of free electrons. Nevertheless, it remarkably exceeds that of
the nitrogen steel and, of course, of austenitic carbon steels.
Finally, the measured concentration of free electrons in austenitic steels with
carbon, nitrogen and carbon ? nitrogen is presented in Fig. 3.6. One can see that
carbon does not really change the concentration of free electrons in austenitic
steels. In other words, carbon supplies its valence electrons to the energy levels
below the Fermi level. Alloying with nitrogen increases it up to some critical
value.
One can conclude that some optimum should exist for the nitrogen content in
austenitic steels. The arrows in Fig. 3.6 show how the concentration of free
28 3 Structure

Fig. 3.6 Effect of carbon, nitrogen and carbon ? nitrogen content on the concentration of free
electrons in austenitic steels

electrons increases if, at the same content of substitutional alloying elements, a


part of nitrogen is replaced by carbon.
The combined alloying with C ? N remarkably increases the concentration of
free electrons and shifts its maximum towards higher interstitial contents, which,
in comparison with austenitic nitrogen steels, can be used to increase the content of
interstitials without pressure metallurgy (see Sect. 5.1).

3.1.2 Atomic Distribution

So-called ideal solid solutions, where the atoms do not interact with each other and
the gain in the Gibbs free energy is obtained only due to the increased entropy, do
not really exist in nature. A different electron constitution of the host and solute
metal atoms predetermines different atomic interactions.
Short-range atomic order designates any deviation from the statistical atomic
distribution. The prevailing bonds of the host metal atoms with the solute atoms,
M–S, are characterized by the term ‘‘short-range atomic ordering’’, whereas the
favour for M–M and S–S bonds means ‘‘short-range atomic decomposition’’.
The steel is a multicomponent solid solution, and the segregation of alloying
elements, well known to metallurgists, is the utmost result of short-range
decomposition. High temperature treatments are used in order to reach the more or
less homogeneous state of steels and, in such a way, provide their stability to
precipitation phenomena in the course of technological operations. Nevertheless,
clusters, i.e. accumulations of one kind of atoms, exist even in the liquid alloys,
3.1 As-Quenched 29

which is well known from the studies of rapid quenching from the liquid state (see,
e.g. [19] for a FeSiB alloy).
This is why, except for some special cases, the aim of steel designers is to
enhance the tendency to short-range atomic ordering, in order to reach the highest
possible thermodynamic stability of phases. Because of the interaction of solute
atoms with dislocations and a tendency to precipitation, the effect of atomic dis-
tribution on mechanical and corrosion properties is not less important.
Among the available experimental technique, Mössbauer spectroscopy is suit-
able for studies of short-range atomic order in solid solutions because the change
in the nearest atomic neighbourhood of iron atoms strongly affects the main
parameters of spectra, namely, the hyperfine field at the atomic nuclei, which
causes the splitting of a single line of paramagnetic austenite into six lines of
ferromagnetic martensite proportional to the atomic magnetic moment (so-called
Zeemann splitting), the isomer shift (the shift of the spectrum gravity center
caused by a change of the electron density at the nuclei) and the quadrupol
interaction (e.g. the splitting of the single line of paramagnetic austenite into the
doublet, if the nearest solute atoms cause a local deviation of the crystal lattice
from its cubic symmetry).
For the paramagnetic solid solutions like austenitic steels, the Mössbauer
spectrum consists of a single line and doublets of which the quadrupol splitting is
proportional to the local distortions of the crystal lattice. The study of interstitial
solid solutions is most informative because the interstitial atoms cause higher
distortions in comparison with the substitutional ones. For this reason, we start
with the binary Fe–C and Fe–N austenites. Thereafter, the ternary Fe–C–N solid
solution will be analyzed, which allows to identify important features of the atomic
distribution due to combined alloying with C ? N as a consequence of a change in
the electron structure.
As the atomic distribution in the austenite is inherited by the martensitic phase
because of the diffusionless martensitic transformation, the Zeemann sextet of the
martensite was used for characterization of the atomic distribution in austenitic
steels. This approach is particularly informative because the interstitial as well as
substitutional alloying elements decrease the hyperfine field at the iron nuclei (i.e.
the extent of Zeemann splitting) and this decrease is proportional to the number of
solute atoms as nearest neighbours of the iron ones.
Mössbauer spectra of binary Fe–C and Fe–N solid solutions are presented in
Fig. 3.7 (see also [20, 21]). The samples were prepared using the saturation of pure
iron with carbon or nitrogen in CH4 ? H2 or NH3 ? H2 mixtures at 1150 and
700 °C, respectively. The carbon and nitrogen concentration were determined by
X-ray diffraction. The atomic configurations corresponding to the components in
the spectra are shown to the left.
The Fe–C spectrum consists of a single line belonging to iron atoms Fe0 having
no carbon atoms as nearest neighbours. The doublet comes from iron atoms Fe1
with one carbon atom as nearest neighbour and Fe290 with two carbon atoms in
neighbouring nearest interstitial sites. These two different configurations cause the
30 3 Structure

Fig. 3.7 Mössbauer spectra of binary austenitic solid solutions, atom %. a Fe-9.1 C. b Fe-9.3 N
and c–f corresponding Fe–C(N) atomic configurations

same electric field gradient in the crystal lattice (only its sign is different), which is
displayed by the same quadrupol splitting of the spectrum.
Such 90°-pairs are never met in Fe–N austenite. Instead, in addition to the Fe1
component, the spectrum contains a doublet from the Fe2180 configuration
caused by nitrogen atoms occupying interstitial sites within the second coordi-
nation sphere. Such a dumbbell-like configuration is an element of the ordered
Fe4N c0 -phase.
In order to obtain the values of C–C and N–N interaction energies consistent
with the fractions of atomic configurations derived from Mössbauer spectra, a
modelling of these solid solutions was carried out using the Monte Carlo method
(Fig. 3.8). W1 and W2 are the energies of interaction between any two interstitial
atoms in the first and second coordination spheres, respectively, if one of them is
located in the coordinate origin. The areas marked as C–C and N–N correspond to
the values of C–C and N–N interactions which are consistent with the fractions Fe1
(d), Fe290 (e) and Fe2180 (f) atoms in Fig. 3.7.
It is seen that the carbon distribution in austenitic steels is characterized by a
soft repulsion between C atoms in nearest interstitial sites (a small W1 for the C–C
3.1 As-Quenched 31

Fig. 3.8 Areas of C–C and


N–N atomic interactions
(marked with gray colour)
within the first and second
coordination spheres in the
sublattice of interstitial sites
corresponding to the fractions
of Fe1 (d) Fe2–90 (e) and
Fe2-180 (f) iron atoms
obtained from Mössbauer
spectra in Fig. 3.7

area), so that, along with single carbon atoms, some fraction of carbon pairs
Fe2–90° should exist. However, a hard C–C repulsion (large W2) is revealed for
carbon atoms as neighbours in the second coordination sphere of the interstitial
sublattice. This means that dumbbell-like C–Fe–C configurations Fe2–180° are not
met in the austenitic carbon steels, which makes the existence of an ordered Fe4C
type structure impossible.
In contrast, the N–N repulsion in the first coordination sphere is so hard (large
W1) that nitrogen atoms cannot be nearest neighbours in the austenitic lattice,
whereas the soft repulsion in the second coordination sphere (small W2) allows
N–Fe–N pairs which are clearly identified in the Mössbauer spectra of austenitic
nitrogen steels. This is why the ordered c0 -phase Fe4N exists, and, based on the
quoted studies [20, 21], one can state that carbon atoms in austenitic steels are
prone to form clusters, whereas the distribution of nitrogen atoms is characterized
by short-range atomic ordering.
For studies of atomic distribution in the solid solutions alloyed with C ? N, a
ternary alloy of iron with mass % 0.93C and 0.91N was used [22]. The samples
were prepared by Dr. Rawers, Albany Research Center, USA, and obtained by a
special technique of melting an iron-carbon steel in a hot isostatic pressing (HIP)
furnace under a nitrogen gas pressure of 160 MPa (see [23]). According to the X-
ray diffraction measurements, the alloy consisted of 60 austenite and 40 % mar-
tensite. The C ? N content was 1.57 mass in martensite and 2.40 mass % in
austenite, estimated from crystal lattice dilatation. A Mössbauer spectrum of this
alloy is presented in Fig. 3.9.
The sextets represent the ferromagnetic martensite, whereas the central part of a
single line and a doublet come from the paramagnetic austenite. Two sextets Fea0
and Fea1 belong to atomic configurations of iron atoms with no or one interstitial
atom as nearest neighbour, respectively.
The following features distinguish this spectrum in comparison with those of
Fe–C and Fe–N austenite and martensite. The austenitic part of the spectrum
contains a single line Fec0 of the iron atoms having no interstitials in the nearest
neighbourhood and the doublet Fec1 of the iron atoms with one interstitial atom as
nearest neighbour. In comparison with the spectrum of Fe–N austenite (see
32 3 Structure

Fig. 3.9 Mössbauer spectrum of iron with (mass %) 0.93C and 0.91N

Table 3.3 Abundance of different atomic configurations in Fe-0.93C–0.91N alloy, according to


Mössbauer data
Phase (%) Austenite Martensite
Configuration Fec0 Fec1 ? Fec2-90° Fea0 Fea1 Fea4
Spectrum mode single line doublet sextet 1 sextet 2 sextet 3
(clusters)
Abundance in spectrum 20 18 47 8 7
Normalized abundance 53 47 86 14 –

Figs. 3.7, 3.8), the component from the N–Fe–N dumbbell-like configurations
(Fec2180 atoms) is not present. In comparison with the spectrum of Fe–C mar-
tensite (see e.g. [24]), the component Fea290 with two carbon atoms as nearest
neighbours is also absent. The occurrence of the low intensive sextet Fea4 suggests
that some clusters of interstitial atoms exist in the martensite, which can be a
consequence of the rather slow cooling of the 5 kg ingot.
The obtained results evidence a distribution of interstitial atoms in the studied
Fe–C–N solid solution that is characterized by a tendency to the hard repulsion
within both the first and second coordination spheres in the sublattice of interstitial
sites.
The abundances of different atomic sites in relation to the whole alloy are
presented in Table 3.3 and are normalized to 100 % of the austenite or martensite.
In order to clarify the distribution of interstitial atoms responsible for the
observed abundances of different configurations of iron atoms and interstitial ones,
a Monte Carlo computer simulation was used. The detailed procedure is described
in [20, 21]. Results obtained for the austenitic and martensitic phases are presented
in Fig. 3.10.
3.1 As-Quenched 33

Fig. 3.10 The areas of i–i interaction in two coordination spheres satisfying Mössbauer data of
alloy Fe-0.93C-0.91N. a Austenitic phase, b martensitic phase

For the austenitic phase, the abundances of configurations Fec1 obtained from
Mössbauer studies can be reached only in a narrow range of concentrations, i.e.
Ni/NFe & 0.085, which corresponds to 1.9572.21 mass %. Ni and NFe are the
number of interstitial and iron atoms involved. It can be seen that in case of
Ni/NFe & 0.098, the data look like the Fe–C austenite, and Ni/NFe & 0.09 leads
to the combination of the interaction energy profiles found for C–C and N–N
interactions in binary Fe–C and Fe–N austenites, respectively (see Fig. 3.8). At
smaller concentrations of interstitials only a strong repulsion in two coordination
spheres can lead to the agreement with Mössbauer data, and apparently it is this
situation that prevails in the studied alloy. As a result of the simulation made for
the martensitic phase, only the concentration ratio Ni/NFe & 0.064 provides the
agreement with Mössbauer data of repulsion between interstitial atoms within the
two coordination spheres.
At the higher contents of interstitials, an attraction within the second coordi-
nation sphere should occur, which is, however, not observed in the Mössbauer
spectrum, possibly because of some clustering in martensite resulting from too
slow cooling of the 5 kg ingot after melting and nitrogen saturation (see sextet 3,
correspondingly atomic configuration Fea4 in Fig. 3.9 and Table 3.3).
Summing up, one can state that the values of interaction energies between
interstitial atoms obtained from the Monte Carlo simulation of the Fe–C-N solid
solution correspond to such a distribution of carbon and nitrogen atoms that is not
prone to form clusters within the first two coordination spheres. This result is quite
different from the distribution of carbon atoms in binary Fe–C austenite and mar-
tensite where interstitials can occupy neighbouring interstitial sites (see Fig. 3.7e
for the austenite) and the effective C–C repulsion exists in the second coordination
sphere only. Compared to the data of binary Fe–N alloys where nitrogen atoms can
occupy neighbouring sites in the second coordination sphere, these 1808 N–N
configurations (Fig. 3.7f) are not revealed in the ternary Fe–C–N alloy.
34 3 Structure

Fig. 3.11 Mössbauer spectra of steel Cr15Mo1 alloyed with (mass %) a 0.6C, b 0.62N and
c 0.29C ? 0.35N after solution treatment at 1100°C followed by quenching in water

Such a strong i–i repulsion within the first two coordination spheres should
provide a higher thermodynamic stability of the iron-based Fe–C-N solid solu-
tions, which is consistent with the results of the calculated phase equilibrium (see
Chap. 2, Constitution). Experimental evidence of an increased stability of iron-
based C ? N solutions to phase transformation was obtained by Mössbauer
spectroscopy. The steel Cr15Mo1 alloyed with (mass %) 0.6C, 0.62N or
0.35C ? 0.29N was used in order to compare the stability to martensitic trans-
formation. Mössbauer spectra are shown in Fig. 3.11. The single component
belongs to austenite and the sextet to martensite.
A striking difference between fractions of the retained austenite in these steels
after quenching demonstrates the non-additivity of carbon and nitrogen effects in
austenitic steels. The physical nature of this phenomenon lies in the correlation
between the electron structure, short-range atomic order and thermodynamic sta-
bility of solid solutions. Like binary Fe–C, Fe–N and ternary Fe–C–N austenites,
the alloying of austenitic steels with nitrogen and particularly with car-
bon ? nitrogen increases the concentration of free electrons, which promotes
short-range atomic ordering and, as consequence, increases the thermodynamic
stability of solid solutions.
It is remarkable that the same effect is observed in the tempered martensite.
Dilatometric curves of tempering are presented in Fig. 3.12 for all three steels. The
first, second and third transformation during tempering, which correspond to the
3.1 As-Quenched 35

Fig. 3.12 Dilatometry of as-


quenched Cr15Mo1C0.6,
Cr15Mo1N0.62 and
Cr15Mo1C0.29N0.35
martensites during tempering
at a heating rate of 1.5 K/min

precipitation of an intermediate e-carbide, the decomposition of the retained


austenite and the transformation e ? h to cementite, are clearly identified in
carbon martensite. Because of the increased stability of the retained austenite in
the alloyed steel, the second transformation is shifted to temperatures above that of
cementite precipitation. The precipitation of M7C3 carbide occurs at temperatures
above that of austenite decomposition.
The dilatometric curve of the nitrogen steel is rather smooth, and the effect of
nitride precipitation is shifted to higher temperatures extending over a broad
temperature range. The retained austenite is not decomposed during heating but is
transformed into martensite in the course of subsequent cooling. No tempering
effects can be identified in C ? N martensite by dilatometry.
As follows from Mössbauer studies, this difference in the tempering behaviour
is caused by different atomic distributions (Fig. 3.13). Tempering at 650 °C
removes the interstitial elements from the solid solution. Therefore, the spectra in
Fig. 3.13 are controlled mainly by the distribution of chromium atoms and, to an
insignificant extent, molybdenum atoms. The component Fe0 belongs to the iron
atoms with no substitutional atom as nearest neighbour and it decreases if carbon is
substituted by nitrogen or nitrogen by carbon ? nitrogen.
This is clear evidence that the chromium distribution is more homogeneous in
the C ? N martensite and it remains so even after tempering at rather high
temperatures.
36 3 Structure

Fig. 3.13 Lines V6 (nuclear


transition +1/2 ? +3/2) in
Mössbauer spectra of steel
Cr15Mo1 alloyed with
carbon, nitrogen or
carbon ? nitrogen after
quenching from 1100 and
tempering at 650 °C for 2 h

3.1.3 Chemical Nanoscale Homogeneity

In the previous section, Mössbauer spectroscopy was employed to analyse the


atomic distribution in alloyed steels and to detect changes from short-range
ordering to short-range decomposition, i.e. clustering. Such changes come with a
reduction of the chemical homogeneity, i.e. a higher concentration of alloying
atoms in clusters. It is the aim of this section to investigate their size which
requires methods different from Mössbauer spectroscopy. It will be shown that
alloying with C ? N improves the chemical homogeneity on a nanoscale (see also
[25]). Three steels CN85, CN96 and CN107 together with the Hadfield steel MnC
were chosen for comparative studies (see their composition in Table 2.1).
The following experimental techniques were used: conduction electron spin
resonance (CESR), magnetic measurements and transmission electron microscopy
(TEM) of stacking fault energy (SFE).
As shown in Sect. 3.1.1, the CESR allows to measure the concentration of free
electrons in the paramagnetic objects. At the same time, one can investigate the
interaction between free and localized electrons, which provides quantitative data
about the fractions of single substitutional atoms and their clusters. This is made
possible by measuring (i) the temperature dependence of the magnetic suscepti-
bility derived from the CESR spectra and (ii) the g-factor which stands for the
splitting of the electron energy levels under an applied magnetic field (Fig. 3.5)
and is different for different chemicals elements.
Three electron subsystems contribute to the total magnetic susceptibility of
austenitic steels: free electrons, single atoms of transition metals (d-atoms) and
superparamagnetic clusters.
3.1 As-Quenched 37

The magnetic susceptibility of free electrons vs0, the so-called Pauli suscepti-
bility, does not depend on the temperature. The magnetic susceptibility of isolated
localised d-electrons vd0 changes with the temperature according to the Curie-
Weiss law:
vd0 ¼ C1 =ðT  hp Þ; ð3:1Þ

where C1 and hp are the Curie constant and the paramagnetic Curie temperature,
respectively.
The subsystems of free s-electrons and of isolated localised d-electrons can
exchange their electrons and the contributions of their interaction to the magnetic
susceptibility are described as
vs ¼ vs0 ð1 þ a1 vd0 Þ and vd ¼ vd0 ð1 þ a1 vs0 Þ; ð3:2Þ

where a1 is the parameter of exchange interaction between the free electrons and
isolated localised d-electrons (isolated atoms of d-elements). It is seen that the
ratio of magnetic susceptibilities of s- and d-electron subsystems v-1 r = vs/vd has
to be a linear function of temperature.
The superparamagnetic clusters are formed in austenitic steels by the atoms
having magnetic moments (the atoms of transition metals with a not-filled electron
d-shell like Cr, Ni, Mn etc.). If they are collected in a cluster, its total magnetic
moment M is proportional to the number of atoms in the cluster. These moments
have an occasional orientation in the crystal lattice, however they can be polarized
by the applied external magnetic field.
The clusters of d-atoms contribute to the whole susceptibility and change the
relative parts of s- and d-contributions. The changed value of vs0 is written as
v0s0 ¼ vs0 =ð1  a2 vd2 ðT ÞÞ; ð3:3Þ

where a2 is the parameter of exchange interaction between the free electrons and
clusters, vd2 is the magnetic susceptibility of the superparamagnetic cluster system
which obeys the Langevin law (see [26]):
vd2 ¼ C2  Lðh=TÞ; C ¼ vd2 at T ¼ 1 K; ð3:4Þ

Lðh=TÞ ¼ cothðh=TÞ  T=h; h ¼ MH=kB B;

where h is the Curie temperature of clusters (proportional to the number of atoms


in the cluster), H is an external magnetic field, B is a constant.
The temperature dependence of the total relative magnetic susceptibility v1
r ¼
vs =vd is described by the equation

v1 1 1
r ðT Þ ¼ a1 vs0 þ vs0 vd1 þ a2 vd2 vs0 vd1 : ð3:5Þ

.
In other words, the CESR technique allows to estimate the distribution of
transition metal atoms in solid solutions (see [7–9] about details).
38 3 Structure

Fig. 3.14 Spectrum of


conduction electron spin
resonance in steel CN85. Two
components are marked with
the short dash and dash dot
lines

Measuring the temperature dependence of magnetization provides one with the


information about single d-atoms (the Curie-Weiss law) and superparamagnetic
clusters (the Langevin law) in addition to that given by the CESR measurements.
Both techniques allow to estimate the chemical homogeneity of solid solutions on
a scale of smaller than 5 nm.
We also used measurements of the stacking fault energy in a non-standard
manner to estimate the short-range decomposition in solid solutions. The SFE in
any solid solution depends only on its chemical composition. Some statements
about the dependence of SFE on the grain size or dislocation density do not take
into account the effect of grain boundary segregation on the change of composition
in the grain bulk and the change of dislocation splitting with increasing dislocation
density due to elastic stresses of neighbouring dislocations. In fact, the SFE is
mainly controlled by the electron structure, namely by the density of electron
states at the Fermi level (see e.g. [27]).
For measurements of SFE, the triple node technique was used (see about details
[28]). Thus, these three techniques are complementary for studies of the atomic
distribution in the solid solutions.

3.1.3.1 Free and Localized Electrons: Electron Spin Resonance

Of the four steels studied, the steel MnC steel is the only one for which CESR was
not observed, which indicates a very low concentration of free electrons. As
example, Fig. 3.14 demonstrates the CESR spectrum of steel CN85, recorded at
T = 130 K.
It consists of two main components. The wide spectral line belongs to free
electrons, whereas the second spectral line is a narrow signal typical for localized
paramagnetic centres, i.e. for electrons localized at the atoms.
3.1 As-Quenched 39

Fig. 3.15 a Temperature


behaviour of CESR spectra in
steel CN85 above 85 K and
b Ferromagnetic resonance in
CN85 at temperatures B85 K

The spectra of this steel measured in two temperature regions are shown in
Fig. 3.15. The spectrum at T = 320 K (see Fig. 3.15a) was recorded simulta-
neously with the CESR signal of the reference sample containing 8 9 1014 spins,
which is seen at a magnetic field of H = 0.16 T.
Starting from 85 K, a reversible phase transformation is observed from para-
magnetic to some new magnetic state within a temperature range of about 3 K
(Fig. 3.15b). The broad absorption signal is very intensive and looks like a
ferromagnetic absorption. At temperatures below 83 K, the signal is not remark-
ably changed, the broad spectral line of free electrons disappears, the narrow line
of the localized paramagnetic centres decreases and can be observed in the
background of this broad signal.
This transformation can indicate the transition from the paramagnetic to a
ferromagnetic state or to blocked magnetic moments of superparamagnetic clus-
ters. The temperature dependence of the signal integral intensities is shown in
40 3 Structure

Fig. 3.16 a Temperature


dependence of the integral
intensities for the CESR
signals from free (Ie/I0) and
localized (Iloc/I0) electrons
and b Resonance magnetic
fields Hres,e and Hres,loc in
steel CN85. The dashed lines
represent the calculated
values

Fig. 3.16a in comparison with that of the reference sample. Ie, Iloc are the reso-
nance signal integral intensities of free electrons and those localized at the para-
magnetic atoms. I0 means the integral intensity of the CESR signal of the reference
sample reduced to T = 1 K. Its temperature dependence is described by the
Curie–Weiss law, Iref = I0/T.
As expected, the integral intensity of signal Ie of free electrons does not depend
on temperature. A small contribution of the localized spin absorption to the signal
of free electrons occurs due to the exchange interaction between these two
subsystems.
Based on the obtained data, the following concentration of free electrons, ne,
was obtained for steels CN107, CN96 and CN85, respectively: ne = 2.0 9 1022
cm-3, 2.9 9 1022 cm-3, and 2.8 9 1022 (see Table 3.2).
The concentration of superparamagnetic clusters, i.e. clusters of substitutional
atoms of alloying elements, is found from the temperature dependence of the
resonance fields presented in Fig. 3.16b.
3.1 As-Quenched 41

They cause local magnetic fields on the free electrons. In this case, the tem-
perature dependence of Hres,e of free electrons is described by the temperature
function of the local magnetic field caused by superparamagnetic clusters, i.e. by
the Langevin function L = coth(h/T)-T/h:

Hres;e ¼ Hð0Þ
res;e  B  ðcothðh=TÞ  T=hÞ: ð3:6Þ

Here, H(0)
res,e is the resonance magnetic field for the CESR in the Fe-based alloys in
the absence of any internal local magnetic fields; B = 4pMcl is the maximum local
magnetic field caused by a system of superparamagnetic clusters with magnetic
moment M; h is the energy of an individual superparamagnetic cluster in temper-
ature units of which the magnetic moment under external magnetic field H is M:
h ¼ MH=kB ¼ 2lB HN=kB ; ð3:7Þ
where lB is the Bohr magneton, N is the number of spins in the cluster, kB is the
Boltzmann constant.
The function (3.6) is shown in Fig. 3.16b by the dashed line with the following
values: H(0)
res,e = 340 mT; B = 180 mT; h = 200 K. By substituting h from (3.7),
the number of magnetic atoms in the cluster was estimated to be about 500 atoms,
which corresponds to an average cluster size of about 3 nm. Using the obtained
B = 180 mT and M = 0.9210-17 erg1/2cm3/2, one can estimate the concentration
of superparamagnetic clusters ncl = B/(4pM) = 1.55  1019 cm-3.
Thus, based on the obtained CESR spectra and their analysis, one can conclude
that all studied steels have a high concentration of free electrons with its maximum
in steels CN96 and CN85. At the same time, along with the localized electrons, i.e.
single atoms of substitutional elements, clusters exist in the studied steels.

3.1.3.2 Clusters: Magnetic Measurements

Aiming at a clarification of the unusual behavior of the CESR measurements


below 85 K and the estimation of the size of magnetic non-homogeneities, we
measured the static magnetization of steels CN85, CN107 and MnC within a wide
range of magnetic fields.
Results are given in Fig. 3.17a for steel CN85 at different temperatures and for
steels CN107 and MnC in Fig. 3.17b, c at a temperature of 300 K. A feature of
these data is the absence of magnetic saturation and the paramagnetic character of
magnetization, which means that the transition at 85 K is not a paramagnetic-
ferromagnetic one.
In Fig. 3.17a one can see a negligible temperature dependence of the magnetic
moment. Versus magnetic field H, it was satisfactorily described for all steels by
Eq. (3.8) as a combination of a linear M(H) function and the Langevin function
M = const(cothan(MclH/kT)-kT/MclH), which is typical for paramagnetism
caused by superparamagnetic clusters (see in detail [25]).
For this reason, the curves in Fig. 3.17 were fitted using the function:
42 3 Structure

Fig. 3.17 Magnetic moment versus magnetic field: a steel CN85 at different temperatures.
b fitted curve of magnetic moment at 300 K for steel CN107 and c for steel MnC

Table 3.4 Parameters of the magnetization curves in Eq. (3.8)


Steel T, K M0, emu/g H0, T H1, T
CN107 300 0.12 0.08 7
CN85 300 0.22 0.2 6.7
240 0.23 0.2 6.5
100 0.23 0.2 7.1
50 0.22 0.15 7.1
4 0.24 0.2 7.2
MnC 300 0.01 0.01 5.0

M ¼ M0  ðcotanhðH=H0 Þ  H0 =HÞ þ H=H1 ð3:8Þ


with parameters M0, H0, H1 presented in Table 3.4. The constant M0 is equal to
glBncl/q, where ncl is the concentration of clusters in the sample, g is factor of
spectroscopic splitting, lB is Bohr magneton, q is the density of the alloy.
In the Langevin function (cotanh(H/H0)-H0/H), H0 is determined as kT/
(glBns), where ns is the number of paramagnetic spins in a cluster. It is supposed
that all clusters have approximately the same size. The linear function versus H
can be presented by the Curie-Weiss paramagnetic susceptibility, as well as by the
Van Vleck atomic paramagnetism. In case of C–W susceptibility, H1 is obtained
3.1 As-Quenched 43

Table 3.5 Concentration of superparamagnetic clusters, ncl, and average number of paramag-
netic atoms per cluster ns
Steel ncl, cm-3 ns
CN107 4.6  1019 2570
CN85 8.4 1019 1012
MnC 3.5  1018 20000

from the Curie-Weiss law as 4kT/g2l2BnPC, where nPC is the number of the local
paramagnetic centers, i.e. that of substitutional solute atoms. The estimation of nPC
shows that the value of H1 = 7 T in Table 3.4 could be obtained if the concen-
tration of paramagnetic centers nPC is equal to about 3.4  1022 at T = 300 K and
is by one order larger at T = 4 K, which is impossible.
Thus, H1 does not depend on temperature and, therefore, the C–W paramag-
netism is not justified by the temperature dependence of H1 up to magnetic fields
H = 8 T. The only kind of the paramagnetic system with a linear function M(H)
and not dependent of T is the Van Vleck paramagnetism [26]. All the iron atoms
take part in the formation of the Van Vleck paramagnetic susceptibility. At smaller
fields, the C–W paramagnetic susceptibility becomes actual.
Using the above determinations, we find the concentration of superparamag-
netic clusters ncl and average number of paramagnetic atoms in clusters ns for the
steels CN107, CN85 and MnC (Table 3.5).
The size of superparamagnetic clusters can be estimated using the number of
paramagnetic atoms in the cluster ns. The concentration of clusters in the CN steels
is consistent with that obtained for steel CN85 using CESR, although their size is a
bit higher. According to the obtained data, steel CN107 is characterized by a low
concentration of clusters ncl and a size of about 5 nm, whereas ncl in steel CN85 is
higher, however, the clusters have a smaller size of about 3.5 nm. This result
suggests that chemical homogeneity in the studied CrMnCN steels is improved
with decreasing C/N ratio.
The carbon steel MnC is characterized by the smallest concentration of su-
perparamagnetic clusters which have the highest size of about 8 nm. In [25], using
small angle scattering of neutrons, the size of chemical clusters in the CrMnCN
steels was determined at about 10 nm with an error of about 30 %. The point is
that, in comparison with magnetization measurements, the resolution of this
technique is not sufficient to study clusters with a size smaller than 5 nm. A much
higher neutron scattering in steel MnC was studied with sufficient precision, and
the measured size of chemical clusters was about 30 nm.
This discrepancy can be evidence of a larger size of chemical clusters in
comparison with superparamagnetic ones and needs a special comment. It should
be noted that the formation of clusters from substitutional solute atoms can have a
different effect on the polarization of the magnetic moments leading to the su-
perparamagnetism and on the enhanced neutron scattering which depends only on
the different neutron scattering amplitudes of the nuclei.
44 3 Structure

Fig. 3.18 Frequency distribution of measured dislocation node radii in austenitic steels. a CN96
and b MnC

3.1.3.3 Stacking Fault Energy as a Tool to Estimate the Chemical


Homogeneity

In addition to the experiments described above, we used the measurement of SFE


to obtain information on the chemical inhomogeneity of the studied steels. Such
measurements are also useful for testing a possible correlation between the elec-
tron structure and SFE in view of the data in [27] where an inverse correlation was
observed between the SFE and the density of electron states at the Fermi level.
The SFE c was estimated using the following equation (see [28]):

c ¼ 0:26 lb2p =ri : ð3:9Þ

where l is the shear modulus and bp is the Burgers vector of a partial dislocation, ri
the radius of a circle inscribed in the dislocation triple node.
As an example, the results of measuring the dislocation splitting are shown for
steel CN96 and MnC in Fig. 3.18 and the obtained dislocation node radii ri and
SFE for all studied steels are presented in Table 3.6.
Two maxima of dislocation splitting (see Fig. 3.18) give evidence of short-
range decomposition in the studied steels. The fitting of the experimental data was
carried out using a Gauss distribution, which is natural for the statistical scattering
of the experimental data. One can see that the homogeneity of the solid solutions is
improved with decreasing content of carbon in the CN steels, consequently with
decreasing C/N ratio, and it is significantly worse in the carbon steel MnC where
the gap between the measured SFE values is the highest (Table 3.6).
The obtained values of SFE are presented in Fig. 3.19 along with the data on
the concentration of free electrons which are proportional to the density of electron
states at the Fermi level D(EF).
Qualitatively, the obtained results are consistent with those in [29, 30], where
the short-range decomposition in neutron- and electron-irradiated austenitic CrNi
steels was studied by local X-ray emission spectroscopy. It was shown that, with a
modulation wave of about several microns, the areas enriched in iron and
3.1 As-Quenched 45

Table 3.6 Dislocation node radius ri and SFE


Steel r1, nm r2, nm SFE1, mJ/m2 SFE2, mJ/m2 DSFE, mJ/m2
CN107 7.3 10.2 59 43 16
CN96 11.5 16.0 38 27 11
CN85 10.0 12.0 43 36 7
MnC 6.4 10.6 68 41 27

Fig. 3.19 Stacking fault


energy, SFE, and
concentration of free
electrons ne in CrMnCN
steels versus C ? N content
in atom %

chromium alternate with those enriched in nickel. Since the electron irradiation
does not change the atomic interactions and only, due to enhanced diffusion,
assists the thermodynamic equilibrium, it is clear that the atomic distribution in
austenitic steels is far from that in ideal solid solutions.
Both SFE1 and SFE2 are inversely proportional to the concentration of con-
duction electrons, which is consistent with the tendency observed in [27] for pure
metals. It is relevant to note that the same inverse relation between SFE and D(EF)
was observed in [31] for austenitic CrMn steels with nitrogen, whereas a direct
proportionality occurred in the same steels if additionally alloyed with nickel.

3.1.3.4 Effect of C/N Ratio on the Chemical Homogeneity

The main question is how the C/N ratio in the austenitic CrMnCN steels affects
their chemical homogeneity. The following analysis can be made based on the
measurements of electron spin resonance, magnetization and stacking fault energy.
In comparison with nitrogen steels studied in [7–9], the ones with
carbon ? nitrogen are characterized by an increase in the concentration of free
electrons. The data presented in Table 3.2 show that the concentration of free
electrons generally increases with a decreasing C/N ratio. Also there is no
remarkable difference between steels CN96 and CN85, which points to some
optimum of the C/N ratio. In addition the data of CESR (Sect. 3.1.3.1) and
magnetization (Sect. 3.1.3.2) show that, with decreasing C/N ratio, the density of
clusters increases, whereas their size decreases. In other words, the atomic
46 3 Structure

distribution becomes more homogeneous, which suggests a tendency to short-


range atomic ordering. This result is consistent with earlier data on the correlation
between the metallic character of interatomic bonds and the atomic distribution in
austenitic steels with nitrogen and carbon [32] and allows to assume that an
appropriate C/N ratio can provide the highest thermodynamic stability of austenitic
steels.
A correlation between atomic interaction and atomic distribution is also con-
firmed by the markedly increased size of clusters in Hadfield steel MnC (see the
data of magnetization in Sect. 3.1.3.2) and the absence of the CESR signal in this
steel, which is a sign of an extremely low concentration of free electrons. As a
result, the Hadfield steel reveals a rather low stability in relation to precipitation at
low temperature heating. In view of the decreased concentration of free electrons,
i.e. the prevailing covalent character of interatomic bonds in the Hadfield steel, a
special comment can be given in relation to its good impact toughness, though it is
smaller compared with austenitic steels containing nitrogen or carbon ? nitrogen
(see [15]). The stacking fault energy of about 50 mJ/m2 [33] is not favourable for
strain-induced c ? e transformation. At the same time, the carbon enhanced
covalent bonds locally increase the shear modulus within the carbon clouds around
dislocation, which increases the line tension of dislocations and, correspondingly,
decreases their mobility. These two factors should create appropriate conditions
for the extremely intensive twinning in the Hadfield steel by cold working. In fact,
the Hadfield steel is a good example of the TWIP effect, which is the real reason
for its tough mechanical behaviour. As shown in [33], such an intensive twinning
is not observed in austenitic steels with nitrogen or carbon ? nitrogen.
Finally, let us discuss the applicability of SFE measurements to estimate the
chemical homogeneity of solid solutions. The distance between the SFE maxima
DSFE in Table 3.6, decreases with decreasing C/N ratio, which is consistent with
the data of chemical inhomogeneities obtained by other experimental methods.
Summing up one should state that the studied austenitic CrMnCN steels are
characterized by a high concentration of free electrons, which evidences the
enhanced metallic character of interatomic bonds. A feature of the atomic distri-
bution in the austenitic solid solutions is the existence of clusters rich in Cr and
Mn, which can be detected by their superparamagnetism or the enhanced small
angle scattering of neutrons on the nuclei of atoms in the chemical
inhomogeneities.
The size of clusters decreases with a decreasing C/N ratio, whereas their
number increases, which suggests that the chemical homogeneity of steels
increases. The same conclusion follows from the measurements of the stacking
fault energy. A clear gap in the frequency curves, showing the distribution of
dislocation splitting, corresponds to the chemical inhomogeneity of solid solutions
on a nano-scale. The difference in the measured values of the stacking fault energy
DSFE corresponds to the short-range decomposition which decreases with
decreasing C/N ratio.
3.2 Structural Change by Loading 47

3.2 Structural Change by Loading

3.2.1 Tensile Straining

The aim of this section is to analyse the influence of carbon, nitrogen and
carbon ? nitrogen on the change of structure in austenitic steels caused by cold
work and, based on the obtained data, discuss the controlling mechanism of cold
work hardening.
The enhanced strengthening by plastic deformation is a general feature of iron-
based interstitial solid solutions. In relation to austenitic steels, a number of studies
were devoted to Hadfield steel MnC (e.g. [34–50], and, nevertheless, the nature of
its extremely high cold work hardening remains a controversial issue. In early
publications, it was attributed to strain-induced c ? e [34–37] or c ? a [38, 39]
martensitic transformations. These observations were later explained also as a
result of segregation [39], precipitation [40] or decarburization [41].
Perhaps for the first time, the idea of twinning as a reason for increasing the
cold work hardening of Hadfield steel was expressed in [42]. Another mechanism
based on dynamic strain ageing, namely the break-up of pinned dislocations from
carbon clouds and the restored pinning due to accelerated pipe diffusion of carbon
atoms along the dislocations was proposed in [43], where both the enhanced cold
work hardening and the serrated flow were explained in such a way.
However, it should be noted in this relation that the enhanced migration of
atoms by pipe diffusion occurs only for substitutional solutes because only the
migration of vacancies is accelerated along the dislocation pipes, and that is not a
migration mechanism for interstitials. On the contrary, in comparison with the
bulk, the interstitial atoms lose their mobility in the vicinity of dislocation cores
because the elastic term practically disappears from the gradient of chemical
potential (see, e.g., experimental data [44] about the retarded carbon diffusion in
previously cold worked iron and other metals).
In further studies of Hadfield type steels the formation of twins or slip
depending on the crystallographic orientation was confirmed [45], whereas the
decisive role of dynamic strain ageing was denied [46].
In studies of austenitic high nitrogen CrMnNi steels, the cold work hardening
was attributed from the very beginning to the nitrogen-caused planar slip at lower
strains and deformation twinning at higher strains [47–49]. At the same time, the
increase in dislocation density during cold working was considered as a factor
preventing brittle fracture [50]: the dislocations shield stresses created by inter-
secting twins where the cracks are thought to be nucleated. Alloying with nickel
assists the dislocation mode of plastic flow and shifts twinning to higher strains.
With decreasing deformation temperature or increasing nitrogen and manganese
contents, the onset of twinning is shifted to lower strains, whereas the dislocation
density decreases in between the twins.
A particular role in cold work hardening of nitrogen steels was ascribed to
secondary twinning, i.e. the formation of twins within the space between the
48 3 Structure

primary twins [51]. The multiple twin system was analysed in terms of an increase
in the efficiency of twins as hardening obstacles for gliding dislocations and
formation of new twins.
At the same time, no remarkable twinning was found in other studies of cold
worked austenitic nitrogen steels. For example Kubota et al. [52] examined the
structure of SUS316L steels (i.e. CrNi ones) with 0.02 and 0.56 mass % nitrogen
and Cr18Mn18 type steels with 0.51 and 0.84 mass % nitrogen after tension tests
and observed an increase in work hardening with increasing nitrogen content and
no deformation twinning but planar dislocation arrays forming Lomer-Cottrell
barriers at their intersections.
In a similar way, Saller et al. [53] demonstrated the microstructures of strained
CrMn austenitic steels alloyed with 0.3 or 0.9 mass % nitrogen, where planar slip
was the main structural feature in case of high nitrogen content, whereas twins and
e-martensite appeared occasionally at rather high strains.
This discrepancy between the observations made by different authors could be
caused by a different crystallographic texture. One of the factors controlling
twinning in austenitic steels, as well as strain-induced c ? e formation, is the
stacking fault energy: the lower the SFE the smaller is the stress for twin for-
mation. According to [54, 55], the actual stress on the leading partials can be
higher or lower than that on the trailing ones depending on the crystal orientation
in relation to the tensile axis (see more about details in the next Section). Cor-
respondingly, the dislocation splitting becomes larger or smaller, which should
affect twinning or c ? e transformation. As shown by Lee et al. [56], this results in
the orientation dependence of twinning during tensile deformation of high nitrogen
austenitic steel Cr18Mn18Mo2N0.9. In relation to the tensile axis, two twinning
systems, primary and conjugating, were observed in \111[ grains, whereas only
one twinning system was activated in \110[ grains and no deformation twinning
occurred in \100[ grains.

3.2.1.1 Yielding

The yield strength belongs to the most important characteristics of engineering


materials used by designers of parts or constructions which are exploited under
loading. In absence of precipitates, it is determined by the following structural
factors: (i) the pinning of dislocation sources by clouds of solute atoms, (ii) the
potential relief of crystal lattice created by interactions of host atoms resulting in
the so-called Peierls barriers, the height of which increases with decreasing tem-
perature (see, e.g. [57]), (iii) the interaction between gliding dislocations and
‘‘dislocation forest’’ consisting of the dislocations intersecting the slip plane, (iv)
the effect of solute atoms which create fields of elastic stresses to be overcome by
dislocations, (v) the grain size which limits the length of the slip plane and,
thereby, the number of dislocations in their plane ensemble, called pile-up. The
latter is particularly important because the shear stress sl acting on the leading
3.2 Structural Change by Loading 49

dislocation in the pile-up increases with the number of dislocations n in the pile-up
as sl = ns, where s is the applied stress.
The dislocation segments in hot forged and solution annealed steels are usually
surrounded by clouds of interstitial atoms, which requires an additional stress for
slip activation. If so, the start of yielding can be accompanied by a sudden decrease
of loading, ‘‘yielding tooth’’, which in early studies was attributed to the breaking
of dislocations from interstitial clouds [58]. However, as shown later by experi-
ment [59, 60], as well as in the theoretical estimations [61], the stress needed to
break the pinned dislocations from interstitial clouds is so high that it exceeds even
the ultimate strength and, therefore, the emission of new fresh dislocations by
dislocation sources should occur before unpinning. In turn, the yielding tooth can
result from dynamics of dislocation multiplication in a solid solution where dis-
locations are blocked by solute atoms.
In contrast to bcc materials, the Peierls relief in metals with an fcc crystal
lattice, which is just the case of austenitic steels, is shallow and cannot really affect
dislocation slip. The same applies to a possible asymmetry of dislocation nuclei,
which in bcc crystals results from a small splitting of dislocations in two or three
atomic planes and is proposed instead of the Peierls relief to be responsible for a
striking increase of the yield strength with decreasing temperature [62].
Nitrogen in austenitic steels remarkably enhances the temperature dependence
of the yield strength, which is not typical of materials with an fcc crystal lattice.
This feature stimulated an idea supported by theoretical calculations [63],
according to which nitrogen in austenitic steels causes an asymmetry of dislocation
nuclei and, for this reason, makes the temperature behaviour of the yield strength
similar to that of bcc materials. However, the experimental data [64, 65] have
clearly shown that the temperature behaviour of the yield strength in austenitic
nitrogen steels is controlled by the intersection of dislocations, as predicted by
Seeger’s theory for fcc crystals [66]. Therefore, some other mechanisms should
exist for the favourable nitrogen effect on the yielding of austenitic steels at low
temperatures (see Sect. 3.2.2).
In interstitial solid solutions, moving dislocations overcome elastic stresses
created by interstitial atoms, which increases the term r0 in the Hall–Petch
equation for the yield strength:

ry ¼ r0 þ ky d1=2 ; ð3:10Þ

where d is the grain size and the Hall–Petch coefficient ky is an ‘‘unblocking


constant’’ characterizing the stress needed for the activation of dislocation sources
in a neighbouring grain in order to transfer deformation through the grain
boundary (see, e.g., [67] for details).
Both r0 and ky depend on the affinity of interstitial atoms to dislocations, which
is characterized by the binding enthalpy. In contrast to a-iron where the binding
enthalpy of carbon and nitrogen atoms with dislocations is practically the same, it
was shown [68] that austenitic steels are distinguished by a significantly stronger
affinity of nitrogen atoms to dislocations. This feature has its nature in a higher
50 3 Structure

elastic interaction between nitrogen atoms and dislocations in comparison with


that of carbon atoms in iron austenite, as well as in the additional chemical
interaction due to dislocation splitting.
Judging from the concentration dependence of the a-iron lattice parameters
[69], there is no remarkable difference between carbon and nitrogen atoms in their
elastic contribution to the dislocation-interstitial interaction, which is dominant in
bcc metals. On the contrary, nitrogen atoms in iron austenite cause higher lattice
distortions in comparison with carbon ones. Moreover, their chemical interaction
with dislocations stems from a different solubility of interstitial atoms in the fcc
solid solution and in the stacking faults which in fcc crystals have a hcp structure.
Nitrogen can be dissolved in hcp iron in a wide concentration range (see, e.g., the
binary Fe–N diagram in [70]), whereas the carbon solubility in hcp iron is limited.
The oversaturation of stacking faults with carbon is clearly illustrated by precip-
itation of cementite in plates of the e-martensite transformed in the course of
fatigue tests of steel CN96 (see Sect. 3.2.4.2). Thus, so-called Suzuki’s atmo-
spheres of carbon and nitrogen atoms at split dislocations add to the term r0 which
generally consists of contributions from the lattice friction stress (Peierls relief),
the stress needed to overcome distortions created by solute atoms and the inter-
secting dislocations penetrating the slip plane.
The interaction between gliding dislocations and dislocation forest is enhanced
due to their splitting in austenitic steels because the stress needed for the formation
of a constriction at the point of intersection is the higher the larger the splitting is.
Along with this, the carbon and particularly nitrogen atoms effectively
contribute to the increase of the coefficient ky enhancing thereby the grain
boundary strengthening of austenitic steels, which was extensively studied in [71–
75] and interpreted in [67].
Based on findings mentioned above, one can predict that both carbon and
nitrogen in austenitic steels increase the yield strength and the effect of nitrogen is
expected to be more pronounced. Results of mechanical tests of C ? N austenitic
steels will be presented in Sect. 4.1.

3.2.1.2 Effect of Stacking Fault Energy

It is natural to suppose that a different mechanical behaviour is controlled by a


different substructure formed during straining of austenitic steels with carbon,
nitrogen and carbon ? nitrogen. The stacking fault energy SFE is responsible for
dislocation slip, as well as for twinning or formation of e during plastic defor-
mation. Usually, an SFE smaller than 100 mJ/m2 is considered to be low (see, e.g.,
[76]).
Dislocation slip prevails at high SFE, which makes the dislocations narrow and
eases a change of the slip plane. Some middle SFE value assists twinning along
with dislocation slip. Strain-induced c ? e transformation occurs at sufficiently
low SFE.
3.2 Structural Change by Loading 51

Fig. 3.20 Effective splitting


of leading (1/6½211) and
trailing (1/6½112) partial
dislocations under applied
shear stress in the fcc lattice

It is worth noting that the SFE values should be measured under condition of
extremely low density of dislocations, which is achieved due to solution treatment.
This ‘‘true’’ stacking fault energy depends only on the chemical composition of
solid solutions and determines their thermodynamic stability.
In cold worked steels, a so-called ‘‘effective’’ stacking fault energy (see [54,
55]) is controlled by an increase or decrease of dislocation splitting depending on
the orientation of the Burgers vectors of the leading and trailing partial disloca-
tions in relation to the acting applied stress or to the retained stresses after
deformation (Fig. 3.20).
This is why the change of crystallographic texture by deformation and the
different crystallographic orientation of grains require the measurement of a
number of triple node radii and statistical analysis of the experimental data as done
in Sect. 3.1.3.3 (see Fig. 3.18).
In the studied steels, the SFE was measured to be equal to 52 in steel MnC, 36
in steel CrMnN and 40 in steel CN96 with an accuracy of about ± 3 mJ/m2. The
value obtained for steel MnC is much higher than follows from the thermody-
namically estimated 22 mJ/m2 in [77] and correlates better with the well known
effect of increasing carbon on the SFE of austenite (see e.g. [78, 79]). Within the
error of measurements, the SFE values of steels CrMnN and CN96 are the same,
which suggests a nearly similar deformation behaviour.

3.2.1.3 Substructure of Cold Worked Steels

Structural and phase transformations caused by cold working the carbon steel
MnC, nitrogen steel CrMnN and carbon ? nitrogen steel CN96 of Table 2.1 are
going to be analyzed in the attempt to find features responsible for their
mechanical behaviour.
For this purpose, tensile specimens were strained to a pre-selected elongation
between 10 and 50 %. The respective true stress r/true strain u curves were
derived from engineering stress/strain curves up to necking (see Fig. 4.1).
The exponent and coefficient of Ludwik’s Eq. (3.11) were taken from the final
slope of these curves (Table 3.7).
52 3 Structure

Table 3.7 Exponents and coefficients of the Ludwik Eq. (3.11) after pre-selected elongation
Elongation e [%] 10 20 30 50
(true strain u) (0.09) (0.18) (0.26) (0.41)
MnC n 0.29 0.49 0.65 0.77
K 1195 1770 2250 2600
CrMnN n 0.28 0.41 0.43 0.41a
K 1700 2185 2270 2200a
CN96 n 0.28 0.41 0.46 0.53
K 1585 2060 2315 2430
a
taken at e = 40 %, i.e. below Au

r ¼ K/n : ð3:11Þ
As follows from the obtained data, the intensive cold work hardening of
Hadfield steel MnC is supported by Ludwik’s exponent n which tops the other
steels at the high end of elongation. While n of MnC and CN96 increases from 10
to 50 % elongation, the cold work hardening of CrMnN was slightly reduced at
40 % leading to the lowest uniform elongation Au of the three steels investigated
(Table A1). At the elongation of 50 %, K changes proportional to the content of
interstitial atoms, be it carbon or nitrogen.
The change in the structure of three selected steels is analysed. It is expedient to
start from a general picture which can be obtained using the optical microscopy
even if it could not be decisive in clarifying the difference between mechanisms of
cold work hardening in the studied steels.
As an example, the sequence of structural change with increasing tensile strain
is presented in Fig. 3.21 for steel CN96 (see also [80]).
The strain of 10 % causes separate slip lines in some grains (Fig. 3.21a). Even
at 30 % of strain, not all the grains are involved in slip (Fig. 3.21b), possibly
because of the difference in their orientation to the applied stress. At the same
time, a sign of the slip on secondary planes can be already seen. At 50 % of strain,
which is the limit of the uniform elongation, the density of slip lines increases
significantly and the slip becomes more homogeneous (Fig. 3.21c). Nevertheless,
the deformation degree is different for different grains.
An important feature of the microstructure after 50 % strain is also the cur-
vature of slip lines, which is clearly due to the interdependence in the deformation
of the neighbouring grains having a different crystallographic orientation. Such a
character of the slip lines also indicates that twinning and c ? e transformation are
not preferential as the deformation mode of this steel.
Transmission electron microscopy allows one to distinguish further features in
structure of the three studied steels which can be responsible for their different cold
work hardening and plasticity limit (see about details [33]).
An intensive twinning is observed in steel MnC at a rather small elongation of
10 % (Fig. 3.22a). Remarkable is the large thickness of the twin plates. Twinning
continues at higher strains and, additionally, the dislocation density significantly
3.2 Structural Change by Loading 53

Fig. 3.21 Microstructure of steel CN96 after tension tests with deformation degree. a 10 %.
b 30 %. c 50 %

Fig. 3.22 Structure of steel MnC after a 10 % elongation with zone axis [112], twin zone axis
½1
12and twinning plane ð
111Þ. b 30 % elongation with zone axis [110], twin zone axis ½
1
10 and
twinning plane ð111Þ
54 3 Structure

Fig. 3.23 a Planar slip and b tangled dislocations in steel CrMnN after 10 % elongation

increases (Fig. 3.22b). The dislocations are tangled and reveal a weak tendency to
form a cellular structure. No traces of a or e martensite are observed.
Planar dislocation slip (Fig. 3.23a) and a tangled dislocation structure
(Fig. 3.23b) are developed in the range of small elongations of the nitrogen steel
CrMnN. The nitrogen-induced short-range atomic ordering is the obvious reason
for the formation of planar dislocation arrays.
An increase in elongation up to 30 % results in planar gliding on intersected
slip planes (Fig. 3.24a). At the same time an essential twinning appears in this
strain range (Fig. 3.24b).
Moreover, two systems of twinning are found, one of which, namely
{113} \120[ , is uncommon. Possibly, new twinning systems are due to condi-
tions of retarded strain in the highly strengthened fcc lattice. The occurrence of
intensive twinning is at variance with the data of [53] where the authors observed
very rare twinning in the same steel after tensile deformation.
Intensive twinning, including that on intersecting planes, occurs also if steel
CrMnN is elongated by 50 % (Fig. 3.25). Again, an uncommon twinning plane is
observed.
A small tensile deformation of steel CN96 causes planar slip like that in the
pure nitrogen steel CrMnN (Fig. 3.26).
Twinning is involved in plastic deformation at higher elongation (Fig. 3.27a).
At an elongation of 50 %, twinning continues and one can observe intersecting
twinning systems (Fig. 3.27b).
Planar slip continues to be an essential feature of the substructure up to the
elongation of 50 % (Fig. 3.28), which is a bit different from the observations in
steel CrMnN where it was not met.
In both deformed steels, CrMnN and CN96, one can rarely find strain-induced e
martensite. As an example, e plates in steel CN96 subjected to an elongation of
50 % are shown in Fig. 3.29. As follows from the diffraction analysis, the
Nishiyama-Wassermann orientation relation occurs between the fcc and the hcp
crystal lattices.
3.2 Structural Change by Loading 55

Fig. 3.24 a Planar slip on intersecting planes and b twinning on intersecting twinning systems:
with zone axis [231], twin zone axis ½ 2
31, twinning plane ð1
11Þ and twin zone axis ½ 1
20,

twinning plane ð113Þ in steel CrMnN after 30 % elongation

Fig. 3.25 Twins in steel


CrMnN after 50 %
elongation with zone axis
½0
13 and two systems of
twinning: with twin zone axis
½013, twinning plane (131),
and twin zone axis ½35 4,
twinning plane (111)

Fig. 3.26 Planar slip in steel


CN96 after 10 % elongation
56 3 Structure

Fig. 3.27 a Twins in steel CN96 after 30 and b twinning at intersecting planes after 50 %
elongation

Fig. 3.28 Planar slip in steel


CN96 after 50 % elongation

3.2.1.4 Mechanisms of Cold Work Hardening

It is natural to discuss possible mechanisms of cold work hardening in austenitic


steels starting from Seeger‘s theory for the yield stress and cold work hardening in
fcc crystals of low stacking fault energy [66, 81]. According to Seeger, in absence
of twinning or strain-induced fcc ? hcp transformation, the yield stress and cold
work hardening of fcc crystals of low stacking fault energy is controlled by gliding
dislocations cutting through the dislocation forest. In this case, the splitting of
dislocations, i.e. the stacking fault energy, plays an important role because the jog
formation at dislocation intersections needs a preliminary constriction of split
dislocations.
If alloying with interstitial elements does not lead to a remarkable dislocation
splitting, they contribute mainly to solid solution hardening, i.e. to the yield stress,
whereas the flow stress reveals an insignificant dependence on the content of
interstitials.
3.2 Structural Change by Loading 57

Fig. 3.29 a Plates of e martensite in steel CN96 after 50 % elongation. b Diffraction pattern
from the area 1 in a with fcc zone axis [111]. c Diffraction pattern from area 2 in a with fcc zone
axis ½111 and hcp zone axis ½ 361 which are in Nishiyama-Wassermann orientation relation.
d Dark field in the light of e reflection marked with a circle in c

Such a conclusion is confirmed by mechanical tests of the high nickel austenitic


nitrogen steel Cr18Ni16Mn10 (Fig. 3.30). Nickel significantly increases the SFE
in austenitic steels [79], making dislocations narrow, whereas nitrogen in this steel,
increased from 0.06 to 0.5 mass %, changes the SFE non-monotonously between
43 and 65 mJ/m2 (see [31]), which does not cause a remarkable dislocation
58 3 Structure

Fig. 3.30 Effect of nitrogen


on microhardness of steel
Cr18Ni16Mn10 after solution
treatment (e = 0 %) and its
increment due to cold rolling
[70]

splitting in comparison with the nitrogen-free steel. This is why nitrogen con-
tributes mainly to solid solution hardening of this steel (see the curve at e = 0 %),
whereas cold work hardening is not changed with an increasing degree of defor-
mation. A similar result was obtained in [82], where the authors did not find any
effect of nitrogen on the hardness of the cold rolled steel Cr18Ni16Mo2.
An increase in the chromium content can compensate this Ni effect and recover
the nitrogen-induced cold work hardening in high nickel austenitic steels (see e.g.
[83]), because chromium decreases the SFE in iron austenite [79].
The effects of carbon, nitrogen and carbon ? nitrogen on the substructure
under conditions where twinning and/or c ? e transformation accompany the
plastic flow are to be compared. The extremely high cold work hardening in
Hadfield steel MnC is obviously connected with intensive twinning which starts at
small strains (see Fig. 3.22). Twinning occurs in spite of the rather high stacking
fault energy of 52 mJ/m2, i.e. in spite of the reduced dislocation splitting. This is
thought to be related to a strong elastic interaction between carbon atoms and
dislocations. Because of the high concentration of interstitial atoms, their elastic
fields hinder dislocation slip and, as an alternative mode of plastic deformation,
promote twinning even in case of comparatively narrow dislocations.
It is also worth noting that the twins in steel MnC are unusually thick, which
additionally enhances cold work hardening. As mentioned in Sect. 3.1.3.4, Had-
field steel is a good example of the TWIP effect.
At higher strains, along with twinning, a significant increase in the density of
tangled dislocations without any inclination to localized slip should effectively
contribute to cold work hardening. It is worth noting that the chaotic distribution of
dislocations contributes to cold work hardening much more than a cellular dis-
location structure.
The cold work hardening is not as high as in steels CrMnN and CN96 at small
elongations up to 10 %. The only feature of the substructure is planar slip on one
of acting slip systems (see Figs. 3.23 and 3.26). This planar slip is usually
attributed to short-range atomic ordering based on the studies [54, 55]. However,
the role of the nitrogen-induced planar slip in cold work hardening is sometimes
3.2 Structural Change by Loading 59

exaggerated (e.g. [53]), at least at moderate strains where planar slip bands do not
intersect each other.
To discuss this topic again, the example of high nickel austenitic nitrogen steels
is used. As mentioned above, nitrogen does not contribute to cold work hardening
if the nickel content is high. At the same time, pronounced planar slip occurs
during deformation (see, e.g., data on planar slip in steel Cr26Ni32N0.36 [84]).
Thus, it is hardly possible to attribute cold work hardening in steels CrMnN and
CN96 to planar slip, at least at moderate strains. Only the intersection of planar
slip systems at higher strains should contribute to cold work hardening due to
formation of Lomer-Cottrell barriers which effectively block the slip in active
{111} planes. In this relation one should remark that, at large strains, slip planarity
on intersecting slip systems is met statistically more often in steel CN96 than in
CrMnN, which can provide a higher cold work hardening in steel CN96. This can
be also attributed to a stronger short-range atomic ordering in carbon ? nitrogen
austenitic steels as compared with the nitrogen steel (see [22]).
A remarkable twinning, although not as intensive as in steel MnC, along with a
moderate c ? e transformation is observed at elongations of 50 %, and one cannot
really find a clear difference in the structure evolution in both steels alloyed with
nitrogen or nitrogen ? carbon except for some prolonged planar slip in steel
CN96. A similar change of structure during plastic deformation is possibly con-
trolled by nearly the same stacking fault energy.
It seems that additional arguments could be obtained by measuring the binding
enthalpy between dislocations and interstitials in both steels. As mentioned earlier,
according to [68], the pinning of dislocations by nitrogen atoms in austenitic steels
is stronger than that by carbon atoms. A higher affinity of nitrogen atoms to
dislocations is one of the reasons for some additional stress needed for cutting the
gliding dislocations through a dislocation forest pinned by interstitial atoms. It is
so far unknown how strong the pinning of dislocations by carbon ? nitrogen
atoms is.
One can also predict that the enhanced metallic character of interatomic bonds,
which is brought about by the higher concentration of free electrons in nitrogen
and particularly in carbon ? nitrogen steels (see Sect. 3.1), has to retard the
opening of cracks during plastic deformation and assist a higher fracture elonga-
tion of N and C ? N steels.

3.2.2 Effect of Subzero Temperature

In contrast to austenitic carbon steels, a feature of the nitrogen ones is an enhanced


temperature dependence of the yield and flow stress. It was shown in [65, 85–88]
that, while at room temperature the difference between the tensile strength of
austenitic carbon and nitrogen steels does not really exceed the error of mea-
surements, it increases with decreasing temperature. According to [88], nitrogen at
60 3 Structure

Fig. 3.31 Temperature


dependence of the yield
strength in steel
Cr18Ni16Mn10 not
purposefully alloyed with
carbon or nitrogen and
alloyed with carbon or
nitrogen [65]. The data for
steels MnCr82, CN85, CN96,
CN107 and CrMnN stem
from [80]

4 K is more effective strengthener of austenitic steels, by a factor of 1.4, than


carbon.
Based on these studies, the idea was proposed [89] that nitrogen assists a bcc-
like behaviour of the flow stress of austenitic steels with temperature, which is
strongly different from the behaviour of fcc substitutional solid solutions, and
theoretical calculations [63, 90] were performed, according to which the alloying
of austenitic steels by nitrogen causes a splitting of the core of screw dislocations
on two or more non-parallel planes, exactly as it takes place in the crystals with a
bcc lattice.
In contrast to these calculations, Obst and Nyilas [64, 76, 91] obtained by
precise measurements that a three-stage temperature dependence of the flow stress
occurs in austenitic nitrogen steels in accordance with Seeger’s theory [66, 81] for
fcc crystals of low stacking fault energy (i.e. lower than 100 mJ/m2). These three
stages include the intersection of dislocation forest by edge dislocations (stage 1),
screw dislocations (stage 2) and emission of vacancies which is needed for the slip
of a jog formed in case of two intersecting screw dislocations (stage 3).
According to [65], nitrogen in austenitic steels affects the third stage (see
Fig. 3.31). As the constriction of split dislocations is a prerequisite for jog for-
mation, the effect of nitrogen on the low temperature strengthening has its origin in
the nitrogen-enhanced temperature dependence of the stacking fault energy: the
more split the dislocations are, the higher is the stress needed for their constriction.
As shown in [65], nitrogen enhances a decrease of the stacking fault energy with
decreasing temperature in consistency with the temperature dependence of the
density of electron states D(EF), i.e. the concentration of free electrons, and the
inverse correlation between SFE and D(EF).
Results of steels listed in Table 2.1 are also presented in Fig. 3.31. Quantitative
data are given in Table 3.7 and A6. From these it follows that, at higher contents of
interstitials, nitrogen and carbon ? nitrogen increase the yield strength and its
temperature dependence within the range of the 2nd stage, where the intersection
of gliding screw dislocations with the dislocation forest controls their slip, as well
3.2 Structural Change by Loading 61

as at the 3rd stage, where the emission of vacancies is needed for slip of a jog
formed due to the intersection of screw dislocations. This is obviously due to the
C- and (C ? N)-caused temperature dependence of the stacking fault energy
which is enhanced by nitrogen and should be even more enhanced by car-
bon ? nitrogen (see effect of C, N and C ? N on the concentration of free elec-
trons in Fig. 3.6).
One can see that the uniform as well as the total elongation remains high at
173 K, which evidences a good low temperature plasticity of the CrMnCN steels
(Table A6). The coefficients K and n in the Ludwik Eq. (3.11) are increased with
decreasing temperature. As follows from the data of [64] for the austenitic nitrogen
steel Cr25Ni15N0.35, the coefficient n can be even decreased within the tem-
perature range of stage 3 in Fig. 3.31. Considering an increasing dislocation
density, the constriction of the split dislocations should require a lower applied
stress because of stresses caused by neighbouring dislocations.
Thus, a high ductility accompanies the strengthening of austenitic (C ? N)
steels at temperatures as low as -100 °C (Sect. 4.1.2). As will be shown in Sect.
4.1.6, this is also true for the impact toughness except for the phenomenon of
ductile-to-brittle transition, which is very unusual for materials with an fcc crystal
lattice and will be discussed in the following Section.

3.2.3 Effect of Strain Rate

3.2.3.1 Rapid Tensile Tests

Tensile tests of steel MnCr82 alloyed with C ? N (Table 2.1) were carried out at
an initial strain rate between 3.3 9 10-4 and 3.3 9 10-2 [92]. As the rate was
raised, the yield strength increased, the ultimate tensile strength and elongation
decreased, whereas the reduction of area remained nearly unchanged.
The strain rate dependence of plastic flow and fracture of austenitic nitrogen
steels was studied by Tomota et al. in more detail [93]. They tested CrMn and
CrNi steels varying the strain rate by 3 orders of magnitude from 2.7 9 10-3 s-1
to 2.7 9 10-1 s-1. The authors changed the nitrogen content (mass %) from 0.513
to 0.844 in a CrMn steel and substituted Mn by Ni in a steel with 0.556N.
Raising the strain rate at RT increases the yield strength and reduces the
elongation and reduction of area. The ultimate strength behaves non-monotonous,
depending on the nitrogen content.
At 77 K, no plastic deformation and cleavage fracture occurred in steel
Cr19Mn19N0.844 having the highest nitrogen content, whereas some elongation
was obtained in steel Cr17Mn19N0.513, which is generally expected. However,
surprising is the absence of necking and a mixture of dimple and cleavage-like
fracture in steel Cr17Mn19N0.513 at the small strain rate of 2.7 9 10-3 s-1,
whereas necking and ductile dimple fracture occurred at the higher strain rate of
2.7 9 10-1 s-1. It is remarkable that the same effect (necking and dimple fracture)
62 3 Structure

is obtained at the strain rate of 2.7 9 10-3 s-1 in steel Cr17Mn1Ni13N0.556,


where, at the same chromium content, manganese is substituted by nickel.
Tomota et al. [93] interpreted their results of steel Cr17Mn19N0.513 in terms of
‘‘work softening’’ caused by an abrupt increase of local temperature in slip bands
due to rapid loading.
Another possible explanation can be based on the effect of nitrogen on the
atomic interactions, as discussed in Sect. 3.1. Increasing the concentration of free
electrons by nitrogen enhances the metallic character of interatomic bonds, which
should assist ductile fracture. However, as follows from Fig. 3.6, the effect of
nitrogen on the concentration of free electrons in austenitic nitrogen steels is non-
monotonous. Reaching a maximum at about 0.5 mass % of nitrogen, the con-
centration of free electrons decreases with a further increase in the nitrogen content
and is rather small in steel Cr18Mn20N0.88 the composition of which is close to
that of steel Cr19Mn19N0.844. The favourable effect of nickel can be attributed to
the nickel-caused increase of the concentration of free electrons in iron austenite
(see [1]), which increases plasticity at RT, as well as at 77 K, and provides a
ductile fracture.
Taking into account this correlation between the concentration of free electrons
and mechanical properties, one can foresee a favourable effect of alloying CrMn
steels with C ? N (see Fig. 3.6), which will be shown in Sect. 4.1.
A further increase in the strain rate up to that of impact loading leads to the
phenomenon of a ductile-to-brittle transition, DBT, which is typical for metallic
materials with a bcc crystal lattice but is unique for those with an fcc one.
Tobler and Meyn [94] were probably the first to observe the quasi-cleavage of
austenitic nitrogen steels during impact tests. It was a special feature that the
fracture occurred along the crystallographic {111} planes, i.e. along the close–
packed slip planes, whereas the cleavage in the bcc crystals proceeds in {001}
planes which are not close-packed. This cleavage-like fracture was observed in
notch impact as well as in tensile tests performed at subzero temperatures. In
contrast, the toughness of austenitic carbon steels gradually decreases with
decreasing temperature and increasing strain rate.
The following features characterise this kind of brittle fracture: (i) alloying
elements have a strong effect (nickel decreases, whereas nitrogen, manganese and
chromium assist cleavage, see [95–97]; (ii) the grain size has a negligible effect
[98], which is quite different from the behaviour of bcc materials, where cleavage
is markedly enhanced by increasing the grain size.
A number of hypotheses were proposed for interpretation of this unusual
phenomenon. Tobler and Meyn [94] suggested a critical role of planar slip
resulting in an accumulation of shear strain in the narrow bands, their weakening
by accumulated dislocations and final separation. According to Müllner et al. [99],
the intersection of deformation twins is a reason for the nucleation of cracks.
Nitrogen, as well as manganese, shift the onset of twinning to lower strains where
the density of dislocations is low and a critical stress for brittle fracture can be
achieved before the crack tip blunts due to dislocation slip. This mechanism is at
3.2 Structural Change by Loading 63

clear variance with the data about the absence of twinning and the high density of
dislocations just under the fracture surface, as obtained by Tomota et al. [100].
Moreover, these authors have clearly demonstrated the localization of plastic
flow, the nitrogen-enhanced slip on the {111} planes (‘‘slipping-off’’, according to
[100]), the nucleation of submicron cracks in these planes and their fusion on non-
active {111} slip planes, which creates cleavage-like facets on the fracture surface.
Based on the data about the effect of nitrogen on the electron structure and
short-range atomic order, we can add the following details to the interpretation of
DBT in austenitic nitrogen steels proposed by Tomota et al.
First, the localization of dislocation slip is obviously related with nitrogen-
caused short-range atomic ordering. It is worth noting that this ordering is not
simply related with the preferential Cr–N atomic bonds, as is often claimed. Due to
the increase in the concentration of free electrons, nitrogen changes the distribution
of substitutional solutes assisting preferential bonds between atoms of different
kind. In contrast, carbon enhances covalent interatomic bonds, which assists
clustering of one kind of atoms.
Second, the nitrogen-caused ‘‘slipping-off’’, according to the terminology
proposed in [100], is caused by the increase in the concentration of free electrons
resulting in a decreased shear modulus l and, consequently, line tension of dis-
locations C & lb2, thereby increasing their mobility. Additionally, the nitrogen-
enhanced Snoek-Köster relaxation and strain dependence of damping can be
mentioned as an experimental evidence of nitrogen-enhanced mobility of dislo-
cations (see e.g. [101]).
Third, the impact loading is accompanied by an ‘‘abnormal mass transfer’’ (e.g.
[102]), of which the transport of solute atoms by dislocations is a possible
mechanism [103, 104]. If so, the nitrogen atoms should accompany dislocations in
the course of impact loading and locally enhance the metallic character of inter-
atomic bonds within the nitrogen clouds around the dislocations. As a result, the
shear modulus l is locally decreased in the vicinity of dislocations. Their line
tension decreases and their mobility increases. Moreover, in case of dislocation
slip accompanied by a transfer of nitrogen atoms, the distance d between the
dislocations in their plain arrays (pileups) decreases as d & (plb)/16(1-m)ns,
where l is the shear modulus, b is the Burgers vector, m is the Poisson coefficient,
n is a number of dislocations in the pileup, s is an applied stress. Consequently, at
the same applied stress s, the number of dislocations n in the pileup increases and,
correspondingly, the stress sl on the leading dislocation in the pileup, sl = ns,
increases, which assists an earlier opening of microcracks because of slip being
blocked by Lomer-Cottrell barriers formed by dislocations at their intersection (see
Fig. 3.32).
In this relation, it is relevant to note that interstitial atoms affect the mobility of
dislocations in two ways (see [101]). If they are essentially immobile, they always
pin the dislocations. If they are sufficiently mobile to follow the dislocations in the
course of deformation, the mobility of dislocations depends on the effect of in-
terstitials on the electron structure, i.e. the mobility increases with an increasing
concentration of free electrons and decreases in the opposite case.
64 3 Structure

Fig. 3.32 Pileups of gliding dislocations in two intersecting (111) planes in the fcc crystal
lattice. Each dislocation is split, i.e. consists of a stacking fault encased between two partial
dislocations (not shown in the Figure). An immobile sessile dislocation is formed due to a
reaction between two gliding dislocations at the point of their intersection. This sessile
dislocation and two neighbouring split gliding dislocations form a Lomer-Cottrell barrier. It is
also immobile and blocks dislocation slip on both intersecting planes

This is why alloying of austenitic steels with nitrogen increases their impact
toughness at RT and down to a temperature which still provides conditions for
relaxation of stresses caused by the Lomer-Cottrell barriers at the intersection of
localized slip planes. Therefore, a pseudo-brittle fracture of austenitic nitrogen
steels under impact loading occurs at temperatures at which the stresses caused by
the nitrogen-intensified and nitrogen-localized slip cannot be relaxed. This is not
the case for austenitic carbon steels, where dislocations are not split, the slip is not
localized and the mobility of dislocations is not increased. For this reason, the
impact toughness of austenitic carbon steels is much smaller at RT and decreases
gradually with decreasing temperature.
As alloying with carbon ? nitrogen enhances the concentration of free elec-
trons and, consequently, the atomic ordering, it would be reasonable to expect a
further enhancement of the pseudo-brittleness of austenitic C ? N steels. But at a
given interstitial content the DBTT of C ? N steels is lower than that of N steels
(see [15, 105], and Table A11). The same effect is achieved by alloying austenitic
nitrogen steels with nickel (see e.g. [93]).
We can suppose that the reason for this is the significant increase in the con-
centration of free electrons, either by carbon ? nitrogen [10, 25] or by nickel [1].
This increase raises the mobility of dislocations so much, that conditions for the
relaxation of stresses due to dislocation slip are created before submicron cracks
can open along the intersecting slip planes.
3.2 Structural Change by Loading 65

Fig. 3.33 Localised slip in a radial-axial metallographic section of CN96 at the entrance side
near the hole, LOM. a Steps of deformation and ASB. b Band of localized slip. c Cracks in such a
band

3.2.3.2 Ballistic Impact

The joint addition of C ? N to stainless austenitic CrMn steels improves the


specific tensile fracture energy Ws (Table A1) composed of strength and ductility
to the highest level measured so far (see also [14]). The aim of this Section is to
find out if the high energy consumption measured in standard tensile tests can be
retained in case of high velocity impact loading and to analyse structural changes
by the ballistic impact (see about details [106]).
Ballistic tests were carried out with steel CN96 and soft core standard
ammunition 7.62 9 51 WK (NATO Level 1, STANAG 4569) at standard velocity
VZ. = 835 m/s. The firing system was of type Steyr (20 m) in which the plane of a
10 mm thick disc was positioned perpendicular to the trajectory.
All specimens were perforated and thus did not withstand the bullet. This
implies that Vz was above the ballistic limit which marks the transition from
penetration to perforation. The ballistic limit of the nitrogen steel Cr19Mn11N0.37
was measured at 535 m/s (see Sect. 6.2.3). Compared to this HNS the new HIS
failed to meet the expectations. Light optical (LOM) and TEM were carried out to
analyse the reason for this weak performance.
66 3 Structure

Fig. 3.34 Plates of


e-martensite in the hardly
deformed matrix, radial-axial
section taken at a distance of
about 10 mm from the hole

Steps of deformation were found on the entrance side of the bullet near the hole.
The microstructure in between was marked by slip lines pointing to even defor-
mation (Fig. 3.33a). The steps extended into adiabatic shear bands (ASB) in which
a large shear displacement occurred while the material in between was much less
deformed. In some ASB the shear was so high that cracks occurred (Fig. 3.33c).
The phenomenon of ASB by joint localisation of slip and heat has been described
for ballistic loading before (see e.g. [107]). The heat generated by slip has no time
to dissipate (adiabatic), thus heating the slipped region which locally lowers the
yield point and gives rise to more slip just there. At higher magnification an ASB
of about 20 lm in width branches into thinner ones which are only a few lm wide
(Fig. 3.33b).
The inhomogeneous deformation is also visible on the exit side near the bullet
hole where a net of deformation steps encases areas of less but even deformation.
At half-thickness of the disk the hardness reached about 500 HV0.5 at the wall of
the hole. The hardness penetration in radial direction was almost 10 mm.
At a radial distance of 10 mm from the hole no signs of deformation were
visible by LOM. However, occasional plates of e-martensite were revealed by
TEM (Fig. 3.34).
A high density of twins was observed at a distance of 6 mm from the hole
(Fig. 3.35). About 3 mm away from the hole, e-martensite occurred together with
an increased density of dislocations, Fig. 3.36. The dark field image of the e-
martensite is obtained in the light of reflection ( 210)e which is overlapped with
(
323)f reflection of f-Fe2N nitride (see the key diagram in Fig. 3.36d). The pre-
cipitates of f-Fe2N nitride are also lighted in Fig. 3.36c along with dislocations.
At a distance of only 0.1 mm from the hole, localised slip without precipitates
appeared (Fig. 3.37). This suggests that 3 mm away from the hole the temperature
was locally high enough to provoke nitride precipitation while at a distance of
0.1 mm it was locally raised to the solution range.
3.2 Structural Change by Loading 67

Fig. 3.35 TEM structure in a radial-axial section taken at a distance of about 6 mm from the
hole. a High density of twins in the austenite. b Diffraction pattern, zone axis [110]c

Fig. 3.36 TEM structure in a radial-axial section at a distance of about 3 mm from the hole.
a Two systems of e-martensite and small f-Fe2N precipitates. b Diffraction pattern showing
reflections of e-martensite, zone axis [121]e, and f-nitride, zone axis [133]f. c Dark field image of
e-plates in the light of reflection (
210)e and Fe2N nitride in reflection (
323)f. d Key diagram of
electron diffraction
68 3 Structure

Fig. 3.37 TEM structure in a


tangential–axial section at a
distance of about 0.1 mm
from the hole, zone axis
(111)c. Localized slip on
intersecting planes in
directions \110[ , no
precipitates

Planar glide, twinning, e-martensite and an increase of the dislocation density


were also observed after slow tensile straining of HIS steels (see Sect. 3.2.1.3). The
really new deformation feature is the formation of adiabatic shear bands [107]
combined with the precipitation of nitrides. Because of this localised slip, steel
CN96 fails before the high fracture energy Ws, available at low strain rates, is
spent. In Sect. 6.2.3 we discuss if an application seems feasible in spite of the
perforation encountered.

3.2.4 Effect of Cyclic Loading

The previous sections dealt with single loading and the effect of testing temper-
ature, strain rate and stress state (notch) on structural changes. In the following the
effect of cyclic loading is studied at room temperature. At first the structure is
described after repeated impacts which are characterised by predominantly com-
pressive stresses and a high velocity. Next, slow push/pull tests are carried out to
investigate their effect on structure.

3.2.4.1 Repeated Impact

The aim of this Section is to describe structural changes after repeated impact
loading with the attempt to clarify the controlling mechanism of the wear resis-
tance. The tests of impact wear are described in Sect. 4.2.2. The experimental
3.2 Structural Change by Loading 69

Fig. 3.38 Fragment of X-ray


diffraction of steel CN96 after
impact treatment and after
electropolishing of the
impact-treated surface. CuKa
radiation

technique included the use of mineral particles of greywacke with a hardness of


760 HV0.1 and a grid size of 11–8 mm, which hit a plane steel surface perpen-
dicularly at a velocity of 25 m/s.
The mechanical response of steels to impact loading is usually described in
terms of cold work hardening, however different reasons for strengthening are
discussed. A short review of proposed mechanisms for cold work hardening was
given in Sect. 3.2.1. Recent observations point also to an important role of impact-
induced partial amorphisation and formation of a nanocrystalline structure in a thin
surface layer [108, 109]. The surface of high interstitial steels was studied after
impact tests by X-ray diffraction, back scattering Mössbauer spectroscopy and
TEM (see about details [110]).
Figure 3.38 shows the X-ray diffraction obtained from the very surface of the
impact-treated steel CN96 and after subsequently removing a surface layer of 10
lm by electropolishing. It is seen that reflections of silicon oxide SiO2 stemming
from greywacke debris disappear by electropolishing. The austenite remains the
only phase and there is no sign of strain-induced c ? a transformation which is
often claimed in studies of Hadfield type steel exposed to impact surface hard-
ening. In case of a-phase, the {110} martensitic doublet would be located near the
[111] austenitic reflection.
Nevertheless, the Mössbauer spectrum obtained from the surface layer of about
150 nm using the conversion electron mode consists of a single line of para-
magnetic austenite and Zeemann sextet of a ferromagnetic phase (Fig. 3.39).
The surface ferromagnetism after impact treatment is not new and usually
interpreted as a sign of a strain-induced martensitic c ? a transformation. Traces
of Zeemann’s sextets are observed even at the depth of about 10 lm (Fig. 3.40),
whereas, after further electropolishing, the spectrum consists only of the para-
magnetic austenitic component.
It is remarkable that the relative emission of electrons from the impact-treated
surface is rather small (Fig. 3.39) because it is subdued by the implanted SiO2
70 3 Structure

Fig. 3.39 Mössbauer


spectrum obtained from the
impact-treated surface of
steel CN96 in the mode of
conversion electrons. Fec is
an austenitic component. Fe0
and Fe1 belong to Fe atoms in
the martensite having no and
one foreign atom as nearest
neighbours

Fig. 3.40 As Fig. 3.39 but


10 lm below the impact-
treated surface

Fig. 3.41 Mössbauer


spectrum obtained from the
impact-treated surface of
Hadfield steel MnC in the
mode of conversion electrons.
Fecl component belongs to
iron atoms located in clusters
enriched by Mn and C

debris but increases after polishing (Fig. 3.40). Qualitatively the same spectra were
obtained from the impact-treated surface of steel MnC (Fig. 3.41) and steel
CN107. The comparison of Mössbauer results with the data of X-ray diffraction
suggests that the surface ferromagnetism of the studied impact treated steels is not
concerned with a-martensite.
3.2 Structural Change by Loading 71

Fig. 3.42 Structure of the surface of steel CN96 after impact treatment. a Amorphous state.
b Twin structure of austenite

Fig. 3.43 Structure of the impact-treated steels. a CN96 at a distance of 5 lm. b CN107, 15 lm
under the surface

TEM observations of the surface layer after impact treatment of CN steels and
Hadfield steel revealed a complicated picture which is characterized by three types
of substructure: an amorphous state, nanocrystals in an amorphous matrix and a
strongly deformed twinned structure of austenite. Figures 3.42–3.44 demonstrate
examples of the observed structures in steels CN96 and CN107.
The very surface layer consists of an amorphous structure (Fig. 3.42a) with rare
islands of heavily twinned austenite (Fig. 3.42b). The amorphous structure is
observed in steel CN96 at a distance of 5 lm below the surface along with
nanocrystals witnessed by separate reflections at the Debye ring (Fig. 3.43a). In
contrast, a preferentially amorphous state occurs in steel CN107 at a distance of
about 15 lm (Fig. 3.43b). It can be concluded from this comparison that the higher
72 3 Structure

Fig. 3.44 Structure of impact-treated steels at a distance of 45 lm under the surface. a CN96
and b CN107

carbon content assists the formation of an amorphous structure at a larger distance


from the surface. The same tendency is observed after removing a layer of 45 lm:
the structure is characterized by dislocation slip bands and rather high density of
tangled dislocations in steel CN96 (Fig. 3.44a), whereas amorphous areas still can
be found in CN107 (Fig. 3.44b).
In consistency with X-ray diffraction data, TEM studies also confirm the
absence of a-martensite. Therefore, the ferromagnetism of the impact-treated
surface layer is obviously related with its amorphous state, as in amorphous iron-
based ribbons (see e.g. [19]).
A physical reason for the formation of an amorphous structure in austenitic
steels with carbon or carbon ? nitrogen under heavy impact treatment needs a
special discussion because it does not occur in similarly treated austenitic steels
free of interstitials (see e.g. [111]). This phenomenon can be explained in terms of
an increased concentration of vacancies in the interstitial solid solution. Smirnov
[112] was the first to theoretically predict that the dissolution of interstitial atoms
in metals increases the thermodynamic equilibrium concentration of vacancies.
Later on, the formation of superabundant vacancies in Ni, Pd and Fe hydrides
under a high hydrogen pressure has been proven by Fukai et al. [113, 114]. An
increase in the concentration of vacancies caused by dissolution of hydrogen in a
stainless austenitic steel was observed in [115]. Even earlier, McLellan [116] has
shown that the enthalpy of vacancy formation is decreased in the vicinity of carbon
atoms in the iron austenite.
The increased concentration of vacancies eases the formation of an amorphous
state. From this consideration, it is also understandable why the alloys intended for
obtaining the amorphous ribbons by rapid quenching from the liquid state usually
contain much interstitial boron, e.g. Fe80Si4B14.
3.2 Structural Change by Loading 73

Fig. 3.45 Cross-head displacement and specimen temperature in tension/compression tests of


CN96 at ra = ±450 MPa in dependence of the number of cycles. Specimen 1 was run for
N = 1000, specimen 2 for N = 2300 and specimen 3 until fracture at Nf = 3800. The Vickers
hardness HV30 is given in brackets: 1, 2 were measured on the specimen surface, 3 on the fatigue
fracture face

Thus, based on the obtained results, one can conclude that the mechanism of
surface hardening by impact treatment of CN steels, as well as of Hadfield steel, is
based on the formation of a complex amorphous, nanocrystalline and thin-twinned
austenitic structure and is not related with a martensitic transformation in the
surface.

3.2.4.2 Push/Pull Fatigue

Fatigue tests allow to investigate what fraction of the tensile strength may be
retained after cyclic loading. Several criteria are used to describe fatigue proper-
ties: fatigue limit, fatigue life, fatigue crack growth rate. A detailed analysis of
fatigue curves measured in the coordinates of stress amplitude/number of cycles is
given in Sect. 4.1.7. Described here is the substructure of steel CN96 developed at
different stages of the fatigue curve obtained in tension/compression tests.
A sample of steel CN96 was run at a stress amplitude of ra = ±450 MPa until
its fracture at Nf = 3800 (Fig. 3.45). The states 1–3, corresponding to different
parts of the fatigue curve, were taken for TEM studies of the structure at the
surface and in the centre of the tested specimens.
The initial decrease in the displacement of the cross-heads indicates cyclic
hardening. In the final part the displacement increases because of crack initiation
and growth. This observation of cyclic hardening in the CrMnCN steel is at some
variance with the data of Sun et al. [117] for austenitic CrMnN steels, where the
absence of hardening was shown in the whole range of fatigue life. Generally,
cyclic softening prevails if cross slip of dislocations is prevented, which assists the
74 3 Structure

Fig. 3.46 Structure of steel CN96 after 1000 cycles (state 1 in Fig. 3.45, the sample is taken
from the specimen surface). a TEM image. b Electron diffraction with reflections of austenite and
orthorhombic nitride f-Fe2N; zone axes ½0
11c and ½5
4
21

formation of planar dislocation arrays and reversible gliding of dislocations within


narrow slip bands [118, 119]. Such a dislocation structure is just typical for the
nitrogen and nitrogen ? carbon austenitic CrMn steels.
As the fatigue of austenitic nitrogen CrNi steels is concerned, it depends on the
value of strain amplitudes. At low strain amplitudes, planar slip and, corre-
spondingly, softening occurs within the whole fatigue life, whereas, if the strain
amplitude increases, planar slip with hardening occurs within 20–100 cycles and
thereafter is substituted by dislocation subcell structures and softening [120–122].
Thus, despite a similar dislocation distribution, the effect of alloying C ? N on the
fatigue of austenitic CrMn steels is a bit different from that of N.
The following structural change accompanies the fatigue tests of which results
are presented in Fig. 3.46. The initial structure of the tested steel CN96 is char-
acterized by a fully austenitic state and a moderate density of dislocations of about
1010 cm-2. After 1000 cycles (state 1 in Fig. 3.45), the density of dislocations was
increased to between 3 9 1010 and 5 9 1010 cm-2 in the specimen centre and
between 5 9 1010 and 7 9 1010 cm-2 at the surface. Planar slip is clearly
observed. Twinning was also occasionally met in one or two systems, corre-
spondingly. The most distinctive feature is the precipitation of fine iron nitride
f-Fe2N particles, as identified by diffraction (Fig. 3.46b).
After 2300 cycles (state 2 in Fig. 3.45), the structure becomes more or less
homogeneous in the cross section of the sample. The density of dislocations
increases up to 1011 cm-2. Along with nitride precipitates in the austenite, new
observations are the c ? e transformation in one (Fig. 3.47) or two (Fig. 3.48)
intersecting systems and the precipitation of cementite within the e-martensite
plates.
3.2 Structural Change by Loading 75

Fig. 3.47 Structure of steel CN96 after 2300 cycles at ra = ±450 MPa (state 2 in Fig. 3.45).
a TEM image of austenite with nitride precipitates and a plate of e-martensite containing
cementite precipitates. b Reflections of h-cementite and e-martensite in the diffraction pattern
obtained from the e-plate; zone axes ½2
1 2
2e and ½54h

As expected (see, e.g., [122]), a-martensite is formed at the intersection of e-


plates and here the cementite precipitates of 20–50 nm in size are larger than in the
e-martensite plates. Precipitates of f–Fe2N nitride, 2–3 nm in size, are dispersed in
the austenite.
After fracture at 3800 cycles (state 3), some areas depleted of dislocations
appear in the structure along with those of a high density in the range of 1011
cm-2. Again f–Fe2N nitrides are observed in the austenite (Fig. 3.49).
The structure presented above after tension/compression tests resembles in
many features that developed during unidirectional loading: dislocations, twins
and e-plates, all increasing their concentration as the cyclic straining proceeds.
But at the same time, the precipitation of Fe2N in austenite and of Fe3C in e-
martensite and in a-islands transformed at the intersection of e-plates are unique
features of the CrMnCN steel after cyclic loading. It is natural that, because of a
smaller carbon solubility in a bcc lattice, the cementite precipitates in the a-phase
are coarser than in e. Of course, these a-islands are too small and scarce to affect
the fatigue life.
It is a curious question, why cementite is precipitated at the e-martensite plates,
whereas the nitride precipitates are nearly evenly distributed in the austenitic bulk.
One should note in this relation that carbon atoms have a strong affinity to grain
and sub-boundaries in austenite and ferrite [123–125]. Therefore, they are pref-
erentially segregated at stacking faults which are prerequisites for the formation of
e-plates in the deformed austenite. In contrast, nitrogen atoms tend to segregate
much less at interfaces [123, 124, 126]. However, causing larger distortions in the
76 3 Structure

Fig. 3.48 State 2 continued. a TEM image of intersecting e-plates formed in different
crystallographic systems. b Diffraction pattern from the intersection site showing austenite,
a-martensite and cementite; zone axes [115]a and ½2
54h . c Cementite particles in the a- and e-
phases shown in a dark field image obtained by reflection ½20
1h

austenite and having a larger enthalpy of binding to dislocations, they are more
able to form clouds around dislocations [68]. For this reason, they precipitate
preferentially within the austenitic solid solution and, due to their small size, add
to the cyclic strength.
Why Fe3N is precipitated during aging after unidirectional strain (see e.g.
[127]), but Fe2N during cyclic loading, remains unclarified. Cyclic recovery may
enhance the annihilation of dislocations of opposite Burgers vector merging their
interstitial clouds, thus raising the local concentration of nitrogen atoms, which
would favour Fe2N.
After accumulating a critical cyclic strain, cracks are initiated which ‘‘soften’’
the specimen and lead to an increase of the displacement until failure, while the
hardness still increases. However, the dislocation structure breaks down locally
(Fig. 3.49) indicating the start of cyclic softening of the material.
3.2 Structural Change by Loading 77

Fig. 3.49 Structure of steel CN96 fractured after 3800 tension/compression cycles at
ra = ±450 MPa (specimen 3 in Fig. 3.45). a TEM image showing an uneven distribution of
dislocations and precipitates of f-nitride Fe2N. b Diffraction pattern; zone axes [112]c and
[-11.5 9 -4]f

Based on work by Feltner and Laird [128], Lukas and Klesnil [129] reviewed
the microstructural features of substitutional fcc Cu-, Al- and Fe-alloys after
cycling in dependence of the stacking fault energy c and the cycles to failure Nf.
They defined the areas A, B and C in their plot of c over Nf. As to the structure,
steel CN96 belongs to area C comprising planar slip, twinning and e-martensite.
However, its stacking fault energy of 40 mJ/m2 is too high to fit into the scheme.
The high interstitial content of this new alloy affects both, c and Nf. In his study on
the cyclic deformation response of substitutional Cu-alloys, Wang [130] proposed
a criterion for the transition from wavy to planar slip based on the interrelation of c
and the concentration of free electrons per atom e/a. The present high interstitial
steels belong to the planar slip type, but in contrast to the respective Cu-alloys
their c is inversely proportional to e/a (see Fig. 3.19). Both results seem to indicate
a difference between interstitial and substitutional fcc solid solutions in respect to
their response to cyclic loading.

References

1. Shanina BD, Gavriljuk VG, Konchitz AA, Kolesnik SP (1998) The influence of
substitutional atoms upon the electron structure of the iron-based transition metal alloys.
J Phys Condens Matter 10:1825–1838
2. Hohenberg H, Kohn W (1964) Inhomogeneous electron gas. Phys Rev B 136(3):864–871
3. Kohn W, Sham LJ (1965) Self-consistent equations including exchange and correlation
effects. Phys Rev A 140(4):1133–1138
78 3 Structure

4. Blaha P, Schwarz K, Madsen GKH, Kvasnicka D, Luitz J (2001) Wien2 k, an augmented


plane wave ? local orbitals program for calculating crystal properties. Karlheinz Schwarz
Techn. Universität, Wien, Austria, ISBN 3-9501031-1-2
}
5. Seith W, Kubaschewski O (1935) Die elektrolytische Uberführung von Kohlenstoff in festen
Stahl. Zs Electrochem 41(7b):551–558
}
6. Seith W, Daur Th (1938) Elektrische Uberführung in festen Metallegierungen. Zs
Electrochem 44(4):256–260
7. Gavriljuk VG, Shanina BD, Baran NP, Maximenko VM (1993) Electron-spin-resonance
study of electron properties in nitrogen and carbon austenite. Phys Rev B 48(5):3224–3231
8. Shanina BD, Kolesnik SP, Konchitz AA, Gavriljuk VG, Smouk SY, Tarasenko AV (1994)
The influence of nitrogen on the paramagnetic properties of the multicomponent d-element
iron-based alloy. Solid State Commn 90(2):109–113
9. Shanina BD, Gavriljuk VG, Konchitz AA, Kolesnik SP, Tarasenko AV (1995) Exchange
interaction between electron subsystems in iron-based F.C.C. alloys doped by nitrogen or
carbon. Physica Stat Sol (a) 149:711–722
10. Shanina B, Gavriljuk V, Berns H, Schmalt F (2002) Concept of a new high-strength low-
cost stainless steel. Steel Res 73(3):105–113
11. Gavriljuk V, Rawers J, Shanina B, Berns H (2003) Nitrogen and Carbon in Austenitic and
Martensitic steels: atomic interactions and structural stability. Mater Sci Forum 426–
432:943–950
12. Shanina BD, Gavriljuk VG (2004) Effect of carbon and nitrogen on electronic structure of
steel. J Steel Relat Mater suppl ‘‘High Nitrogen Steels 2004’’, pp 45–52
13. Shanina BD, Gavriljuk VG, Berns H (2007) Atomic interactions in stainless austenitic
CrMn steels alloyed with C, N or (C ? N). Mater Sci Forum 539–543:4993–4998
14. Berns Hans, Gavriljuk Valentin G (2007) Steel of highest fracture energy. Key Eng Mater
345–346:421–424
15. Berns H, Gavriljuk VG, Riedner S, Tyshchenko A (2007) High strength stainless austenitic
CrMnCN steels—part I: alloy design and properties. Steel Res Intern 78(9):710–715
16. Gavriljuk VG, Razumov O, Petrov Yu, Surzhenko I, Berns H (2007) High strength stainless
austenitic CrMnCN steels—part II: structural changes by repeated impacts. Steel Res Intern
78(9):716–719
17. Shanina BD, Gavriljuk VG, Berns H (2007) High strength stainless austenitic CrMnCN
steels—part III: electronic properties. Steel Res Intern 78(9):720–724
18. Gavriljuk VG, Shanina BD, Berns H (2008) Ab initio development of a high-strength
corrosion-resistant austenitic steel. Acta Mater 56(18):5071–5082
19. Mogilny GS, Shanina BD, Maslov VV, Nosenko VK, Shevchenko AD, Gavriljuk VG
(2011) Structure and magnetic properties of rapidly quenched FeSiB ribbons. J Non-
Crystalline Solids 357(16–17):3237–3244
20. Sozinov AL, Balanyuk AG, Gavriljuk VG (1997) C–C interaction in iron-base austenite and
interpretation of Mössbauer spectra. Acta Mater 45(1):225–232
21. Sozinov AL, Balanyuk AG, Gavriljuk VG (1999) N–N interaction and nitrogen activity in
the iron base austenite. Acta Mater 47(3):927–935
22. Balanyuk AG, Gavriljuk VG, Shivanyuk VN, Tyshchenko AI, Rawers J (2000) Mössbauer
study and thermodynamic modeling of Fe-C-N alloy. Acta Mater 48(15):3813–3821
23. Rawers JC (1999) High carbon-nitrogen iron alloys. J Mater Sci 34(5):941–944
24. Gavriljuk VG, Tarasenko AV, Tyshchenko AI (2000) Low temperature ageing of the freshly
formed Fe-C and Fe-N martensites. Scripta Mater 43(3):233–238
25. Shanina BD, Tyshchenko AI, Glavatskyy IN, Runov VV, Petrov YuN, Berns H, Gavriljuk
VG (2011) Chemical nano-scale homogeneity of austenitic CrMnCN steels in relation to
electronic and magnetic properties. J Mater Sci 46(24):7725–7736
26. Vonsovsky VS (1971) Magnetism (in Russian). Nauka, Moscow
27. Noskova NI, Pavlov VA, Nemnonov SA (1965) A correlation between the stacking fault
energy and structure of metals (in Russian). Physics Metals Metallogr 20(6):920–924
References 79

28. Ruff AW Jr (1970) Measurement of stacking fault energy from dislocation interactions.
Metall Trans A 1(9):2391–2413
29. Garner FA, McCarthy JM (1990) Spinodal-like decomposition of Fe-Ni and Fe-Ni-Cr
‘‘invar’’ alloys during neutron or ion irradiation. In: Russel KS, Smith DF (eds) Physical
metallurgy of controlled expansion invar-type alloys. TMS-AIME, Warendale, pp 187–206
30. Rotman F, Gilbon D, Lorant H, Dimitrov O (1987) Radial redistribution of elements in
austenitic Fe-Cr-Ni based alloys under HVEM Irradiation. Mater Sci Forum 15–18:1409–
1414
31. Gavriljuk V, Petrov Yu, Shanina B (2006) Effect of nitrogen on the electron structure and
stacking fault energy in austenitic steels. Scripta Mater 55(6):537–540
32. Gavriljuk VG, Shanina BD, Berns H (2000) On the correlation between electron structure
and short range atomic order in iron-based alloys. Acta Mater 48(15):3879–3893
33. Gavriljuk VG, Tyshchenko AI, Bliznuk VV, Yakovleva IL, Riedner S, Berns H (2008) Cold
work hardening of high-strength austenitic steels. Steel Res Intern 79(6):413–422
34. Hadfield RA (1925) Metallurgy and its influence on modern progress, with a survey of
education and research. Chapman and Hall Ltd, London, pp 91–101
35. Hall JH (1929) Trans AIME 84:382–427
36. Krivobok VN (1929) Investigation on the microstructure of Hadfield manganese steel. Trans
Amer Soc Steel Treat 15(6):893–928
37. Berns H, Klärner HF, Schmidtmann E (1967) Einfluß der Wärmebehandlung und Legierung
auf die Kaltverfestigung von Manganhartstahl. Arch. Eisenhüttenwesen 38(7):547–553
38. White CH, Honeycombe RWK (1962) Structural changes during deformation of high-purity
Fe-Mn-C alloys. JISI 200(4):457–466
39. Colette G, Crussard C, Kohn A, Plateau J, Pomey G, Weisz M (1957) Contribution à l’étude
des transformations des austénites à 12 % Mn. Revue de Metallurgie 6:433–481
40. Janssen K, Jellinghaus W (1956) Perlit- und Martensitbildung in Mangan-Hartstahl. Arch.
Eisenhüttenwesen 27(4):573–578
41. Berns H (1968) Entkohlung im Manganhartstahl Z. Wirtschaftl. Fertigung 63(9):437–441
42. Raghaven KS, Sastri SA, Marcinkowsky MJ (1969) Nature of work hardening in Hadfield
steel. Trans TMS-AIME 245:1569–1575
43. Dastur YN, Leslie WC (1981) Mechanism of work hardening in Hadfield manganese steel.
Metall Trans A 12(5):749–759
44. Matosyan MA, Golikov VM (1970) Zashitniye Pokritiya na Metallakh (in Russian). Kiev
Naukova Dumka 3:58–59
45. Karaman I, Huseyin S, Ken G (1998) On the deformation mechanisms in single crystal
hadfield manganese steels. Scripta Mater 38(6):1009–1015
46. Tsakiris V, Edmonds DV (1999) Martensite and deformation twinning in austenitic steels.
Mater Sci Eng A 273–275:430–436
47. Uggowitzer PJ, Harzenmoser M (1989) Strengthening of austenitic steels by nitrogen. In:
Foct J, Hendry A (eds) High nitrogen steels, HNS 88:174–79. Institute of Metals, London
48. Gavriljuk VG, Duz’ VA, Yephimenko SP (1990) Structure and mechanical properties of
cold-worked high-nitrogen austenite. In: Stein G, Witulsky H (eds) High nitrogen steels,
HNS 90, pp 100–103. Stahl&Eisen, Düsseldorf
49. Müllner P, Solenthaler C, Ugowitzer P, Speidel MO (1993) Martensite and deformation
twinning in austenitic steels. Mater Sci Eng A 164:164–169
50. Müllner P, Solenthaler C, Ugowitzer P, Speidel MO (1994) Brittle fracture in austenitic
steel. Acta Metal Mater 42(7):2211–2217
51. Müllner P, Solenthaler C, Speidel MO (1994) Second order twinning in austenitic steel.
Acta Metal Mater 42(5):1727–1732
52. Kubota S, Xia Y, Tomota Y (1998) Work-hardening behavior and evolution of dislocation-
microstructures in high-nitrogen bearing austenitic steels. ISIJ Intern 38(5):474–481
53. Saller G, Spiradek-Hahn K, Sgheu C, Clemens H (2006) Microstructural evolution of Cr–
Mn–N austenitic steels during cold work hardening. Mater Sci Eng A 427:246–254
80 3 Structure

54. Copley SM, Kear BH (1968) The dependence of the width of a dissociated dislocation on
dislocation velocity. Acta Metal 16(2):227–231
55. Kestenbach HJ (1977) The effect of applied stress on partial dislocation separation and
dislocation substructure in austenitic stainless steels. Phil Mag 36(6):1509–1515
56. Lee T-H, Oh C-S, Kim S-J, Takaki S (2007) Brittle fracture in austenitic steel. Acta Mater
55(11):3649–3662
57. Dorn John E, Raynak Stanley (1964) Nucleation of kink pairs and the Peierls mechanism of
plastic deformation. TMS AIME 230(5):1052–1064
58. Cottrell AH, Bilby BA (1949) Dislocation theory of yielding and strain ageing of iron. Proc
Phys Soc A 62(1):49–62
59. Keh AS (1963) In: Relation between structural and mechanical properties of metals. H. M.
Stationery Office, London, p 436
60. Baird JD (1963) Strain aging of steel-A critical review. Iron Steel 36:450–457
61. Hirth JP, Lothe J (1968) Theory of dislocations, Ch. 18, McGraw-Hill Book Co., New York,
pp 433–472
62. Vitek V, Kroupa F (1966) Dislocation theory of slip geometry and temperature dependence
of flow stress in B.C.C. metals. Phys Stat Sol a 18(2):703–713
63. Grujicic M (1994) The core structure of (a/2) \ 110 [ screw dislocations in Fe–Ni–Cr–N–
austenite. Mater Sci Eng A 183:223–232
64. Nyilas A, Obst B, Nakajima H (1993) Tensile properties, fracture and crack growth of a
nitrogen strengthened new stainless steel (Fe-25Cr-15Ni-0.35N) for cryogenic use. In:
Gavriljuk VG, Nadutov VM (eds) High nitrogen steels, HNS 93, Institute for Metal Physics,
Kiev, pp 339–344
65. Gavriljuk VG, Sozinov AL, Foct J, Petrov YuN, Polushkin YuA (1998) Effect of nitrogen
on the temperature dependence of the yield strength of austenitic steels. Acta Mater
46(3):1157–1163
66. Seeger A (1955) The generation of lattice defects by moving dislocations, and its
application to the temperature dependence of the flow stress of fcc crystals. Phil Mag
46(382):1194–121
67. Gavriljuk VG, Berns H, Escher Ch, Glavatskaya NI, Sozinov AL, Petrov YuN (1999) Grain
boundary strengthening in austenitic nitrogen steels. Mater Sci Eng A 271:14–21
68. Gavriljuk VG, Duz’ VA, Yephimenko SP, Kvasnevski OG (1987) Interaction between
carbon/nitrogen atoms and dislocations in austenite (In Russian). Phys Metals Metallogr
64(6):1132–1135
69. Chen SR, Tang D (1990) Effect of interstitial atom concentration on lattice parameters of
martensite and retained austenite in iron-carbon-nitrogen alloys. Mater Sci Forum 56–
58:201–206
70. Gavriljuk VG, Berns H (1999) High Nitrogen steels. Springer, Berlin
71. Norström LA (1977) The influence of nitrogen and grain size on yield strength in type AISI
316L austenitic stainless steel. Metal Sci 11(6):208–112
72. Degallaix S, Foct J, Hendry A (1986) Mechanical behaviour of high-nitrogen stainless
steels. Mater Sci Technol 2(9):946–950
73. Varin RA, Kurjidlowski KJ (1988) The effects of nitrogen content and twin boundaries on
the yield strength of various commercial heats of type 316 austenitic stainless steel. Mater
Sci Eng A 101:221–226
74. Werner E (1988) Solid solution and grain size hardening of nitrogen-alloyed austenitic
steels. Mater Sci Eng A 101:93–98
75. Uggowitzer PJ, Speidel MO (1990) Ultrahigh-strength austenitic steels. In: Stein G,
Witulsky H (eds) High Nitrogen Steels, HNS 90. Stahl & Eisen, Düsseldorf, pp 156–160
76. Obst B, Nyilas A (1991) Ultrahigh-strength austenitic steels. Mater Sci & Eng A137:141–
150
77. Adler PH, Olson GB, Owen WS (1986) Strain hardening of Hadfield manganese steel.
Metal Trans A 17(10):1725–1737
References 81

78. Petrov YuN (1978) Defects and diffusionless transformation in steel (in Russian). Naukova
dumka, Kiev
79. Schramm RE, Reed RP (1975) Stacking fault energies of seven commercial austenitic
stainless steels. Metall Trans A 6(7):1345–1351
80. Riedner S (2009) Höchstfeste nichtrostende austenitische CrMn-Stähle mit (C ? N). Doctor
thesis. Ruhr Universität Bochum, Lehrstuhl Werkstofftechnik, Bochum
81. Seeger A (1954) The temperature dependence of the critical shear stress and of work-
hardening of metal crystals. Phil Mag 45(366):771–773
82. Akdut N, Keichel J, Foct J (1997) The influence of nitrogen and orientation on the rolling
deformation mechanisms of austenitic single crystals. Steel Res 68(11):495–500
83. Speidel HJC, Speidel M (2003) Nickel and chromium based high nitrogen alloys. In:
Speidel MO, Kowanda C, Diener M (eds) High Nitrogen steels, HNS 2003. Institute of
Metallurgy Swiss Federal Institute of Technology, Zürich, pp 101–112
84. Sassen J, Carratt-Reed AJ, Owen WS (1989) Electron microscopy of austenitic Fe-Ni-Cr
alloys containing nitrogen. In: Foct J, Hendry A (eds) High Nitrogen steels., HNS
88Institute of Metals, London, pp 159–162
85. Sandström R, Bergqvist H (1977) Temperature dependence of tensile properties and
strengthening of nitrogen alloyed austenitic stainless steels. Scand J Metallurgy 6:156–169
86. Nilsson J-O, Thorvaldsson T (1985) The influence of nitrogen on microstructure and
strength of a high-alloy austenitic stainless steel. Scand J Metallurgy 15:83–89
87. Byrnes MLG, Grujicic M, Owen WS (1987) Nitrogen strengthening of a stable austenitic
stainless steel. Acta Metall 35(7):1853–1862
88. Reed RP, Simon NJ (1989) Nitrogen strengthening of austenitic stainless steels at low
temperatures. In: Foct J, Hendry A (eds) High Nitrogen steels, HNS 88. Institute of Metals,
London, pp 180–188
89. Grujicic M, Owen WS (1995) Models of short range order in a face-centered cubic Fe-Ni-Cr
alloy with a high concentration of nitrogen. Acta Metall Mater 43(11):4201–4211
90. Grujicic M (1995) The effect of nitrogen on the structure and mobility of dislocations in Fe-
Ni-Cr austenite. J Mater Sci 30(22):5799–5807
91. Obst B (1998) Basic aspects of tensile properties. In: Seeber B (ed) Handbook of applied
superconductivity, 2, F1.11, Institute of Physics Publishing, Bristol-Philadelphia, pp 969–
993
92. Schmalt F (2005) Nutzung der Löslichkeit von C ? N in nichtrostenden Stählen. Doctor
thesis. Ruhr Universität Bochum, Lehrstuhl Werkstofftechnik, Bochum-Düsseldorf
93. Tomota Y, Nakano J, Xia Y, Inoue K (1998) Unusual strain rate dependence of low
temperature fracture behaviour in high nitrogen bearing austenitic steels. Acta Mater
46(9):3099–3108
94. Tobler RL, Meyn D (1988) Cleavage-like fracture along slip planes in Fe-18Cr-3Ni-13Mn-
0.37 N austenitic stainless steel at liquid helium temperature. Metall Trans A 19(6):1626–
1631
95. Harzenmoser MAE (1990) Massive aufgestickte austenitisch-rostfreie Stähle und
Duplexstähle. Doctoral thesis. Eidgenössische Technische Hochschule, Zürich
96. Ishizaki J, Orita K, Terao K (1992) The influence of chemical composition and pre-strain on
impact toughness transition behaviour of 18 %MN-18 %Cr-N austenitic steels. J ISIJ (Tetsu
to Hagane) 78(12):1846–1853
97. Uggowitzer PJ, Paulus N, Speidel MO (1992) Ductile-to-brittle transition in nitrogen
alloyed austenitic stainles steels. In: Nordberg H, Börklund J (eds) Application of stainless
steels’92. The Institute of Metals ASM Intern, Stockholm, pp 62–72
98. Tomota Y, Endo S (1990) Cleavage-like fracture at low temperatures in an 18Mn-18Cr-
0.5 N austenitic steel. ISIJ Intern 30(8):656–662
99. Müllner P (1997) On the ductile to brittle transition in austenitic steel. Mater Sci Eng A
234–236:94–97
100. Tomota Y, Xia Y, Inoue K (1998) Mechanism of low temperature brittle fracture in high
nitrogen bearing austenitic steels. Acta Mater 46(5):1577–1587
82 3 Structure

101. Gavriljuk VG, Shivanyuk VN, Shanina BD (2005) Change in the electron structure caused
by C, N and H atoms in iron and its effect on their interaction with dislocations. Acta Mater
53(19):5017–5024
102. Larikov LN, Falchenko VM, Mazanko VF, Gurevich SM, Kharchenko GI, Ignatenko AI
(1975) Abnormal acceleration of diffusion during impulse loading of metals. Rep Acad Sci
USSR (in Russian) 221:1073–1075
103. Pogorelov AE, Ryaboshapka KP, Zhuravlev AF (2002) Mass transfer mechanism in real
crystals by pulsed laser irradiation. J Appl Phys 92(10):5766–5771
104. Karnaukhov IN, Pogorelov AE, Chernolevsky MS (2006) Non-activated mechanism of
mass transfer by moving dislocations (in Russian). Metallofiz Noveishie Technol
28(6):827–835
105. Bernauer J (2004) Einfluß von Kohlenstoff als Legierungselement in stickstofflegierten
Chrom-Mangan-Stählen. Doctor thesis. Eidgenössische Technische Hochschule, Zürich
106. Berns H, Riedner S, Gavriljuk V, Petrov Y, Weihrauch A (2011) Microstructural changes in
high interstitial stainless austenitic steels due to ballistic impact. Mater Sci Eng A 528(13–
14):4669–4675
107. Nakkalil R (1991) Formation of adiabatic shear bands in eutectoid steels in high strain rate
compression. Acta Metal Mater 39(11):2553–2563
108. Xu Y, Chen Y, Xiong J, Zhu J (2001) Acta Metal Sinica 37(2):165–170
109. Petrov YN, Gavriljuk VG, Berns H, Schmalt F (2006) Surface structure of stainless and
Hadfield steel after impact wear. Wear 409(6):687–691
110. Gavriljuk VG, Tyshchenko AI, Razumov ON, Petrov YuN, Shanina BD, Berns H (2006)
Corrosion-resistant analogue of Hadfield steel. Mater Sci Eng A 420(1–2):47–54
111. Prokopenko GI, Petrov YuN, Vasilyev MA, Trophimova LN, Bliznyuk VV (2008)
Structural-phase transformations in austenitic steel under ultrasonic impact surface
treartment (in Russian). Metal Phys Adv Technol 30(1):115–131
112. Smirnov AA (1991) Theory of vacancies in interstitial alloys (in Russian). Metal Phys
13(9):40–44
113. Fukai Y, Okuma N (1993) Evidence of copious vacancy formation in Ni and Pd under a
high hydrogen pressure. Jpn J Appl Phys part 2, 32(9A):L1256–1259
114. Fukai Y (2003) Formation of superabundant vacancies in M-H alloys and some of its
consequences: a review. J Alloys Compd 356–357:263–269
115. Gavriljuk VG, Bugaev VN, Petrov YuN, Tarasenko AV, Yanchitsky BZ (1996) Hydrogen-
induced equilibrium vacancies in FCC iron-base alloys. Scripta Mater 34(6):903–907
116. McLellan RB (1988) The thermodynamics of interstitial-vacancy interactions in solid
solutions. J Phys Chem Solids 49(10):1213–1217
117. Sun H, Diener M, Uggowitzer PJ, Speidel MO (1990). Low cycle fatigue behaviour of high
nitrogen steels. In: Stein G, Witulski H (eds) High Nitrogen steels, HNS 90. Stahl & Eisen,
Düsseldorf, pp 220–223
118. Grosskreutz JC (1972) Strengthening and fracture in fatigue (Approaches for achieving high
fatigue strength). Metall Trans 3(5):1255–1262
119. Margolin H, Mahajan Y, Saleh Y (1976) Grain boundaries, stress gradients and fatigue
crack initiation. Scripta Metall 10(12):1115–1118
120. Taillard R, Foct J (1989) Mechanisms of the action of nitrogen interstitials upon low cicle
fatigue behaviour of 316 stainless steels. In: Foct J, Hendry A (eds) High Nitrogen Steels,
HNS 88. The Institute of Metals, London, pp 387–391
121. Degallaix S, Dickson JI, Foct J (1989) Effect of nitrogen on fatique and creep-fatigue
behaviour of austenitic stainless steels. In: Foct J, Hendry A (eds) High Nitrogen Steels,
HNS 88. Institute of Metals, London, pp 380–386
122. Venables JA (1962) The martensite transformation in stainless steel. Phil Mag 7(1):35–44
123. Rudy ML, Huggins RA (1966) Grain boundary segregation and the cold work peak in iron
containing carbon and nitrogen. TMS AIME 236(12):1662–1666
124. Lagerberg G, Josefsson A (1955) Influence of grain boundaries on the behaviour of carbon
and nitrogen in a-iron. Acta Metal 3(5):236–244
References 83

125. Petrov YuN (1993) On the carbon distribution at structural imperfections in manganese
austenite. Scripta Metall et Mater 29(11):1471–1476
126. Petrov YuN, Gavriljuk VG, Berns H, Escher Ch (1999) Nitrogen partitioning between
matrix, grain boundaries and precipitates in high-alloyed austenitic steels. Scripta Mater
40(6):669–674
127. Gavriljuk VG, Stein G, Berns H (2003) Structural stability of austenitic CrMn(Mo)N steels
for high-strength, corrosion-resistant retaining rings. Steel Res Int 74(7):444–452
128. Feltner CE, Laird C (1968) Factors influencing the dislocation structure in fatigue metals.
TMS AIME 242(8):1253–1257
129. Lukas P, Klesnil M (1972) In: Devereux O, Owen F (eds) National Association of Corrosion
Engineers. Houston, pp 118–132
130. Wang Z (2004) Cyclic deformation response of planar-slip materials and a new criterion for
the wavy-to-planar-slip transition. Phil Mag 84:351–379
Chapter 4
Properties

This chapter deals with key properties relevant in service. Those important for
manufacturing are discussed in Chap. 5. The position of new HIS is revealed by
comparing them to reference steels given in Table 2.1. The mechanical properties
are in the center of attention, because steels of higher strength are to be developed.
These stainless grades require a corrosion resistance comparable to some standard
steels which has to be proven by respective tests. The high austenite stability of
HIS poses the question, if they are useful for nonmagnetic applications. Last, not
least, a higher strength of ductile austenite offers a chance to improve the resis-
tance to impact wear and cavitation. See [1] for more details.

4.1 Mechanical Properties

Tensile tests are carried out to characterise steels under slow uniform loading
within a wide range of temperature. Notch impact tests come with a multiaxial
stress state and increased velocity, which in connection with subzero testing
temperatures offer insight into impending embrittlement. Fatigue is responsible for
many failures and respective tests help to find out what fraction of tensile strength
is retained under cyclic loading.

4.1.1 Tensile Properties at Room Temperature

Specimens of Ø5 mm were tested according to EN10002-1 at room temperature


(RT) and a rate of 0.5 mm/min. The engineering stress–strain curves of hot worked
steels (longitudinal taking) are presented in Fig. 4.1a. From these, common prop-
erties of (i) proof, ultimate and true fracture strength (Rp0.2, Rm, R), (ii) ductility in

H. Berns et al., High Interstitial Stainless Austenitic Steels, 85


Engineering Materials, DOI: 10.1007/978-3-642-33701-7_4,
Ó Springer-Verlag Berlin Heidelberg 2013
86 4 Properties

Fig. 4.1 Stress strain curves


derived from tensile tests at
RT of new HIS and reference
steels, (a) engineering stress
r0 and elongation e, (b) true
stress r and true plastic strain
u up to Au, thin lines:
Ludwigson fit, see Table A2

terms of uniform and fracture elongation (Au, A), reduction of area (Z) and (iii)
toughness expressed by the specific fracture energy Ws (deduced from the area under
the curves in Fig. 4.1a) are derived and listed in Table A1 (Appendix A). Assuming
constant volume the engineering stress–strain curves are converted into true stress–
strain curves within the range of uniform elongation (Fig. 4.1b). A fit of the
Ludwigson equation [2]

r ¼ K1  un1 þ eK2 þn2 u ð4:1Þ

to the experimental data yielded the constants presented in Table A2. The first term
represents the Ludwik equation [3], which is adjusted to smaller strains by the
second term. The meaning of each constant is discussed e.g. in [2]. However, the
quality of the fit depends on the starting conditions, which in turn influences the
constants. Here, the procedure is used to compare the exponent n1 of cold work
hardening.
The accumulated tensile properties are used to compare the new HIS with ref-
erence steels (Fig. 4.2). CN96 starts e.g. from a proof strength of 600 MPa and work
hardens to a true fracture strength of more than 2500 MPa within 74 % of elongation
reaching a work hardening exponent of 0.92 (see Sect. 3.2.1). The HNS CrMnN
4.1 Mechanical Properties 87

Fig. 4.2 Comparing key tensile properties of new HIS (shaded area) with those of reference
steels (Table A1), Rp0.2 = proof strength, R = true fracture strength, n1 = exponent of cold
work hardening (Ludwigson), Ws = specific fracture energy, the steel number refers to the
chemical composition in Table 2.1

slightly exceeds this strength level but comes with a distinctly lower elongation.
Necking starts earlier and proceeds to a higher reduction of area (Table A1,
Fig. 4.1a). The high carbon grade MnC begins at a significantly lower proof strength,
work hardens intensely but fails with hardly any necking after only 46 % elongation.
As expected, the low interstitial steel CrNi offers the lowest proof strength, yet work
hardens steadily at n1 = 0.66 up to A = 83 %. The high manganese TWIP steel
MnCr77 surpasses this ductility and together with a higher strength level and work
hardening (n1 = 1) the highest fracture energy Ws is achieved.
The addition of (mass %) 1 Mo or 2 Cu to the new HIS lowers the strength level
only slightly but the reduction of area by about one third (Table A3). The fracture
energy remains quite high, though.
The interstitial content of HIS, saturated with about 0.6 mass % N at normal
pressure of air, is raised by adding carbon. This increases the strength of centrifugal
castings (Fig. 4.3, Table A4) while the ductility is diminished. The properties of a
sand casting are given for comparison. The strength level of castings is well below
that of forgings, because the grain size is considerably coarser (Table A5). In fact,
some of the high proof strength of CrMnN may be due to the fine-grained structure.
88 4 Properties

Fig. 4.3 Effect of interstitial


content on strength (Rp0.2,
Rm) and ductility
(Z = reduction of area) of
centrifugal castings GCN65
to GCN115,
encircled = results of sand
casting GCN85 (Table 2.1
and A4)

4.1.2 Tensile Properties at Subzero Temperatures

Testing temperatures TT to -100 °C cover the climatic range and some industrial
cooling operations. Embrittelment is known of HNS [1] and therefore two new
HIS were also tested at -196 °C. The proof strength increases as TT is lowered
(Fig. 4.4, Table A6, Sect. 3.2.2). The ductility of HIS and HNS is hardly affected
down to -100 °C, but then drops to one tenth in liquid nitrogen. This ductile to
brittle transition is not observed for CrNi. The Ludwigson fit for CN96 in Table A2
reveals that the exponent n1 of cold work hardening is raised only marginally, but
that eK2 increases considerably, which is in line with the enhancement of Rp0.2 at
lower TT. The true strength R grows as well which is reflected by a higher K1.
The fracture face of CN96 tested at RT is covered with ductile dimples
(Fig. 4.5). At -100 °C some intercrystalline fracture appears. The fracture in
liquid nitrogen is transcrystalline, i.e. cleavage-like brittle. This corresponds to the
low values of A and Z in Table A6. At -100 °C the ductility accumulated before
the onset of fracture counts. The brittle patches of intercrystalline fracture develop
as the crack passes by.
The as-cast steel GCN85 reacts to a decrease of TT in about the same way as the
hot worked grade CN85x but at a lower level of strength and ductility (Table A6).

4.1.3 Tensile Properties at Elevated Temperatures

Strengthening of cold worked HNS was observed by aging at 500 °C [4].


Carbonitride precipitation starts between 700 and 750 °C after 1 h as depicted in
Fig. 5.9. This poses the question, if HIS may be used at temperatures up to 700 °C
without severe embrittlement or a loss of austenite stability.
After heating, tensile specimens were soaked for 1 h before loading and tested at a
crosshead speed of 0.5 mm/min. The results of hot tensile test are listed in Table A7.
The engineering stress–strain curves of CN85 are smooth at 20, 400 and 500 °C but
show serrations at 300 °C accompanied by softening and strain aging at 400° C
4.1 Mechanical Properties 89

Fig. 4.4 Effect of subzero testing temperature TT on proof strength Rp0.2 and fracture elongation
A of HIS CN859, CN96, CN107 (shaded area, Table A6) and reference steels CrMnN and CrNi

(Fig. 4.6). A different type of waviness appears between 550 and 650 °C, the peaks of
which are farther apart than those at 300° C. According to [5] this Portevin-Le-
Chatelier effect is caused by interaction of dislocations and interstitial atoms at 300
°C and of substitutional atoms at 550–650 °C. The curve runs smoothly again at
700 °C. The start of precipitation at C550 °C is confirmed by intercrystalline cor-
rosion after etching of metallographic sections taken from the gauge length. How-
ever, no precipitates were found by SEM within the austenite grains nor on grain
boundaries of the neck at 600 °C (Fig. 4.7a). Even at 700 °C most grain boundaries
in the gauge length did not appear to be decorated (Fig. 4.7b). Here, the failure mode
turned from microvoid coalescence to intergranular cracking perpendicular to the
direction of stress as TT was increased from 500 to 700 °C.
90 4 Properties

Fig. 4.5 Fractography (SEM) of CN96 after tensile tests at a 22, b -100, c -196 °C

Fig. 4.6 Proof strength Rp0.2


and fracture elongation A of
CN85 in dependence of the
testing temperature TT (Table
A7)

In conclusion, precipitation seems to be responsible for a loss of ductility at


TT [ 500 °C but also for a strengthening effect (Fig. 4.6). The proof strength of
CN85 at 700 °C is as high as that of CrNi at RT.
Cold expanded retaining rings are shrink-fitted on generator shafts where they
are moderately heated in service. Table A8 answers the question what loss of
strength is to be expected of HIS, cold worked by 20 elongation and aged above
shrink temperature, by heating up to 150 °C. At 150 Rp0.2 is about 10 % lower
than at 100 °C.
4.1 Mechanical Properties 91

Fig. 4.7 Metallographic sections after hot tensile tests (CN85, SEM), a Longitudinal section
through neck showing pores (600 °C), b Transverse section from gauge length (700 °C),
GB = grain boundary

4.1.4 Creep Properties

The new HIS are not meant for long-time creep service, because too much
embrittling precipitation has to be anticipated. However, the short-time creep
behaviour up to several 100 h may be of interest for tooling applications. Steel CN85
was selected to cut down on precipitation. It was tested in the as-quenched state and
also after subsequent cold working by 20 % elongation at constant creep stresses and
TT of 600, 650 and 700 °C in a range of 15–420 h. According to [6] creep properties
are improved most by cold deformation between 10 and 30 %. The present inves-
tigation is of tentative nature because of a limited number of specimens.
At the example of tests at 650 °C and a stress of 350 MPa the effect of cold
working is explained in Fig. 4.8a,b The high initial creep rate of the solution
annealed state comes down to the minimum creep rate e_ min of 0.0614 %/h after a
creep elongation e(_emin ) of 5.39 %. In the cold worked condition the respective
values are 0.00841 %/h and 0.13 % which points to extensive strengthening by
precipitation. The time t(_emin ) to reach the minimum creep rate is, however, hardly
affected by cold working, which slightly raises the time to fracture tf but lowers the
elongation to fracture ef by more than an order of magnitude.
A metallographic section was taken from the gauge length of the solution
annealed specimen that fractured after 15 h at 650 °C, 400 MPa and 9 % elon-
gation (Fig. 4.8c). Large precipitates and creep cracks impaired thinning of
specimens for TEM. The figure was taken in SEM mode while selected area
diffraction (SAD) of transparent precipitates was used to identify the phases.
Carbides of type M7C3 and M23C6 have precipitated along grain and twin
boundaries, while a eutectoid of M2N type nitrides in austenite grows into the
grains. A dispersion of fine M3C carbides is distributed throughout the austenite
grains. No traces of bcc structure were found by X-ray diffraction, which implies
that the austenite remains stable in spite of less solute atoms. The embrittling effect
of grainboundary precipitation promotes respective creep cracks (Fig 4.8d). The
dispersed iron carbides are bound to contribute to the creep resistance but are too
fast growing to keep this effect up.
92 4 Properties

Fig. 4.8 Results of creep test at 650 °C, steel CN85 solution annealed (SA) or subsequently cold
worked by 20 % elongation (CW), a Creep rate e_ at a creep stress of 350 MPa in dependence of
creep elongation e defining the minimum creep rate e_ min , the elongation to reach it e(_emin ) and the
elongation to fracture ef, b creep rate e_ at the same stress in dependence of time t defining the time
to reach the minimum creep rate t(_emin ) and the time to fracture tf. c Microstructure of SA
specimen within the range of uniform creep elongation after fracture at 400 MPa, SEM, d As
before showing creep cracks, arrow indicating direction of creep stress, LOM

The Larson-Miller parameter PLM = T (C ? log tf) is a means to combine


creep results obtained at different temperatures. For T in K and tf in h those authors
demonstrated in [7] a good fit for ferritic or austenitic steels and even for non-
ferrous metals, if C = 20. Notwithstanding the subsequent discussion on C the
stress is plotted over PLM at C = 20 in Fig. 4.9 based on the results in Table A9. In
spite of the enhanced strengthening, PLM is hardly changed by cold working. The
results are clearly to the right of those measured for hot work tool steel
X38CrMoV5-1 (AISI H11) quenched and tempered to a tensile strength
Rm = 1425 MPa [8]. The better performance of fcc HIS in comparison to bcc tool
steel is based mainly on the lower rate of diffusion in close packed crystals.
Nucleation and growth of precipitates contribute to the creep behaviour as well.
4.1 Mechanical Properties 93

Fig. 4.9 Interdependence of


creep stress and Larson-
Miller parameter PLM
comparing CN85 solution
annealed (filled symbols) and
cold worked by 20 %
elongation (open symbols)
(Table A9) with hot work tool
steel H11 (X38CrMoV5-1)
QT, Rm = 1425 MPa at
20 °C, CN85 was tested
under constant stress, H11
under constant load [8]

4.1.5 Hardness

The macrohardness (ISO 6507) only moderately increases with the interstitial
content of centrifugal castings (Table A10). The respective values of the hot worked
HIS (270–278 HV30) are almost twice as high as of CrNi (141 HV30). The
macrohardness of MnC (211 HV30) stays below that of HIS, but the microhardness
on the tensile fracture face (741 HV0.1) exceeds all HIS and reference steels. Cold
work hardening is also evident after cold drawing CN96 (B423 HV30). Heating to
the precipitation range raises the macrohardness of CN85 if accompanied by
straining (B476 HV30).

4.1.6 Notch Impact Toughness

The V-notch impact bending energy KV was measured according to EN10045. The
specimens of hot worked steel were taken mostly in longitudinal direction, those of
as-cast steel perpendicular to the direction of transcrystallisation. A typical result of
HIS is shown in Fig. 4.10. Starting from a high toughness at RT a ductile-to-brittle
transition occurs as TT is lowered. The respective DBTT is taken at KV = 100 J,
which is still in the range of tough constructional steels.
Steels CN96 attains KV = 364 J which is the highest of all measured values
(Table A11). The respective steel GCN98 is the toughest of the as-cast grades at
KV = 317 J. In general the new HIS are as tough as or tougher than low-interstitial
CrNi which does not show a DBTT, though, but just a gradual decrease of KV as TT is
lowered. The DBTT of HIS is close to -90 °C except for the high carbon grades
CN107 and GCN115 which are just above -50 °C. In comparison the DBTT of the
high nitrogen grade CrMnN is up to -21 °C. At room temperature KV of CN107,
CN96 and CN85 is reduced by transverse taking but unexpectedly DBTT of CN85 is
reduced as well.
94 4 Properties

Fig. 4.10 ISO-V impact


toughness KV in dependence
of testing temperature TT
comparing steel CN96 and
CN107, the ductile-to-brittle
transition temperature DBTT
is taken at KV = 100 J
(Table A11)

The appearance of the fracture face turns from ductile dimples at RT to


cleavage-like brittle fracture at -196 °C (Fig. 4.11). In the transition range at
-80 °C dimple and intercrystalline areas are visible. Thus, about the same features
of microfractography are encountered as after tensile tests (Fig. 4.5), but the
ductile to brittle transition occurs at higher TT (see Sect. 3.2.3.1).

4.1.7 Rotating Bending Fatigue

Rotating bending tests were carried out at RT, a frequency of 40 Hz and constant
stress amplitudes ra until fracture or 107 number of cycles N. Specimens of
hourglass shape and 5 mm smallest diameter were ground in axial direction by a
flap wheel composed of abrasive cloth strips and loaded by rotating four-point-
bending [9]. Up to 7 specimens were run at preselected ra levels to allow a
statistical evaluation of the fatigue life Nf50 and the fatigue limit rf50 at N = 107
with a probability P = 50 % following the arcsin HP procedure [10]. Before the
investigation a single specimen was run at 50 Hz under the highest load and
observed by a thermo-camera. The temperature stayed below 40 °C until the very
final cycles to fracture when it rose to 74 °C. Polished specimens were stopped
after every 5000 cycles and searched for cracks by a microscope mounted on the
test rig. The number of cycles to crack initiation Ni was taken at a crack length of
0.2 mm and compared to the number of cycles to fracture Nf.
In the order of GCN85, CN96 and CN96 pre-streched by 20 % elongation the
fatigue limit increases from 229 to 415 to 552 MPa and, divided by the respective
proof strength of 447, 600 and 1022 MPa, we obtain rf50/Rp0.2 equal to 0.47, 0.69
and 0.54 (Fig. 4.12). Short cracks were initiated at slip bands on the specimen
surface which is best to be seen on the coarse grained as-cast specimens
4.1 Mechanical Properties 95

Fig. 4.11 Fractography (SEM) of CN96 after ISO-V impact test at a 22, b -80, c -196 °C

(Fig. 4.13b). They follow crystallographic planes until after two or three grains
they merge into the main crack which runs perpendicularly to the bending stress up
to fracture. The ratio of (Ni/Nf)100 is generally well above 50 % except for one
specimen where the crack started early, i.e. after 39 % of the fatigue life
(Fig. 4.13a). The results of this limited study suggest that the fatigue limit at
N = 107 attains about half of the proof strength. At higher stress amplitudes the
fatigue life is shortened and mostly accompanied by earlier crack initiation.

4.2 Wear Resistance

It is well known that wear changes the surface layer of materials, so that further
wear is governed by the surface properties. Ductile austenitic steels tend to
workharden in the wear surface and debris of wear-inducing counterbodies may
become inserted into the surface. These processes depend on the type of loading.
Abrasion by grooving wear comes with less workhardening than impact wear by
mineral particles. The implosion of bubbles during cavitation is comparable to a
particle impingement. Hadfield steel MnC has been used in applications where
impact wear prevails because of its capacity to workharden. The new HIS were
investigated in respect to their resistance to impact wear, abrasion and cavitation
because their rate of workhardening is close to that of MnC, which is not stainless,
though.
96 4 Properties

Fig. 4.12 Wöhler diagram of rotating bending fatigue tests plotting the stress amplitude ra over
the number of cycles N, full lines = fatigue life (50 % probability), dashed lines = fatigue limit
at N = 107 (50 % probability), digits = number of specimens not broken after N = 107

4.2.1 Abrasive Wear

Grooving wear by mineral particles is met e.g. in mining, processing and transpor-
tation of ore and rock. This abrasion was investigated by a pin-on-plate test. The end
face of area A of a pin specimen Ø 6 9 20 mm is moved under a load of 37 N at a
speed of 4.8 mm/s over fresh flint grinding paper of 80 or 220 mesh size for a
length L. The mass loss Dm is measured and converted to a loss of volume by means
of the density q to give the dimensionless wear resistance W-1 ab = qAL/Dm. The
results of pin-on-plate test are shown in Fig. 4.14. The wear resistance increases in
the order of CrNi, CN96, MnC and is higher against the finer 220 mesh abrasive. The
reciprocal value Wab is the wear rate (Table 4.1).

4.2.2 Impact Wear

A schematic representation of the impact wear test is given in Fig. 4.15a:


Individually counted particles of greywacke, mesh size 11–8 mm, impinge per-
pendicularly on a sample plate at a speed of 25 m/s. In fact, two samples are
mounted on a rotor in adjacent positions and the wear loss of both is summed up to
give the plot in Fig. 4.15b.
4.2 Wear Resistance 97

Fig. 4.13 Stages of fatigue, a Crack initiation at Ni (filled symbols) in percent of Nf (open
symbols) of one specimen per ra. The Wöhler lines are taken from Fig. 4.12 for comparison,
b Crack initiation at slip bands at the example of GCN85

Fig. 4.14 Abrasive wear resistance W-1


ab against 80 or 220 mesh flint paper
98 4 Properties

Table 4.1 Wear rate by abrasion Wab, wear rate by impact wear Wimp and wear rate by cavi-
tation Wcav of different steels after running in. The factor f stands for the improvement of the
wear resistance in relation to the low interstitial standard steel CrNi
Wab f Wimp f Wcav f
[10-5] [lg/impact] [lg/s]
CrNi 3.91 1 9.21 1 4.77 1
MnC 3.09 1.62 4.75 1.94 2.83 1.69
CN96 3.54 1.10 4.67 1.97 0.41 11.6
GCN115 - - - - 0.25 19.6

Fig. 4.15 Impact wear, a Schematic representation of impact conditions, b Mass loss depending
on the number of impacts (Table A12)

After a running-in period, in which the wear surface is formed, the wear loss
increases almost linearly, so that the wear rate Wimp can be derived (Table 4.1).
Greywacke debris are embedded in the wear surface of the most ductile steel
MnCr77 to such an extent that mass is picked up during the first thousand impacts.
The wear resistance of the new HIS is about as good as that of Hadfield steel MnC.
In fact all steels of high C, N or C ? N content lie in a rather narrow scatter band
of wear loss, while the low interstitial steel CrNi wears about twice as fast.
In a section normal to the wear surface the micro-hardness of a cold worked
layer is highest at the surface and then drops as the distance from the surface
grows. The hardness penetration is deepest for the softest steel CrNi where it
exceeds 1 mm (Table A12).
Cold work hardening is visible in the microstructure and described in Sect.
3.2.4.1. The most interesting feature of HIS and MnC is a thin amorphous top layer
followed by a nanocrystalline zone.
4.2 Wear Resistance 99

Fig. 4.16 Mass loss by


cavitation in dependence of
test duration (courtesy of S.
Huth)

4.2.3 Wear by Cavitation

A local change of pressure in fluid energy machines may cause gas bubbles which
subsequently implode. This is accompanied by small jets of fluid which exert
impacts on a surface and degrade it until small particles are detached. This cav-
itation was investigated by an ultrasonic sonotrode of Ø16 mm mounted vertically
on a piezo-quartz which oscillated at a frequency of 20 kHz with an amplitude of
40 lm dipped in destilled water at RT. Adjacent to the end face of the sonotrode a
specimen was mounted below at a distance of 0.5 mm. After certain intervals the
mass loss was measured. The results are plotted in Fig. 4.16. After an incubation
time the mass loss increases about linearly to give the wear rate Wcav (Table 4.1).
The lower the rate, the longer is the incubation time during which the surface is
cold worked [11]. At an equal interstitial content the mass loss rate of GCN115 is
more than an order of magnitude lower than that of Hadfield steel MnC. This is
based on the greater toughness of HIS reflected e.g. by the higher specific fracture
energy WS (Table A1, A4). The wear of CN96 proceeds more than an order of
magnitude slower than that of CrNi.

4.3 Corrosion Resistance

The corrosion resistance of stainless steels depends on a thin passive layer of


complex structure, simply called Cr2O3 [12]. It has to be build up by the sur-
rounding media and is under their constant attack which is influenced by chemical
composition, temperature and fluid flow. We discern between general corrosion of
evenly distributed mass loss and localised corrosion. It is therefore an extensive
task to characterise the corrosion resistance of a new material and we have to
confine the present study to a number of test examples. They comprise submersion
tests, current density-potential-tests and tests on intercrystalline corrosion.
100 4 Properties

4.3.1 Submersion Tests

Ground specimens of about 10 9 10 9 20 mm were submerged according to


DIN 50905 in aqueous solutions of H2SO4 or HCl at room temperature for 120 h
to measure the mass loss entailed by general corrosion. Initial tests showed, that
specimens, which had been stored in the office for some weeks, were fully pas-
sivated and not attacked by 10 % H2SO4. Therefore all further specimens received
a grinding finish by 1000 mesh SiC paper immediately before submersion. Usually
the test is interrupted every 24 h to clean and weigh the specimens and to renew
the test fluid. However, the exposure to air led to the passivation of some speci-
mens thus inhibiting further corrosion and stopping the evolution of hydrogen
bubbles. Therefore other tests were run for 120 h without interruption but with a
renewal of the test fluid every 24 h during which the specimens stayed submerged.
The mass loss Dm [g] is related to time and area to give the corrosion rate v [g/
m2h], which may be converted to the removal w [mm/a] per surface. The required
densities of the new HIS were measured at 7.6–7.67 g/cm3. Early passivation
means that there is no further mass loss and v as well as w decrease as time goes
on, e.g. to the technically accepted range of w B 0.1 mm/a.
The hot worked steels CN96, CN107 and the related alloys with Mo or Cu were
weighed after 24 h in aqueous solutions of 5 or 10 mass % H2SO4 and in
1 mass % HCl. The exposure to air caused passivation and prevented any further
mass loss or bubbling until the end of the 120 h interrupted test. Only CN103Mo1
met the target of w = 0.1 mm/a after 120 h in 5 % H2SO4. The remainder of
specimens would have done so only after prolonged test duration. In spite of
interruptions the specimens did not immediately passivate in 3 % HCl. The mass
loss ceased after the fourth weighing of the almost completely dissolved CN96
specimen and after the third of CN94Mo1. Steel CN96Cu2 continued to lose mass
until end of test. As for Mo the mass loss during active corrosion is lowered by Cu
(Fig. 4.17) [13].
Uninterrupted test were run with the same steels in 1 and 3 % HCl and 10 %
H2SO4. In the latter all specimens started with a strong effusion of hydrogen which
continued for CN96, was slowed down for CN107 and CN94Mo1, after 48 h almost
stopped in case of CN96Cu2 and completely ceased for CN103Mo. The corrosion
products fell off easily during handling of the specimens and dissolved again in the
fluid except for CN96Cu2. After removing the coat of corrosion products the surface
of this specimen appeared copper-coulored and 6 mass % Cu were measured by
EDX at 7 kV. The coat contained 36 mass % Cu (EDX, 15 kV). A protective layer of
Cu is apparently formed on the surface during early corrosion which impedes further
mass loss. But Mo and C reduce the corrosion rate as well (Fig. 4.18). However,
hydrogen is emitted in 3 % HCl throughout the test. The corrosion rates are high and
the effects of alloying less striking (Table A13).
The series of centrifugal castings was submitted to uninterrupted submersion
tests to evaluate the effect of carbon on corrosion in aqueous solutions of 1 and 3
HCl and 10 % H2SO4 [14]. In HCl the mass loss and corrosion rate were reduced
4.3 Corrosion Resistance 101

Fig. 4.17 Effect of Mo and


Cu on the mass loss Dm
during interrupted 120 h
submersion tests in an
aqueous solution of 3 % HCl

Fig. 4.18 Effect of C, Cu


and Mo on the removal w
after uninterrupted 120 h
submersion tests in an
aqueous solution of 10 %
H2SO4

as the carbon content increased (Table A14). Little effect of carbon is noticed in
H2SO4, especially if the test duration is prolonged 5 times (Fig. 4.19). Hydrogen
bubbles were observed at the beginning of all tests but the gas evolution had
stopped in 10 % H2SO4 and 1 % HCl, before the test solution was renewed after
the first 24 h. It did not start again throughout the remainder of the 120 or 600 h
test duration. Bubbling continued in 3 % HCl resulting in a high mass loss.
Carbon apparently changes the mechanism of dissolution from shallow pit
corrosion (Fig. 4.20a, b) at a low C content (GCN65) and high mass loss
(Fig. 4.19) to corrosion along crystal planes (Fig. 4.20c, d) at higher C contents
(GCN88 to GCN115) and lower mass loss (Fig. 4.19). These results are consistent
102 4 Properties

Fig. 4.19 Effect of carbon


on the removal w after
uninterrupted submersion
tests in aqueous solutions
lasting 120 or 600 h

with earlier work [15]: Corrosion is impeded by the most resistant crystal planes
(e.g. Fig. 4.20d) but increases, if this barrier is overcome. The resulting shallow pit
corrosion (e.g. Fig. 4.20b) is less dependent on crystal structure.

4.3.2 Current Density/Potential Tests

In submersion tests corrosion proceeds by chemical reactions at the resting


potential. If the potential is influenced from outside, we speak of electro-chemical
corrosion. In a respective test according to DIN 50918 and ASTM G5-94 the
potential between the specimen electrode and a reference electrode is raised and
the resulting current measured at room temperature. The reference calomel elec-
trode of potential UC = +244.3 mV is connected to the electrolyte via a Haber-
Luggin capillary. The dissolution of metal atoms yields an electric current which is
related to the test area and plotted as current density in dependence of the steadily
increasing potential, that is transformed to the potential UH = 0 mV of the stan-
dard hydrogen electrode SHE. Two hours after final grinding with 1000 mesh SiC
paper the specimens were inserted into the electrolyte which was then purged with
nitrogen for 30 min to expel remainders of oxygen. After a cathodic treatment at
-1 V (SHE) for 1 min, the resting potential UR was measured for 30 min. Starting
from 10 mV below UR the current density/potential curve was recorded by raising
the potential at a rate of 0.6 V/h. Tests were run at room temperature in 0.5 molar
H2SO4, corresponding to an aqueous solution of 4.74 mass % and in an aqueous
solution of 3 mass % NaCl. Typical curves are shown in Fig. 4.21 and key
potentials and current densities of the steels investigated are listed in Table A15.
4.3 Corrosion Resistance 103

Fig. 4.20 Surface of centrifugally cast steels after uninterrupted 120 h submersion in (a, c,
d) 1 % HCl, (b) 3 % HCl

Diluted electrolytes of H2SO4 are commonly used to study general corrosion of


stainless steels. The rising potential stands for an increase of the redox potential
encountered in practical corrosion systems. The higher the resting potential UR and
the shorter the distance to the passivating potential UP, the smaller is the range of
active corrosion. In some cases a prepassivation occured at UPP before the pas-
sivating potential UP at which the current density iP drops to the low level i0 of the
passive range. It is terminated at the break down potential UB which is, however,
interrupted by repassivation between URP and the final break down at UB0 .
To visualise the corrosion attack, specimens were polarised at specific poten-
tials. A typical result is depicted in Fig. 4.22. In the active range severe corrosion
along crystal planes and by shallow pits becomes visible. In the passive range the
grains appear but scratches from grinding are still noticeable. At the repassivating
potential only a few local pits have been formed, although the corresponding
current density iRP is high i.e. about a quarter of iP. This means that a major part of
the current is spent on building a layer of corrosion products, which was observed
earlier [16].
The chemical composition of HIS has little effect on the shape of the curves in
0.5 m H2SO4 (Table A15a). There is no influence of carbon on the resistance of
centrifugal castings. Their higher Cr content raises UR and lowers i0, though. The
Mo content has a similar effect on UR and distinctly lowers the passivating current.
As to the reference steels, CrMnN comes with a repassivation just like HIS, while
CrNi does not. The key values of the latter come quite close to some of the HIS as
e.g. GCN98.
104 4 Properties

Fig. 4.21 Characteristic


current density/potential
curves, a In 0.5 m H2SO4,
b In 3 % NaCl, type I with
passive range, type II
without. The indices stand for
A = activating, B = break
down, P = passivating,
PP = prepassivating,
R = resting,
RP = repassivating,
0 = passive

Sea water and salted roads give rise to pitting corrosion of stainless steels by
Cl- ions. A simple method to measure the liability of such a steel to this type of
localised corrosion is a current density-potential test in a 3 % NaCl electrolyte.
The attention is not so much on the passive range of low i0 but on the breakdown
potential. The higher UB, the better is the pitting resistance (Fig. 4.21b). As the
potential is raised the breakdown may not be abrupt and a criterion is necessary to
define UB. The results in Table A15b are based on a lasting increase of i0 by 10
lA/cm2. This sharp criterion works well with type I curves in Fig. 4.21 but may be
misleading in case of type II. This is best documented in linear plots of the curves
(Fig. 4.23). In these examples steel CN85 is of type II, starts to deviate early and
the respective UB is -0.19 V. However, the major breakdown occurs at
UB = 0.69 V. The evaluation of UB is further aggravated by repassivation of early
pits which leads to small intermittent peaks of current before final breakdown.
UB of CN96 (0.2 V) is slightly lower than that of CrNi (0.26 V) but very much
surpassed by UB of CN107 (0.91 V). Together with the higher Cr content the high
carbon content of GCN115 leads to UB = 1.28 V. This points to a beneficial
influence of carbon on the pitting resistance of stainless nitrogen steels. Steel
CrMnN is an example of the latter and in spite of its high Cr content does not
exceed UB = 0.74 V. The grade GCN65 is a nitrogen steel as well, but its low
breakdown potential of 0.18 V is most likely connected with remainders of
4.3 Corrosion Resistance 105

Fig. 4.22 Surface of CN859


after polarization in 0.5 m
H2SO4 at RT for 1 h against
SHE, a In the active range at
U = -0.256 V, b In the
passive range at
U = 1044 V, c At the
repassivating potential
U = 1294 V
106 4 Properties

Fig. 4.23 Current density i


in dependence of potential U
(SHE) at RT in 3 % NaCl

d-ferrite in the as-quenched microstructure. The addition of Mo comes with an


increase of UB from CN96 (0.2) to CN94Mo1 (0.44 V) but with a decrease of UB
from CN107 (0.91) to CN103Mo1 (0.88 V). Two mass % Cu clearly bring UB
down to the lowest value (-0.04 V) of all steels tested.

4.3.3 Tests on Intercrystalline Corrosion

The high interstitial content of the new steels is liable to form precipitates along
grain boundaries during cooling from the temperature range of homogeneous
austenite or during reheating [17]. This may cause a depletion of Cr adjacent to the
boundaries and entail intercrystalline corrosion [18]. The susceptibility to inter-
crystalline corrosion (IC) was investigated by Strauß tests according to
EN ISO 3651-2 method A. Specimens of Ø4 9 10 mm were gas quenched from
TSA, ground to strips of 4 9 1 9 10 mm, subjected to the test solution, bent by 90°
and inspected for cracks by standard 10 times magnification and by SEM. As no
cracks were detected, no IC had occurred. The small size of specimens was chosen
to minimise the case to core difference of cooling time. In Sect. 5.3.3 the cooling
time was prolonged step by step until IC was observed.

4.4 Magnetic Properties

The magnetic moments of individual iron atoms in HIS perform a random motion
between the possible orientations of their projections. If an external magnetic field
is applied, the moments tend to align in the field direction making a Larmor
precession. Temperature causes a chaotic flipping of the moments which prevents
their alignment. The combined action of these processes results in the paramag-
netic state of HIS. However, at technically possible field strengths, the system
4.4 Magnetic Properties 107

Table 4.2 Temperature dependence of the volume magnetic susceptibility v taken at


H = 0.33 T
susceptbility v
T MnC CN107 CN96 CN85
[K] (10-6) (10-4) (10-6) (10-4)
300 2.276 0.815 0.442 2.052
250 2.276 0.825 0.443 2.734
200 2.276 0.828 0.443 3.174
150 2.276 0.828 0.443 3.374
100 2.276 0.828 0.444 3.414
50 2.276 0.828 0.445 3.424
4.2 2.276 0.828 0.448 3.424

Table 4.3 Relative magnetic permeability lr measured by Förster Magnetoscope 1.067-103 with
a handheld sensor. In a first series the specimens were tested at room temperature after (1)
solution annealing, (2) subsequent cold drawing to a reduction of 16–37 % and (3) subsequent
deep freezing in liquid nitrogen for 15 min. In a second series solution annealed specimens were
tested at room temperature (4) on the deformed impact wear surface and (5) after subsequent deep
freezing in liquid nitrogen for 2 h
1 2 3 4 5
Steel Solution Cold drawn Cooled Impacted Cooled
annealed to -196 °C to -196 °C
%
CN85x 1.0010 – – 1.0011 1.0011 1.0011
CN96 1.0008 – – 1.0011 1.0010 1.0011
CN107 1.0007 – – 1.0011 1.0011 1.0011
CN96 – 16 1.0011 1.0011 – –
CN96 – 20 1.0008 1.0008 – –
CN96 – 29 1.0008 1.0008 – –
CN96 – 37 1.0009 1.0009 – –
CN94Mo1 1.0008 – – 1.0008 – –
CN103Mo1 1.0009 – – 1.0009 – –
CN96Cu2 1.0010 – – 1.0010 – –
GCN65 1.0150 – – 1.0150 – –
GCN88 1.0009 – – 1.0009 – –
GCN98 1.0009 – – 1.0009 – –
GCN115 1.0008 – – 1.0008 – –
GCN85 1.0014 – – 1.0014 – –
MnCr82 1.0011 – – 1.0011 1.0011 –
CrMnN 1.0010 – – 1.0012 1.0010 1.0012
CrNi 1.0900 – – – 1.1200 –
MnC 1.0010 – – – 1.0025 –

stays far from a total alignment of the moments and thus far from saturation. The
degree of alignment along the magnetic field, i.e. the magnetic moment M of a unit
volume, is linked to the applied field strength H via the volume susceptibility v, in
108 4 Properties

short M = vH. For convenience it is generally accepted to use Tesla (T) as unit
for M and H, so that v is dimensionless. As pointed out in Sect. 3.1.3.2, the
paramagnetism of HIS is of the Van Vleck type. A transformation to the anti-
ferromagnetic state was not observed down to a temperature of 4 K. As shown in
Table 4.2, the volume susceptibility is quite small and only slightly increases
during cooling.
In a technical sense this persistent paramagnetism translates to a nonmagnetic
behaviour of workpieces. They are not prone to guide magnetic fields or generate
heat in alternating electromagnetic fields. Applications are mentioned in Sect. 6.2.3.
Deviations from a nonmagnetic behaviour may occur by small fractions of a
ferromagnetic phase, e.g. remainders of d-ferrite or strain-induced a-martensite,
embedded in paramagnetic austenite. The relative permeability lr, i.e. the slope of
the virgin magnetization curve, was measured to make sure that no trace of fer-
romagnetic phase was present, neither after deep freezing nor after straining. The
results in Table 4.3 confirm that lr of HIS is quite low except for GCN65 which
contains remainders of d-ferrite. This applies also to CrNi which in addition picks
up some a-martensite by impact deformation. The very thin amorphous surface
layer on impacted HIS, although ferromagnetic (Sect. 3.2.4.1), is not revealed by
this test method.

References

1. Riedner S (2010) Höchstfeste nichtrostende austenitische CrMn-Stähle mit (C ? N), doctoral


thesis, Ruhr-University Bochum
2. Ludwigson DC (1971) Modified stress-strain relation for FCC metals and alloys, Metallurg.
Transactions A2:2825–2828
3. Ludwik P (1909) Elemente der Technologischen Mechanik. Springer, Berlin
4. Uggowitzer PJ, Speidel MO (1991) Ultrahigh-strength Cr-Mn-N steels. In: Stainless
Steels’91, Chiba conference, The Iron and Steel Institute of Japan, pp 762–770
5. Nikulin I, Kaibyshev R (2011) Deformation behavior and the Portevin-Le Chatelier effect in
a modified 18Cr-8Ni stainless steel, Material Sci. and Engin. A 528:1340–1347
6. Böhm H, Schirra M (1973) Einfluss der Kaltverformung auf das Zeitstand- und
Kriechverhalten einiger warmfester austenitischer Stähle. Arch. Eisenhüttenwes. 44:785–791
7. Larson FR, Miller J (1952) A time-temperature relationship for rupture and creep stresses,
Trans ASME 74:765–775
8. Berns H (1975) Das Zeitstandverhalten von Warmarbeitsstählen und seine Bedeutung für die
Auslegung von Blockaufnehmern und Druckgusskammern, habilitation thesis, Technical
University Berlin
9. Berns H, Gavrilujk VG, Nabiran N, Petrov YuN, Riedner S, Trophimova LN (2010) Fatigue
and structural changes of high interstitial stainless austenitic steels, steel research int
81(4):299–307
10. Dengel D, Dahl W (ed) (1978) Verhalten von Stahl bei schwingender Beanspruchung, Verlag
Stahleisen, Düsseldorf, pp 23–46
11. Berns H, Siebert S (1996) High Nitrogen austenitic cases in stainless steels. ISIJ Int 36:927–
931
12. Kruger J (1988) Passivity of metals—a materials science perspective. Int Mat Rev 33:113–
130
References 109

13. Berns H, Riedner S, Hussong B (2010) Influence of molybdenum and copper on the corrosion
resistance of high strength austenitic steels. Mater Sci Forum 638–642:2979–2985
14. Berns H, Hussong B, Riedner S, Wischnowski F (2010) Effect of carbon on stainless
austenitic FeCrMnN steel castings, steel res int 81:245–251
15. Dobbelaar ALJ, Herman CME, Dewit HWJ (1992) The influence of the microstructure on the
corrosion behaviour of Fe-25Cr. Corros Sci 33:779–790
16. Schwenk W (1963) Beobachtungen über die Korrosion nichtrostender Stähle in
Schwefelsäure unter potentiostatischen Bedingungen, Werkstoffe u. Korrosion 14:646–654
17. Mujica Roncery L, Weber S, Theisen W (2011) Nucleation and precipitation kinetics of
M23C6 and M2N in an Fe-Mn-Cr-C-N austenitic matrix and their relationship with the
sensitization phenomenon. Acta Mater 59:6275–6286
18. Dayal RK, Parvathavarthini N, Raj B (2005) Influence of metallurgical variables on
sensitisation kinetics in austenitic stainless steels. Int Mat Rev 50:129–155
Chapter 5
Manufacture

We discern between hot manufacturing processes as e.g. melting, casting, hot


working, heat treatment and welding or cold processes as cold forming and
machining, although some heat may be generated. The hot steps are accompanied
by phase transformations which come the closer to thermodynamic equilibrium the
higher the temperature and the longer the time of treatment are. The sequence of
slow solidification, hot working and solution annealing therefore approximates to
phase diagrams. On the basis of the actual chemical composition of the new HIS a
set of isoplethal diagrams is calculated to explore (a) the variation of C ? N,
(b) the variation of C/N and (c) the mole fraction in dependence of temperature
(Appendix B, Figs. B1–B10). From these, key temperatures are derived. These are
the liquidus and solidus temperatures TL and TS, which mark the range of solid-
ification, and the temperature TP, which stands for the begin of precipitation in
austenite and thus about ends the range of hot working and solution annealing.
These temperatures are influenced by the carbon content (Fig. 5.1). As it increases
TS is lowered more than TL which widens the range of solidification enhancing
segregation and, together with a rise of TP, narrows the range of homogeneous
austenite. This is the more so, if 2 mass % copper are added while the effect of
1 mass % Mo or the higher Cr content in the castings is less significant. At the
example of CN85, CN96 and CN107 (Figs. B1b–B3b) it becomes clear that only
CN85 is close to (C/N)op defined in Fig. 2.3. At the higher carbon contents TP is
raised above Top which is marked by (+) in Fig. B1b for comparison. The higher
TP, the higher is the rate of diffusion and therefore the tendency to precipitation
during quenching of heavy cross-sections.
The cold processes are depending on work hardening which increases as the
deformation proceeds (Fig. 4.1b). The high work hardening exponents n1 in
Table A2 point to high forces required for metal forming or cutting. The respective
microstructural changes are described in Sect. 3.2.1.

H. Berns et al., High Interstitial Stainless Austenitic Steels, 111


Engineering Materials, DOI: 10.1007/978-3-642-33701-7_5,
Ó Springer-Verlag Berlin Heidelberg 2013
112 5 Manufacture

Fig. 5.1 Liquidus and


solidus temperature TL, TS
and begin of precipitation at
TP in dependence of the
carbon content derived from
equilibrium phase diagrams
of HIS in Figs. B1 to B10,
encircled = HIS with
2 mass % Cu (Table A18)

5.1 Melting and Casting

All new HIS were molten in induction furnaces the capacity of which ranged from
4 kg to 5 Mg. The final temperature in the furnace has to be higher, if a ladle is
used. The melt from small furnaces is usually poured directly into the mould. So
there are no general recommendations for the tapping or pouring temperature of
new HIS. Nitrogen was added via nitrided ferrochromium and melting as well as
pouring was done under normal pressure of air.

5.1.1 Ingots

Melts of 5 Mg were produced of CN96 and CN107 (Table 2.1) at Friedrich


Lohmann GmbH in D-58454 Witten and teemed to three square ingot moulds each
of tapered shape that was reduced to cylinders of Ø310 mm by forging. These
electrodes were electro-slag-remelted at Energietechnik in D-45143 Essen, to give
ESR ingots of Ø430 mm. This remelting was done in a pressurised facility (PESR)
because of availability, which was run at a pressure of 5 bar argon.
A 2.5 Mg melt of CN85 was teemed at the Deutsche Nickel in D-58239
Schwerte to a conical mould of 350–315 mm diameter and 3900 mm length. The
ingot was processed in a standard ESR unit under normal pressure of air to give a
remelted ingot of Ø510 mm.
The grades CN94Mo1, CN103Mo1 and CN96Cu2 were molten in a lab furnace
of 4 kg capacity and poured into a Y-shaped steel mould. After removal of the
feeder small slabs of 150 by 70 by 30 mm were ready for hot rolling.
5.1 Melting and Casting 113

Fig. 5.2 Scheil simulation of


GCN88 compared to
equilibrium solidification,
Scheil range of phases:
1 = L, 2 = L ? F,
3 = L ? F ? A,
4 = L ? A,
5 = L ? A ? M23C6

5.1.2 Centrifugal Castings

The alloys GCN65 to GCN115 (Table 2.1) were molten in a 450 kg furnace at
Klaus Kuhn Foundry in D-42477 Radevormwald. Starting from GCN65 portions
of the melt were successively enriched with carbon in the laddle to give a series of
four steels. These were solidified by horizontal centrifugal casting. The mould of
100 mm inner diameter consisted of low carbon steel S355 and was cooled from
outside. Tubes of Ø100 by Ø40 mm in size were produced at a rotational speed of
200 rpm.
The cooled metallic mould raises the rate of solidification and increases the
deviation from equilibrium. This is expressed by a Scheil simulation (Fig. 5.2)
provided by Thermo-Calc which resembles a non-equilibrium cooling that allows
of infinitely rapid diffusion in the liquid but none in the solid state except for the
light elements carbon and nitrogen. Compared to equilibrium the solidus tem-
perature is lowered by almost 100 °C. Rapid solidification is expressed in the
macrostructure by the radial growth of elongated primary grains (Fig. 5.3a).
Remainders of d-ferrite are encountered in GCN65, while some decorated grain
boundaries are visible in GCN115 even after quenching (Fig. 5.3b).

5.1.3 Sand Castings

Sand moulds were prepared by the KSB foundry in D-91257 Pegnitz and cast from
an 80 kg furnace to give castings as depicted in Fig. 5.4 for further investigation.
After slow cooling in the mould the grain boundaries are decorated with pseudo
pearlite (Sect. 5.3.2). These M(C,N) precipitates extend close to the surface,
114 5 Manufacture

Fig. 5.3 Microstructure after centrifugal casting and solution annealing (LOM). a Section
through a cast tube of GCN88. b Decorated grain boundary of GCN115

Fig. 5.4 Drawing (mm) of


sand castings made of
GCN85 with positions of
specimens (grey) for
mechanical testing. Strips
s = 26 mm thick were cut
from castings for further hot
rolling

indicating that there is little decarburisation or loss of nitrogen after casting


(Fig. 5.5). The surface itself is smooth, reflecting just the roughness of the sand
mould. After solution annealing and quenching EDX line scans revealed micro-
segregations of Cr and Mn. The maximum content of these elements divided by
the minimum content stands for the degree of segregation S and amounts to
SCr = 1.27 and SMn = 1.34. The average spacing is about 50 lm [1].

5.1.4 Refractories

It was shown that melting, remelting and different casting procedures of the new
HIS are industrially feasible. A nitrogen content of about 0.6 mass % was reached
without pressure- or powder metallurgy which, together with carbon, is a sound
basis for interstitial strengthening. To reach this high nitrogen content, nickel had to
be replaced by manganese which tends to interact with refractories, i.e. furnace
lining, slag, sand mould and inclusions. The monolithic lining of induction furnaces
used to be made of rammed fireclay-bonded sand to give mullite (3Al2O32SiO2)
which is accompanied by expansion that prevents a strike of the melt to the cooling
coil [2]. This acidic lining matches with an acidic slag. Early work of Körber [3, 4]
suggested that Mn and—to a lesser extent—Cr are capable to reduce SiO2 and the
5.1 Melting and Casting 115

Fig. 5.5 Section normal to the as-cast surface of GCN85 (SEM). a Decorated grain boundaries.
b Pseudo-pearlite growing from grainboundary

more so the higher the temperature and the lower the carbon content are. This
implies that Mn is transferred from the steel to the slag and SiO2 is lost from the
lining which is accompanied by a rising Si content in the melt. This corrosive wear
of the quartz constituent in the lining is most pronounced at the slag line. A basic
MgO lining and a respective slag would be resistant to the attack of Mn but without
expansion of the rammed lining. A compromise is seen in an MgAl2O4 spinell with
excess Al2O3. This refractory material comes with sufficient expansion and is called
neutral between acidic and basic [5]. Melting of larger HIS volumes in an electric
arc furnace may be handled in a hearth consisting of MgO bricks, which stay hot.
In the foundry the interaction of high manganese melts with sand moulds is
inhibited by a proper facing, as e.g. zirconia [6]. This procedure is well known
from Hadfield manganese steel.

5.2 Hot Working

The Ø430 mm ESR ingots of CN96 and CN107 were forged at Friedrich Lohmann
to bars of Ø120 mm. The Ø510 mm ESR ingot of CN85 was rolled at Deutsche
Edelstahlwerke in D-58452 Witten to 180 mm square and further reduced to bars
of Ø65 mm by a radial forging machine.
Sections of CN96, Ø290 9 530 mm in size, were upset forged, punched and
ring-rolled on a tyre rolling mill at Bochumer Verein in D-44793 Bochum to
obtain tyres with rim for rail vehicles. The laboratory slabs were hot rolled to half
of their initial thickness at the Institut für angewandte Materialtechnik of the
University in D-47057 Duisburg.
Last not least a strip was taken from a GCN85 sand casting according to
Fig. 5.4 and rolled from 26 mm initial thickness to 11 mm final thickness at the
Max Planck Institute in D-40237 Düsseldorf. According to Table 2.1 this material
is called CN85x and allows to study the effect of hot working on the mechanical
properties by comparing CN85x in Table A1 with GCN85 in Table A4. The
recrystallized fine grained microstructure of the former (Table A5) considerably
116 5 Manufacture

raises strength and ductility. However, the higher strength entails a small decrease
of the notch impact toughness at room temperature and a slight increase of the
DBTT (Table A11).
Comparing the tensile properties of CN85x and CN85 (Table A1) does not
show much effect of previous electro-slag-remelting the latter. As the testing
temperature is decreased to -100 °C the ESR grade provides a higher proof
strength and respectively less elongation A but more necking, i.e. a higher
reduction Z (Table A6). After ESR the notch impact energy at RT is definitely
raised by about 20 %, while the DBTT is slightly lowered (Table A11). The higher
upper-shelf energy of ESR is consistent with a better cleanness. Smaller nonme-
tallic inclusions induce smaller microvoids during straining and fracture by
macrovoid coalescence requires more deformation. This is more pronounced for
the multiaxial stress state of the notch impact test than for the tensile test.
Hot working with different types of forging and rolling steps covered a wide
range of cross-sections and were done without any unexpected incident or scrap.

5.3 Heat Treatment

Solution annealing followed by quenching is the main heat treatment of HIS.


Aging is sometimes applied to cold worked HIS.

5.3.1 Solution Annealing

Precipitates are observed in the microstructure after slow cooling from casting or
hot working (Fig. 5.5). The purpose of solution annealing is to dissolve these
precipitates and achieve a structure of homogeneous austenite. To this avail the
solution anneal temperature TSA has to be above the equilibrium temperature of
beginning precipitation TP which is shown in Fig. 5.1. An increase of TSA
abbreviates the time for dissolution and allows to dissolve different chemical
compositions locally provoked by microsegregation (see Sect. 5.1.3), which are
not reflected by Thermo-Calc. There is, however, a limit to TSA because of grain
growth which impairs the strength of hot worked steel. The coarser grain size of
castings (Table A5) is less affected by TSA which in general is raised above that of
hot worked steel, because a casting is not exposed to the effect of diffusion
annealing during soaking at the forging temperature. Remainders of undissolved
precipitates may be encountered in castings (Fig. 5.3b) but rarely in forgings.
The steels investigated were solution annealed at different temperatures,
quenched in water and inspected by LOM and SEM to find TSA experimentally.
The results are compiled in Table A16. The equilibrium pressure of nitrogen gas
pN2, calculated by Thermo-Calc for TSA of each steel, is required to keep the
nitrogen content in the surface at the given level. It depends on the chemical
5.3 Heat Treatment 117

Fig. 5.6 Subsurface concentration of elements in CN96 after solution annealing at 1100 °C for
3 h in air. The results stem from glow discharge spectroscopy

composition and the temperature, but stays in all cases below the partial pressure
of nitrogen in air. Therefore some up-take of nitrogen is to be expected rather than
a loss. This is indeed supported by measurements depicted in Fig. 5.6. After
prolonged annealing of CN96 in air, nitrogen is increased towards the surface,
while some carbon is lost by oxidation. In total, the interstitial content of the
subsurface zone is changed little, which is underlined by Fig. 5.7a, which shows
this zone after precipitation annealing at 900 °C to generate pseudo-pearlite. It is
evident that a thin layer of only 0.2 mm in depth is enriched with pearlite,
respectively C ? N.
The exchange of elements between the steel surface and air is inhibited by a
thin coat of oxides which are rich in chromium. This is not the case, if solution
annealing is carried out in an N2 atmosphere. As a result the subsurface content of
nitrogen is raised considerably (Fig. 5.7b) as indicated by the higher content of
precipitates near the surface which is topped by a thin nitride layer. In contrast
solution annealing in vacuum entails a severe loss of N and Mn (Fig. 5.8). Again
pseudo-pearlite is used as a marker. Its content is reduced from core to case
followed by layers of (i) austenite plus grainboundary precipitates, (ii) ferrite plus
carbides and finally (iii) plain ferrite at the surface, the Mn content of which was
down to 0.5 mass % as measured by EDX [7].
118 5 Manufacture

Fig. 5.7 Microstructure (LOM) in a section normal to the surface of CN96 after annealing at
1100 °C for 3 h. a In air of about 1 bar pressure. b In nitrogen of 1 bar pressure, followed by
quenching in water and subsequent annealing at 900 °C for 2 h in air to generate pseudo-pearlite

5.3.2 Interrupted Quenching

The constitutional diagrams in Figs. B1 to B10 give evidence of phases as M2N


nitrides, M23C6 carbides and r-phase below TP. It is the purpose of quenching
from TSA to subdue their precipitation. In this context it is important to know the
temperature range in which the rate of precipitation is highest. Therefore speci-
mens were solution annealed, cooled to an isothermal temperature between 1000
and 700 °C, held for 1 h and quenched in water. The resulting content of pseudo-
pearlite was evaluated by microscopy and plotted in Fig. 5.9. It increases with the
C ? N content which raises the temperature of maximum precipitation. The lower
concentration of grain boundaries in the coarse grained as-cast steel GCN85
reduces the volume of precipitates as compared to hot worked CN85x. The
addition of 1 mass % Mo tends to reduce the volume of pseudo-pearlite while
2 mass % Cu have little effect (Fig. B11).
Pseudo-pearlite is known from high nitrogen steels where it consists of M2N
and austenite lamellae. A specimen of CN96 was intensely etched to locally
remove the austenite and analyse a protruding lamella by EDX (Fig. 5.10). About
half of the atoms consist of C and N and the other half of metal ones. Notwith-
standing the problem of measuring light elements quantitatively the MX lamella is
not an equilibrium phase (Fig. B2). Even after annealing at temperatures between
950 and 1020 °C the specimens did not fully arrive at equilibrium, but grain
boundary precipitates pointed to carbides while those within grains resembled
nitrides. This coincides with results found after creep at 650 °C (Fig. 4.8c) and
with the general experience that carbon has a strong affinity to grain boundaries
while nitrogen has not.
The method of interrupting the quench by isothermal holding was also applied
to notch impact specimens to measure the effect of precipitation on toughness KV
and sensitisation to intercrystalline corrosion IC. Only after a short holding of
\1 min the specimens stay free of IC. Within 10 min KV comes down to a fully
brittle level, especially if the C ? N content is high (Fig. 5.11).
5.3 Heat Treatment 119

Fig. 5.8 Microstructure in a


section normal to the surface
of CN96 after annealing at
1100 °C for 3 h in vacuum of
4  10-5 bar, quenching in
pressurized N2 gas and
subsequent annealing at
900 °C for 2 h in air to
generate pseudo-pearlite.
a Overview of the
sub-surface zone (LOM).
b Near surface zone enlarged
(SEM). c EBSD analysis of
the near surface zone
120 5 Manufacture

Fig. 5.9 Content of pseudo-


pearlite after solution
annealing and isothermal
holding at temperature Tis for
1 h and subsequent
quenching in water

5.3.3 Continuous Quenching

Interrupted quenching has shown that the climax of precipitation occurs between
1000 and 700 °C. The t10/7 cooling time is therefore a suitable measure to search
for the onset of precipitation at the critical cooling time tc10/7 (Fig. 5.12). Pre-
cipitation begins at grain boundaries and may be detected by microscopy or notch
impact testing (Fig. 5.11). However, a most sensitive method is a Strauß test for
intercrystalline corrosion according to EN ISO 3651-2 method A. Specimens of
Ø4 9 10 mm were continuously gas quenched from TSA at different t10/7, ground
to plates of 4 9 1 9 10 mm, subjected to the test solution, bent by 90° and
inspected for cracks at 10 times magnification. After increasing t10/7 step by step
tc10/7 is assigned to the transition of uncracked to cracked specimens. The results
are presented in Fig. 5.13. The critical cooling time is lowered as the C ? N
content or the C/N ratio is raised. The latter stems from an increase of TP above
Top which is shown in Fig. B1b. The higher TP, the more the range of precipitation
during quenching is shifted to faster diffusion which is indirectly reflected in
Fig. 5.9. Copper appears to be an exception, because it raises TP (Fig. B6) but also
tc10/7 (Fig. 5.13) [8]. The grainboundary segregation of carbon is inhibited by
copper just as by nickel which may explain this result.
The cooling time of a work piece depends on the quenchant and on the size and
shape of the cross-section in which it grows from case to core. In the order of
quenching in water, oil or air the notch impact toughness was reduced consider-
ably and the more so the higher the C ? N content. The sensitisation to inter-
crystalline corrosion increased in the same order (Table 5.1).
These results call for quenching in water or even intensive quenching in agi-
tated water to reduce the steam envelop. Published data on natural cooling in water
are a guideline for simple shapes as round or rectangular bars of different size.
Numerical simulations are another way of finding t10/7 from case to core of a
workpiece. The higher the C ? N content or the C/N ratio, the smaller is tc10/7 and
5.3 Heat Treatment 121

Fig. 5.10 Microstructure of


steel CN96 after solution
annealing and isothermal
holding at 900 °C for 10 min.
a Pseudo-pearlite after deep
etching in V2A at 60 °C for
25 min (SEM). b EDX
analysis of protruding
lamella, see frame in a

also the cross section that stays free of precipitation to the core. However, cooling
is faster at the surface which tends to prevent IC. Some loss of toughness may be
acceptable in the core. In other applications IC is not to be expected. Therefore the
admissible cross-section is related to the in-service conditions. During manufac-
ture a sensitisation to pickling agents has to be avoided.
122 5 Manufacture

Fig. 5.11 Effect of solution


annealing and subsequent
holding in a salt bath at
900 °C for tis B 600 s on the
notch impact energy KV and
the sensitisation to
intercrystalline corrosion IC

Fig. 5.12 Schematic


representation of the t10/7
cooling time and the critical
cooling time tc10/7 at which
precipitation begins,
TSA = solution anneal
temperature,
TP = equilibrium
temperature of beginning
precipitation, G = grain
boundary decoration,
D = discontinuous
precipitation of pseudo-
pearlite

Fig. 5.13 Critical cooling


time tc10/7 of new HIS
5.3 Heat Treatment 123

Table 5.1 Notch impact toughness KV [J] and intercrystalline corrosion IC (+) in dependence of
the t10/7 cooling time [s] during quenching from TSA (Table A16) in water, oil or air, respectively,
measured in the core of 10 mm square test bars
steel water oil air
KV t10/7 IC KV t10/7 IC KV t10/7 IC
CN 96 367 1.4 – 278 5.9 ; 204 47.1 +
CN107 348 2.2 – 193 6.9 + 12 52.4 +

Fig. 5.14 Effect of cold working and aging on hardness HV30 of CN96 measured at RT. a After
44 % of cold upsetting. b After reduction of cross-section by cold-drawing

5.3.4 Aging

Aging is caused by a rearrangement of interstitial atoms in cold worked austenite


up to prestages of precipitation. To induce a deformed structure the height of a
solution annealed and quenched cylinder Ø13 mm was cold compressed from 20
to 11.2 mm which corresponds to an upset strain of 44 %. Immediately after cold
working the specimen was consecutively aged at temperatures up to 550 °C. The
first peak of hardness at 400 °C in Fig. 5.14a is most likely caused by aging, the
second at C550 °C by precipitation (see Sect. 4.1.3).

5.4 Cold Drawing

The intensive work hardening of HIS may be used to raise the proof strength. Cold
drawing of rod or wire is one way to go. Solution annealed rods of Ø18 mm were
processed on a Schumag 2B drawing bench of 8 MN force at the Deutsche
Edelstahlwerke in D-58089 Hagen. In several stages the reduction of cross-section
124 5 Manufacture

Fig. 5.15 Effect of reduction


by cold drawing of rods on
the engineering stress–strain-
curves of CN96, the dashed
line represents the true stress

was raised to 37 % which corresponds to the elongation obtained. A plot of


respective engineering stress–strain-curves reveals that the yield point is raised by
proceeding cold reduction and that it follows the true stress curve (Fig. 5.15).
After 37 % reduction the 0.2 % proof strength is about doubled and the ratio Rp0.2/
Rm is raised to 0.84 (Table A17). Fracture elongation and energy as well as the
workhardening exponent n1 are reduced, because part of the plasticity has been
spent by cold drawing. Remarkable is Z = 58 % after 37 % reduction by drawing
starting from Au = 12 % which gives proof of the high workhardening capacity
and ductility of HIS during necking.
Shortly after drawing the hardness was measured and again after 3.5 years. Age
hardening had occurred at room temperature as depicted in Fig. 5.14b.

5.5 Welding

Tungsten inert gas (TIG) welding is widely applied to stainless steels. The electric
arc creates a hot spot on the steel surface that entails a weld pool in which the heat
is distributed by convection. As the heat source moves a solidifying mushy zone is
travelling behind and a heat affected zone (HAZ) is induced in the base metal. A
summary of TIG welding high nitrogen steels is given in [9] and also applies to
high interstitial steels. The main problems encountered are (i) a loss of N and Mn
from the weld pool, (ii) the formation of nitrogen bubbles which do not fully
escape from the bath but are caught in the mushy zone forming pores, (iii) pre-
cipitation in the HAZ.
Admixing a few percent of nitrogen to the argon shielding gas is a means of
counteracting a loss of nitrogen and the formation of bubbles, although the cor-
rosion of the tungsten electrode may increase. Using a filler metal of higher N
solubility is another way of keeping nitrogen in the weld pool. The evaporation of
manganese is impeded by the pressure of the shielding gas and a good coveridge of
5.5 Welding 125

Fig. 5.16 Weldment of MnCr77 sheet using a 3 kW continuous Nd:YAG laser a normal section
through the seam (LOM). b Center of weld pool (LOM), (courtesy of L. Mujica Roncery)

the pool. This is best achieved if the bead is kept narrow. Mixtures of argon and
helium raise the arc temperature to more than 10000 °C and are not recommended.
It is essential to retain the content of N and Mn not only in respect to the strength
level but also to subdue d-ferrite during solidification which would enhance
degassing of N2.
Experience with TIG welding of austenitic HNS demonstrates that flawless
weldments without pores or hot tearing are feasible but that precipitates in the heat
affected zone cause intercrystalline corrosion and some loss of ductility. This was
corroborated recently in an investigation of welding stainless TWIP steels of high
interstitial content as e.g. MnCr77 in Table 2.1. Solution annealing after welding
will restore a fully austenitic structure. Continuous laser welding of sheet 1.2 mm
thick was carried out in [10] without intercrystalline corrosion in the HAZ. The
weld pool was about as wide as the sheet thick, i.e. the pool volume per unit length
was very small and therefore the self-quenching rate high enough to prevent
precipitation (Fig. 5.16). The hardness of the bead is slightly raised.
In contrast to fusion welding experience with pressure welding of HNS suggests
that electric resistance heating or friction goes without a loss of elements and
appears to be applicable to HIS if general rules for high alloy steels are obeyed.
Depending on the local cooling rate, subsequent solution annealing may be
necessary.

5.6 Machining

A high yield strength, intensive cold work hardening, and much ductility entail a
strong resistance of the new HIS to metal cutting. This resistance is related to the
specific fracture energy Ws of HIS (Tables A1, A3, A4) which is about 50 %
higher than that of standard austenitic steel CrNi and 2 to 7 times as high as that of
ferritic-pearlitic or quenched and tempered steels [11]. The chip formation on HIS
is therefore accompanied by high forces and by heat. The latter is not readily
dissipated because of a low thermal conductivity. A large rake angle is a way to
126 5 Manufacture

reduce the cutting force acting on the tool. This implies a sharper cutting edge
because rake angle, wedge angle of the tool tip and clearance angle add up to 90°.
The sharper the cutting edge the more fragile it is and the less brittle the tool
material should be. This is the reason why mostly WC–Co sintered hardmetals are
used which are abbreviated to tungsten carbide or WC tools. Because of their
limited temperature resistance the cutting speed v is kept at a low level. In rough
turning using M15 to M30 tool inserts with a chip breaker, v may be down to
25 m/min and in fine turning with M10 to M25 between 40 and 50 m/min. The
feed (depth of cut) should be high enough to undercut most of the cold worked
surface zone formed during the previous revolution but not overload the tool edge.
In some applications a feed of 0.4 mm/revolution was used. In general the
machine-tools should be quite rigid to keep vibrations low which would add to
workhardening.
From Hadfield steel we know that milling is possible but that drilling of deep
holes may pose a problem. There is no experience with the new HIS, but a similar
behaviour is expected. Hydro-jet-cutting with abrasives was applied to rough cuts
and electro-spark-erosion to fine machining operations and slender holes.

References

1. Berns H, Nabiran N, Mujica L (2012) High interstitial stainless austenitic steel castings, steel
research int. doi:10.1002/srin.201100332
2. Chesters JH (1957) Steelplant refractories. The United Steel Comp. Ltd., Sheffield
3. Körber F (1937) Einfluss der Beimengung auf die Reaktionen zwischen Eisenschmelzen,
Eisen Mangan-Silikaten und fester Kieselsäure. Stahl u. Eisen 57(48):1349–1355
4. Körber F (1936) Zur Metallurgie der Eisenbegleiter. Stahl Eisen 56(3):77–104
5. Schacht CA (ed) (2004) Refractories handbook. CRC Press, Boca Raton
6. Rudolph S (1994) Betrachtungen zum Aufbau von Form- und Kernschlichten unter
besonderer Berücksichtigung ihrer feuerfesten Bestandteile. Gießerei-Praxis 8:165–178
7. Riedner S, Berns H (2008) Wärmebehandlung hochfester, nichtrostender Austenite. HTM
63(2):84–94
8. Berns H, Riedner S (2008) Zusammenhang zwischen Konstitution und Wärmebehandelbarkeit
hochfester austenitischer Stähle. HTM 63:337–341
9. Gavriljuk VG, Berns H (1999) High Nitrogen Steel. Springer, Berlin
10. Mujica Roncery L (2010) Development of high-strength corrosion-resistant austenitic TWIP
Steels with C ? N, doctoral thesis, Ruhr University Bochum, Bochum
11. Berns H, Gavriljuk VG (2007) Steel of highest fracture energy, Trans Tech Publications. Key
Eng Mater 345–346:421–424
Chapter 6
Assessment

The aim of this concluding Chapter is to summerise and assess the contents of
previous chapters. On the scientific side the structure/property relation is of key
interest. The technical side is mainly concerned with manufacture and application
based on the achieved properties. The new high interstitial steels (HIS) are
assessed by comparing them with known austenitic steels. Their acronyms and
chemical compositions are given in Table 2.1.

6.1 From Structure to Properties

The multiscale approach of investigating the new HIS, depicted in Fig. 1.1, is an
attempt to relate macroscopic properties to microstructural features. Starting from
the scale of electron structure it was demonstrated that joint alloying with carbon
and nitrogen increases the concentration of free electrons in the CrMn austenite.
Respective measurements of conduction electron spin resonance (CESR) were
corroborated by ab initio calculations (Sect. 3.1.1). Free electrons enhance the
metallic character of interatomic bonds and thus the ductility. At the same time
they improve the homogeneity of the atomic distribution, i.e. reduce clustering of
alloying atoms and support their short-range ordering. This more even distribution
of alloying atoms in steels with C ? N was recorded by Mössbauer spectroscopy
(Sect. 3.1.2). Three methods were employed to measure the degree of chemical
homogeneity: (i) By CESR the size of clusters in HIS was estimated at 3 nm and
their concentration at 1.55  1019 cm-3 (Sect. 3.1.3.1). (ii) Magnetic measure-
ments revealed a cluster size of about 5 nm for CN107 and 3.5 nm for CN85,
while MnC alloyed only with carbon showed a cluster size of about 8 nm. (iii)
Measurements of the stacking fault energy (SFE) led to two maxima of dislocation
splitting which hints to short-range decomposition. The difference DSFE between
the two maxima decreased from CN107 to CN85. This points to less clustering,

H. Berns et al., High Interstitial Stainless Austenitic Steels, 127


Engineering Materials, DOI: 10.1007/978-3-642-33701-7_6,
Ó Springer-Verlag Berlin Heidelberg 2013
128 6 Assessment

which was highest for MnC (Sects. 3.1.3.3 and 3.1.3.4). In all, the SFE of HIS is
inversely proportional to the concentration of free electrons.
Originally the joint addition of C ? N was meant to circumvent costly pressure
metallurgy: A part of N in high nitrogen steels (HNS) was replaced by C to give
new HIS molten at normal pressure of air. Now it is evident that C ? N not only
ease production but exert a beneficial influence on the electron structure of HIS.
Their ductility is enhanced and their chemical homogeneity is improved on a
nanoscale, i.e. on a length scale more than four orders of magnitude below the
scale of microsegregations (Sect. 5.1.3). As a result of the more even atomic
distribution the stability of austenite to phase transformation is raised. The lower
the alloy concentration in the clusters the less they are prone to initiate precipitates
and the less the areas in between are liable to provoke bcc transformation. This is
reflected by thermodynamic equilibrium simulations which see a wide phase field
of homogenous austenite in the Fe-18Cr-18Mn-C–N system but not in the Fe-
18Cr-18Mn-C system (Fig. 2.1c, a). In fact an optimal C/N ratio exists at which
the austenite is most stable, i.e. extends to the lowest temperature (Fig. 2.3).

6.1.1 Mechanical Properties

A key property of the new HIS is the specific fracture energy Ws which corre-
sponds to the area below an engineering tensile stress/elongation curve. It profits
from a high yield point, from intense cold work hardening and from ductility. Ws
of CN96 is 100 % higher than that of the Hadfield carbon grade MnC, 60 % higher
than that of standard low interstitial steel CrNi and 20 % above that of the nitrogen
grade CrMnN (Table A1), which shows a higher DBTT in notch impact testing,
though (Table A11). CN96 is topped only by the high manganese TWIP steel
MnCr77 which is, however, less corrosion resistant.
The individual contributions of the proof strength Rp0.2, the work hardening
exponent n1 (Eq. 4.1) and the elongation at fracture are compared in Fig. 4.2 for
HIS (shaded area) and reference steels. The high interstitial content of HIS leads to
a proof strength which is almost three times that of low interstitial CrNi. In spite of
a higher interstitial carbon content MnC does not by far live up to the proof
strength of HIS. This is explained by a significantly stronger affinity of nitrogen
atoms to dislocations (Sect. 3.2.1.1).
Starting from a high yield point plastic straining of HIS is accompanied by a
succession of structural changes: planar slip followed by twinning, occasional e-
martensite and a high density of dislocations (Sect. 3.2.1.3). Their joint
strengthening effect delays necking and raises the uniform elongation of CN96 to
61 % compared to 45 % for MnC and 44 % for CrMnN. This evidences the
advantage of alloying with C ? N compared to C or N. Reference steels of higher
uniform or fracture elongation are either of low interstitial CrNi type or contain
less chromium which is bound to enhance the concentration of free electrons
(Fig. 3.6) but reduces the corrosion resistance.
6.1 From Structure to Properties 129

Effect of temperature. The fracture energy Ws is raised as the tensile test tem-
perature is lowered to -100 °C (Table A6). This depends mainly on an increase of
the yield strength, because the work hardening exponent n1 is hardly changed
(Table A2) and the fracture elongation is moderately reduced. The higher yield
strength is related to a decrease of SFE at subzero temperatures, which is caused by
an increase in the concentration of free electrons (Sect. 3.2.2). Between -100 and
-196 °C a severe embrittlement is observed (Fig. 4.4) which is known from high
nitrogen but not from high carbon austenite. In notch impact tests a DBTT in the
range of -100 °C is found for HIS based on a still generous toughness level of
KV = 100 J (Table A11). The respective DBTT of the high nitrogen steel CrMnN
is shifted to -21 °C, while the high carbon grade MnC starts from a lower energy at
RT and reaches 100 J at -59 °C but without a distinct ductile- to-brittle transition.
Compared to alloying with C the mobility of dislocations is enhanced by N or
C ? N because of a higher concentration of free electrons. This improves the
toughness at RT. As the testing temperature is lowered the number of glissile
dislocations in a pile-up is higher for N and C ? N steels which tends to open
microcracks if slip is blocked, e.g. by Lomer-Cotrell barriers. This ‘‘ductile’’ crack
initiation resembles the onset of embrittlement (Sect. 3.2.3.1).
At elevated temperatures serrations are observed in engineering stress/strain-
curves of HIS at 300 °C and again at 550–650 °C which are caused by an inter-
action of dislocations with interstitial or substitutional solute atoms, respectively
[1]. In the upper temperature range a sensitization to intercrystalline corrosion
points to a beginning precipitation. Actually the new HIS are designed to develop
optimal properties in the as-quenched state of homogeneous austenite but some
precipitation assists the short-time creep resistance at temperatures below 700 °C.
The creep elongation is reduced, though, especially by grain boundary decoration
(Sect. 4.1.4).
Effect of strain rate. Increasing the strain rate of HIS MnCr82 at RT raised the
proof strength without affecting the reduction of area [2]. At -196 °C an HNS
showed a mixture of brittle and dimple fracture at low strain rate while fully
dimple fracture prevailed at high strain rate [3]. This unusual behaviour was
explained by local heating caused by a localization of slip (Sect. 3.2.3.1). If it
comes to high velocity ballistic impact at RT the phenomenon of localised slip
leads to adiabatic shear bands and a perforation of CN96 targets. The high fracture
energy Ws measured by a low strain rate is not available at ballistic rates (Sect.
3.2.3.2).
Effect of cyclic loading. In push/pull-tests at RT the microstructural changes
revealed by TEM are similar to those after uniform straining except for additional
nitride precipitates (Sect. 3.2.4.2). In rotating bending short cracks start from slip
lines at the surface and follow crystallographic planes until, after two or three
grains, they extend normal to the bending stress up to fracture. The fatigue limit at
N = 107 cycles is not a true one but bound to come down, if cycling proceeds [4].
It is raised by previous cold working and in either case is about 50 % of the proof
strength. This is a key advantage of HIS in respect to low interstitial steel CrNi
which offers a distinctly lower proof strength and fatigue limit [5].
130 6 Assessment

6.1.2 Wear Behaviour

Scratch tests have shown [2] that the specific scratch energy es (J/mm3) increases
from 5.76 for CrNi to 12.81 for MnCr82 and 15.10 for MnC. The so-called fab-
value of these three steels is about the same, i.e. 0.82–0.85 on a scale of 0 to 1. At
fab = 1 a chip is cut without deformation on either side of the groove, at fab = 0
the groove volume is plastically displaced into ridges along the groove just by
ploughing without any removal of material. These results show that the energy es
to cut a chip is almost tripled by a high content of interstitials. The equally high
fab-value of the three steels supports the view that plasticity is more or less con-
fined to the chip and only a minor part is spent on the vicinity of the grove, i.e. on
work hardening of the surface. This lack of substantial cold work hardening is
most likely the reason why the abrasion resistance of HIS CN96 is by only 10 %
higher than that of low interstitial steel CrNi (Table 4.1).
In contrast impact wear by mineral particles is accompanied by severe defor-
mation of the wear surface into which mineral debris is embedded. The wear
resistance of CN96 is higher by about 100 %. The repeated impacts bring about a
thin surface layer of amorphous structure followed by a nanocrystalline one below,
but only on CN96 and MnC and not on low interstitial CrNi (Sect. 3.2.4.1). This
seems to be related to a vacancy-interstitial interaction [6]. In cavitation the sur-
face is impacted by water jets of imploding bubbles. Here the wear resistance of
CN96 is higher by more than an order of magnitude compared to CrNi (Table 4.1)
and the wear surface is not affected by embedded mineral debris. It is interesting to
note that the high carbon grade MnC is by far less resistant to wear by cavitation
than the high interstitial grade CN96 although their rate of cold work hardening is
equally high (Table A2). It is obvious to assume that the higher concentration of
free electrons in the steel with C ? N (Sect. 3.1.1) promotes ductility and delays
the initiation of cracks in the process of delaminating flakes of material from the
surface. This is seen parallel to the much higher specific fracture energy Ws of
CN96 in tensile tests compared to MnC (Table A1).
Just as the well known high carbon Hadfield steel MnC, the new high interstitial
steels promise a high wear resistance as long as sufficient work hardening of the
surface is involved in the wear process. In contrast to MnC the new HIS passivate
because they contain about 18 mass % Cr. This is of advantage if wear occurs in a
corrosive environment.

6.1.3 Corrosion Resistance

Wet corrosion in aqueous solutions tends to attack the weakest spot on the surface
of stainless steels which is generally characterised by a reduced concentration of
passivating elements. In this respect the chemical homogeneity is of major
importance. As shown in Sect. 3.1.2 a high concentration of free electrons in HIS
6.1 From Structure to Properties 131

enhances short range atomic ordering and thus lowers short range decomposition
on a nanoscale. The accumulation of chromium atoms in clusters is reduced and
the solute atoms are spread more evenly. As microsegregations are the outmost
result of short-range decomposition, joint alloying with C ? N promises a bene-
ficial effect not only on the nanoscale but also on the microscale distribution of
solute atoms and thus on the corrosion resistance.
This is the more important as manganese, added to promote the solubility of
nitrogen, is generally seen as unfavourable in respect to corrosion resistance [7].
On the other hand nitrogen is known to improve the resistance to pitting corrosion
[8] and several explanations have been forewarded [9, 10]. It was also shown that
carbon raises the pitting resistance if alloyed together with nitrogen [11]. Calcu-
lated phase diagrams (Figs. 2.9 and 2.10) reveal that the constitution of HIS allows
of adding molybdenum to further increase the resistance to pitting which was
verified experimentally (Sect. 4.3). The addition of copper to reduce general
corrosion is limited by the evolution of gas from the melt (Fig. 2.11) and by a
steep rise of TP, the temperature of beginning precipitation from austenite
(Fig. 2.12).
In conclusion one can say that the beneficial effect of C ? N makes up for the
detrimental one of manganese, so that the new HIS come close to standard steel
CrNi in respect to general corrosion. In respect to pitting corrosion they are dis-
tinctly better, though. A drawback of HIS is their sensitisation to intercrystalline
corrosion, if the critical cooling time is surpassed. This may limit the size of the as-
quenched cross section.

6.1.4 Nonmagnetic State

The new HIS are paramagnetic down to a temperature of 4 K and their suscep-
tibility is quite small (Sects. 3.1.3.2 and 4.4). Their relative permeability remains
at a low level even after plastic deformation and deep freezing indicating the
absence of ferromagnetic traces. The austenite is very stable which amounts to a
nonmagnetic behaviour of workpieces.

6.2 From Manufacture to Application

The feasibility of manufacturing at reasonable costs is a prerequisite of transferring


the excellent properties of new HIS to products. At this early stage of pilot pro-
duction no serious assessment of costs is possible. However, the exchange of
nickel by manganese makes the melt cheaper and the high strength tends to save
cross-section, respectively weight of workpieces.
132 6 Assessment

6.2.1 Constitution and Hot Manufacture

The higher the temperature and the slower the cooling rate, the closer the met-
allurgical processes come to equilibrium represented by phase diagrams. Therefore
the results of respective calculations, visualised in Figs. B1 to B10, contain
valuable information on the temperature range of solidification, hot working and
solution annealing, especially if slow-cooling heavy cross-sections are concerned.
The phase diagrams including the new HIS clearly point out that below the high
temperature range of homogenous austenite a forbidden zone of carbide and nitride
precipitation exists, which has to be quickly transgressed by quenching.
The solidus temperature TS decreases and the begin of precipitation at TP
increases, if the interstitial content is raised (Fig. 5.1). This reduces the range of
homogenous austenite in which hot working is performed. The situation is
aggravated by enhancing the rate of solidification as shown by a Scheil simulation
in Fig. 5.2. Therefore soaking at the initial temperature of hot working is important
to reduce the effect of segregation or even eliminate minor amounts of liquid phase
which is most detrimental in respect to hot workability. Some deformation below
TP is possible, if continuous straining leads to a dispersion of precipitates instead
of grain boundary decorations which promote tearing. Quenching at the end of hot
working, before precipitation has occurred, saves the cost of solution annealing
and promotes grain refinement which enhances the yield strength.
Of the new HIS (No. 1–11 in Table 2.1) GCN65 contained a few percent of d-
ferrite indicating the lower end of interstitial content. TL and TS of this alloy are
about 1380 and 1340 °C, respectively (Fig. B7). In comparison TL of the low
interstitial standard steels Cr18Ni10 and Cr17Ni12Mo2 amounts to about 1,460
and 1450 °C and TS to about 1430 and 1420 °C, respectively [12]. In these grades
precipitation is not a problem as far as hot working is concerned, but d-ferrite
exists in equilibrium down to 1300 and 1250 °C, respectively. If segregation is
taken into account areas of weaker d-ferrite are to be expected at the beginning of
hot working. Under certain conditions this may lead to tearing. This weakness is
not encountered in HIS which, however, require higher deformation stresses
because of solid solution hardening by the interstitial elements.
The constitution of the new HIS allows of alloying Mo (Fig. 2.9) to enhance the
resistance to pitting corrosion. Grades with 1 mass % Mo have been manufactured
(Figs. B4 and B5). The addition of Cu is restricted by the evolution of N2 gas
(Fig. 2.11) and by a higher TP (Figs. 5.1 and B6).
The forbidden zone of unwanted precipitation starts at TP and ends at about
500 °C. The lower TP, the lower is the temperature range of precipitation which is
thereby retarded. In Fig. 2.3 the phase field of homogeneous austenite extends to
the lowest temperature TP = Top at (C/N)op. This is made use of in steel CN85
(Fig. B1b), which is quite close to the optimal conditions marked by (+). The
higher carbon content of CN96 or CN107 raises TP and thus reduces the critical
cooling time (Fig. 5.13). Below 500 °C, aging occurs in cold worked HIS
(Fig. 5.14a). At temperatures of C550 °C a sensitization to intercrystalline
6.2 From Manufacture to Application 133

corrosion is observed (Sect. 4.3.3) caused by a faint precipitation along grain-


boundaries followed by a severe one between 700 and 1000 °C (Figs. 5.9 and
B11). Any stay in this temperature range from &500 °C to TP (Fig. 5.1) is likely
to impair the structure of homogeneous austenite. It should be avoided or repaired
by solution annealing and quenching. This applies to stress relief annealing,
welding, brazing, hot coating and other hot processes. Creep loading is a tentative
exception and the only diversion from a homogeneously austenitic structure (Sect.
4.1.4).
Quenching from solution anneal temperature TSA (Table A16) must stay below
the critical cooling time tc10/7 to prevent precipitation. This time becomes the
shorter the higher the interstitial content and, as the nitrogen content of all new
HIS is close to 0.6 mass %, the higher the C/N ratio is (Fig. 5.13). Therefore the
steel selection has to be adapted to the as-quenched cross-section of a workpiece.
This intense sensitivity to the cooling rate is not observed in standard stainless
austenitic steels but is known from ferritic steels or ferritic-austenitic duplex steels
[13, 14].
As described in Sects. 5.1 and 5.2 sand castings, centrifugal castings and ingots
were produced on an industrial scale. The latter were electro-slag-remelted and hot
worked by rolling and forging to different semi-finished products, solution
annealed and quenched. No unexpected incident or scrap occurred.

6.2.2 Workhardening and Cold Manufacture

Drawing, cold forging and coining operations are chipless, i.e. the degree of
deformation stays below the tearing limit. Blanking, punching and machining go
beyond this limit to cut off material. In both cases the Rp0.2 proof strength is a
guideline for the stress required to initiate deformation. The ultimate tensile
strength Rm, although not a true stress, corresponds to uniform elongation Au and
just avoids necking in chipless deformation. The true fracture strength R stands for
a separation or cut, if we set aside the different stress states. If we compare e.g. the
high interstitial steel CN96 with the low interstitial standard steel CrNi (Table A1),
we obtain the following ratios: Rp0.2 ? 600/221 = 2.71, Rm ? 1020/592 = 1.72,
R ? 2547/1930 = 1.32. These results indicate that, at a given tool geometry, the
new HIS require distinctly higher stresses to initiate plastic flow and keep it up.
This calls for stronger machines to provide sufficient force but also bears on the
tools. Thus the sheet thickness for blanking may have to be reduced and the
slenderness of punched holes as well.
The degree of deformation that a material will survive is strongly depending on
the stress state. The elongation ratios of the above high and low interstitial steels
are AU ? 61/70 = 0.87 and A ? 74/83 = 0.89. They stay below one but still
promise excellent formability of HIS, provided the tools can take the load.
Moderate heating up to e.g. 200 °C will help to lower the required flow stress (see
Table A8) and enhance the obtainable deformation.
134 6 Assessment

The ratio of the workhardening exponents (Table A2) amounts to n1 ? 0.92/


0.66 = 1.39. It underlines that the resistance to plastic flow of CN96 considerably
surpasses that of CrNi. This is a disadvantage of cold forming HIS but an
advantage of cold formed workpieces which offer a higher strength. It may further
increase by natural age hardening at room temperature (Fig. 5.14b). A heat
treatment of artificial aging will anticipate and define the increase of hardness
(Fig. 5.14a). In Sect. 3.2.1 it was shown, that workhardening is brought about by a
sequence of structural changes at such a rate that necking is delayed and elon-
gation is enhanced. In consequence workhardening promotes strength and ductility
simultaneously which is expressed by the specific fracture energy Ws containing
both properties. The ratio of Ws is 675/422 = 1.60 (Table A1). This demonstrates
that CN96 requires 60 % more energy than CrNi to deform a unit volume, which
has to be provided by the press. In service CN96 may consume 60 % more energy
before fracture.

6.2.3 Application

The new HIS offer some excellent in-service properties (Chap. 4). They are related
to structural features which were studied in great depth (Chap. 3). This sound
structural foundation, summarised in Sect. 6.1, adds reliability to the properties
measured. It is the combination of two or more properties which offers new areas
of application. All of these profit from lower alloy costs.
The forbidden temperature range of precipitation (Sect. 6.2.1) calls for a suf-
ficient quenching rate. Therefore applications of moderately thick cross-sections
are to be preferred. Although thin, sheet material is not seen in the centre of
application because of restrictions in respect to welding (Sect. 5.5). The stress
seems to rest on castings, hot rolled rods and rings as well as forgings, all
machined to final size. Strip may be cold rolled and wire cold formed by drawing
and forging as far as permitted by workhardening (Sect. 6.2.2).
In the following a few combinations of properties are discussed to demonstrate
the potential of new HIS in application.
Yield strength + corrosion resistance. In respect to standard CrNi steel the yield
strength of HIS is higher by a factor more than 2.5. This offers the chance to
reduce the cross-section and save weight, i.e. go for light-weight construction e.g.
in transportation.
Workhardened + nonmagnetic corrosion resistant. Cold forging and surface
rolling led to a surface hardness of 60 HRC in rings made of CN107 for roller
bearings [15]. As this level of hardness is reached without ferromagnetic a-mar-
tensite, such rings combined with ceramic balls may be used as bearings in the
vicinity of strong magnetic fields as demonstrated on a lab scale by the Schaeffler
KG in DE 97424, Schweinfurt (Fig. 6.1a). Intermittent heating hardly affects the
hardness at room temperature, although the hardness tested at elevated temperature
declines (Fig. 6.1b).
6.2 From Manufacture to Application 135

Fig. 6.1 Cold worked


surface of the inner ring of a
roller bearing made of CN107
with a bore diameter of
25 mm. a Scatter band of the
hardness profile HV1
measured along three paths
(see arrows). b(1) Hardness
HV0.05 of the cold worked
surface measured at room
temperature after holding at
elevated temperature for 2 h
or b(2) hot hardness
measured at the elevated
temperature after holding for
2h

Another application is seen in sea water. Here the higher resistance to pitting
corrosion of HIS compared to martensitic stainless steels could be an advantage for
bearings.
Energy consumption + corrosion resistance. In road construction the hillside
flanks are sometimes covered by a net of thick wire to catch falling rock. Their
energy has to be consumed by the protective structure even at subzero tempera-
tures. The exceptionally high specific fracture energy Ws of the new HIS offers an
effective energy consumption which even increases down to -100 °C (Table A6).
As HIS are stainless no protective coating is required. In contrast to their appli-
cation under tensile loading a vehicle crash impact occurs at a lower hydrostatic
stress, i.e. at an even higher ductility.
At ballistic velocity of an impact, adiabatic shear bands are formed [16] which
localize the strain and reduce the consumed energy (Sect. 3.2.3.2). Discs of CN107
and CN96, each 5 or 10 mm thick, were perforated by standard soft-core
ammunition 7.62 9 51 WK (Nato level 1, STANAG 4569) at a velocity VZ of
835–840 m/s [17]. In an austenitic high nitrogen steel a ballistic limit of 535 m/s
was found [18]. This suggests that a hard layer on top of HIS is liable to reduce the
velocity by flattening or fracturing the bullet. In such a hard ductile-sandwich
structure HIS could be a most effective ductile partner. A comparison of tensile
stress–strain curves in Fig. 6.2 points to a higher specific fracture energy of
136 6 Assessment

Fig. 6.2 Engineering stress–


strain curves, full
lines = measured, dashed
line = hypothetic. Steel
A = low alloy martensitic
steel with about 0.3 mass C,
steel B = high alloy
maraging steel X2NiCoMoTi
18-8-5 with 0.7 mass Ti, steel
CN96 = solution annealed,
CN96-40 = cold streched by
40 %, CN96-op = after
optimal cold working

solution annealed steel CN96 (perforated) but a much higher yield strength of
steels A and B (not perforated). After 40 % of cold stretching grade CN96-40
comes closer to steel B and after optimal cold working CN96-op may approach
steel A. This leads to the proposal to generate a hard layer by surface rolling or
severe peening a HIS and create an in situ sandwich structure.
Wear and corrosion resistance. As shown in Table 4.1, the resistance to wear
increases, if the degree of work hardening involved is raised. Compared to low
interstitial standard steel CrNi the resistance Wab1 of high interstitial steel CN96 to
abrasion is improved by a factor of only 1.1 and stays below that of Hadfield steel
1
MnC which is not stainless though. The resistance Wimp of CN96 to impact wear is,
however, raised by a factor of 1.99, which is above that of MnC. Finally the factor
of improving the resistance to cavitation Wcav1 is up to 11.6, i.e. much higher than
that of MnC (1.69). In contrast to impact wear by mineral particles no mineral
debris are imbedded in the wear surface.
These results suggest that the new HIS are suited to replace stainless CrNi steels
in case of wear by impacting particles, and also replace steel MnC in aggressive
surroundings as e.g. in deep pit or marine mining. In these applications CrNi or
MnC castings often suffer from deformations because of unpredictably high
service loads. In this respect the higher strength of HIS is bound to be an
advantage. The highest profit is expected of applying HIS to fluid flow machines as
e.g. pumps and armatures. In this case the corrosion resistance may be enhanced
by alloying molybdenum.
Strong + nonmagnetic. Nonmagnetic steels are used in the vicinity of alter-
nating currents to prevent unwanted heating of structures and tools or near direct
currents to prevent diverting of the magnetic field. The coils of transformers and
magnets exert forces which are e.g. counteracted by nonmagnetic frames. Low cost
and high strength HIS may be used to about -100 °C because of embrittlement
6.2 From Manufacture to Application 137

below (Figs. 4.4 and 4.10). Cold expanded retaining rings of moderate cross-sec-
tion may be suited to hold the wiring on electric generator shafts.
Elevated temperature service. The hot strength of CN85 derived in tensile tests
up to 700 °C (Fig. 4.6) is not above that of hotwork tool steel H11. However,
aging of cold worked CN96 at temperatures up to 550 °C keeps the hardness,
measured at RT, at or above the initial level (Fig. 5.14a). This means that inter-
mittent heating by friction or by residual heat after shut down is not necessarily
accompanied by a drop of strength.
The resistance of CN85 to short-time creep exceeds that of hot work tool steel
H11 the more, the higher the temperature and the longer the duration of the test
(Table A9, Fig. 4.9).
Prestraining seems to have little effect on the time to creep rupture of CN85 but
raises the initial strength. Cold working is, however, confined to smaller cross-
sections. This may be partly overcome by semi-hot working, to raise the initial
strength and induce a dispersion of precipitates. Mechanical loading is usually
higher during the stroke of a hot work press but thermal fatigue on the tool surface
may be more severe in a die-caster and the more so, because the thermal expansion
of austenitic HIS is higher and the thermal conductivity lower than that of mar-
tensitic tool steels. Thus the application of HIS for tooling is open for shop floor
trial.
Body friendly steel. About 10 % of the human population are allergetic to
nickel, which may cause local inflammations of the skin or the tissue inside the
body which stays in prolonged contact with e.g. zippers, wrist watches, costume
jewellery, dental braces, piercings or implants [19]. This does not only apply to Ni
coated parts but also to those made of stainless CrNi steels. High nitrogen steels
were produced by pressure metallurgy and Ni exchanged by Mn. This is in line
with high interstitial steels which do not require costly pressure equipment. By
selecting the scrap the Ni content could be reduced in respect to that of the melts
listed in Table 2.1. The addition of Mo would improve the corrosion resistance.

6.3 Pros and Cons of HIS

The arguments pro and contra the use of new high interstitial austenitic steels are
briefly summarised.

6.3.1 Pros

The unique combination of high proof strength (Rp0.2 & 600 MPa) and cold work
hardening exponent (n1 & 0.9) combined with a fracture elongation of A &70 %
amounts to a tensile fracture energy between 600 and 700 J/cm-3 which is way
138 6 Assessment

above that of conventional stainless and other steels. In short: Stainless austenitic
HIS are strong and tough!
The intensive work hardening enhances the resistance to wear by impacting
mineral particles and to wear by cavitation. In respect to the latter, HIS are proved
to be superior to Hadfield steel, which is not stainless.
The combination of high strength and good corrosion resistance of HIS espe-
cially in respect to pitting is unique. Quenched and tempered stainless steels may
attain a higher proof strength but are less corrosion resistant. This applies vice
versa to stainless duplex steels.
HIS are nonmagnetic even at cryogenic temperatures or after severe cold
working, which can lead to a hardness of 60 HRC without a trace of a-martensite.
They even retain good properties after moderate heating.
The industrial manufacture of HIS at normal pressure of air did not pose any
special problems. The costs are reduced by refraining from alloying nickel.

6.3.2 Cons

The numerous positive properties of HIS are contrasted by two negative ones.
The first one is concerned with a tendency for precipitation, if the cooling rate
during quenching from solution anneal temperature is too slow. This implies that
the cross section of work pieces is limited and has to be matched with the C ? N
content and C/N ratio of the steel, if a sensitisation to intercrystalline corrosion or
embrittling grain boundary decoration is to be avoided. This applies also to the
heat affected zone during welding.
The second one is represented by an embrittling influence of nitrogen at subzero
temperatures which is actually of ductile character. In notch impact tests the
ductile-to-brittle transition temperature at KV = 100 J is close to -100 °C.
Therefore the new HIS may be applied in the total range of climatic temperatures
but not in the deep cryogenic range. Compared to HNS of equal interstitial content
the DBTT of HIS appears to be lower, though [20–22].

References

1. Nikulin I, Kaibyshev R (2011) Deformation behavior and the Portevin-Le Chatelier effect in
a modified 18Cr-8Ni stainless steel. Mater Sci Engin A 528:1340–1347
2. Schmalt F (2004) Nutzung der Löslichkeit von C ? N in nichtrostenden Stählen, doctoral
thesis, Ruhr University Bochum, see also Fortschr. Ber (2005) VDI 5-702, VDI Verlag,
Düsseldorf
3. Tomota Y, Nakano J, Xia Y, Inoue K (1998) Unusual strain rate dependence of low
temperature fracture behavior in high nitrogen bearing austenite steels. Acta Mater
46:3099–3108
References 139

4. Mughrabi H (2010) Fatigue, an everlasting materials problem—still en vogue. 10th


International Fatigue Congress (Fatigue 2010). Procedia Engineering 2:3–26
5. Nebel T, Martin U, Eifler D (2001) Wechselverformungsverhalten metastabiler austenitischer
Stähle. HTM 56:314–320
6. Petrov Y (2012) Surface structure of different interstitial austenitic steels after impact wear.
Int J Mat Res 103:551–553
7. Jargelius-Pettersson RFA (1998) The influence of N, Mo and Mn on the microstructure and
corrosion resistance of austenitic stainless steels, doctoral thesis, Royal Institute of
Technology, Stockholm. ISBN 91-7170-337-3
8. Forchhammer P, Engell HJ (1969) Untersuchungen über den Lochfraß an passiven
austenitischen Chrom-Nickel-Stählen in neutralen Chloridlösungen, Werkstoffe u.
Korrosion 20:1–11
9. Grabke HJ (1996) The role of nitrogen in the corrosion of iron and steels. ISIJ Intern
36:777–786
10. Mudali UK, Ningshen S, Raj B (2009) Passive films and localised corrosion—role of
nitrogen. In: Svyazhin AG, Prokoshkina VG, Kossyrev KL (ed) Proceedings of 10-th
international conference on high nitrogen steels (HNS 09), MISIS, Moskau, 6–8 July 2009,
pp 271–280
11. Bernauer J (2004) Einfluss von Kohlenstoff als Legierungselement in stickstofflegierten
Chrom-Mangan Stählen, doctoral thesis, ETH Zürich, No. 15457
12. Berns H (2002) Stainless steels suited for solution nitriding. Mat-wiss u Werkstofftechn
33:5–11
13. Oppenheim R (1982) Güteeigenschaften des Superferrit X1CrNiMoNb 28-4-2 für den
Chemie-Apparatebau. Thyssen Edelstahl Techn Ber 8:97–110
14. Gümpel P, Chlibec G (1985) Untersuchungen über das Werkstoffverhalten des ferritisch-
austenitischen Stahles X2CrNiMoN22-5. Thyssen Edelstahl Techn Ber 11:3–8
15. Riedner S, Berns H, Tyshchenko AI, Gavriljuk VG, Schulte-Noelle C, Trojahn W (2008)
Nichtmagnetisierbarer warmbeständiger nichtrostender Stahl für Wälzlager. Mat-wiss und
Werkstofftechn 39:448–454
16. Armstrong RW, Walley SM (2008) High strain rate properties of metals and alloys. Int Mater
Rev 53(3):105–128
17. Berns H, Riedner S, Gavriljuk V, Petrov Y, Weihrauch A (2001) Microstructural changes in
high interstitial stainless steels due to ballistic impact. Mater Sci Eng A 528:4669–4675
18. Shi J, Dong H, Liu YL, Gu YL, Rong F, Hui WJ, Speidel MO (2004) Ballistic behavior of
nitrogen alloyed austenitic steel plates for anti-terrorist use: steel grips suppl. High Nitrogen
Steels 2:239–245
19. Uggowitzer PJ, Magdowski R (1996) Nickelfree high nitrogen austenitis steels. ISIJ Intern
36:901–908
20. Harzenmoser MAE (1990) Massiv aufgestickte austenitisch-rostfrei Stähle und Duplexstähle,
doctoral thesis, ETH Zürich
21. Riedner S (2010) Höchstfeste nichtrostende austenitische CrMn-Stähle mit (C ? N), doctoral
thesis, Ruhr-University, Bochum
22. Bernauer J (2004) Einfluss von Kohlenstoff als Legierungselement in stickstofflegierten
Chrom-Mangan Stählen, doctoral thesis, ETH Zürich, No 15457
Appendix A
Tables

Table A.1 Tensile properties of HIS and reference steels at RT


Steel CN85 CN85x CN96 CN107 MnCr82 CrMnN MnC CrNi MnCr77
Rp0,2 [MPa] 552 561 600 604 494 626 370 221 443
Rm [MPa] 1000 1002 1020 1075 951 1014 829 592 881
R [MPa] 2136 2143 2547 2545 2635 2679 1131 1930 1635
Au [%] 53 52 61 62 68 44 45 70 86
A [%] 67 67 74 74 78 63 46 83 100
Z [%] 64 61 69 65 68 77 33 86 –
Ws [J/cm3] 603 607 676 694 651 569 330 422 751

Table A.2 Constants of the Ludwigson fit (Fig. 4.1b) at RT and of CN96 at subzero testing
temperature TT [°C]
Steel K1 n1 K2 eK2 -n2
CN85x 2260 0.451 6.05 424 19.50
CN96 2148 0.919 6.37 582 0.02
CN107 2326 0.924 6.37 584 0.01
GCN65 1916 0.513 5.94 380 20.89
GCN88 2254 0.530 6.07 434 17.86
GCN98 2290 0.546 6.13 460 14.89
GCN115 2600 0.579 6.11 452 12.72
GCN85 2108 0.551 6.03 417 15.58
MnCr82 2343 0.616 6.04 420 10.36
CrMnN 2115 0.842 6.40 603 0.07
(continued)

H. Berns et al., High Interstitial Stainless Austenitic Steels, 141


Engineering Materials, DOI: 10.1007/978-3-642-33701-7,
Ó Springer-Verlag Berlin Heidelberg 2013
142 Appendix A

Table A.2 (continued)


Steel K1 n1 K2 eK2 -n2
MnCr77 1919 1 6.10 446 -10-5
CrNi 1544 0.661 5.39 219 6.85
MnC 2754 0.888 5.98 395 4.679
CN96
TT
0 2222 0.927 6.44 629 0.015
-30 2282 0.936 6.58 718 0.013
-60 2499 0.963 6.66 784 0.007
-80 2490 0.949 6.75 851 0.008
-100 2617 0.940 6.85 944 0.011

Table A.3 Effect of Mo and Cu on the tensile properties of HIS at RT


Steel CN96 CN94Mo1 CN96Cu2 CN107 CN103Mo1
Rp0,2 [MPa] 600 571 544 604 575
Rm [MPa] 1020 999 997 1075 1060
R [MPa] 2547 2428 2243 2545 2441
Au [%] 61 58 56 62 61
A [%] 74 72 69 74 74
Z [%] 69 43 42 65 42
Ws [J/cm3] 676 646 608 694 673

Table A.4 Effect of interstitial content on the tensile properties of HIS castings at RT, N &
0.6 mass %
Casting Centrifugal Sand
Steel GCN65 GCN88 GCN98 GCN1,15 GCN85
Rp0,2 [MPa] 422 501 491 554 477
Rm [MPa] 779 910 919 1040 831
R [MPa] 1586 1661 1597 1823 1278
Au [%] 46 49 47 58 44
A [%] 61 59 54 65 46
Z [%] 63 54 51 49 41
Ws [J/cm3] 423 472 433 585 329

Table A.5 Grain size of as-quenched steels, ASTM number according to ISO 643 C1
Steel ASTM Steel ASTM
Number Number
CN85x 6 GCN85 -0.2
CN96 5.2 GCN82 5.4
CN107 3.2 CrMnN 9
CN94Mo1 7 CrNi 5.7
CN103Mo1 5 MnC 4.8
CN96Cu2 6
Appendix A 143

Table A.6 Tensile properties of HIS and reference steel CrMnN at subzero testing temperatures
TT
TT [°C] 22 0 -30 -60 -80 -100 -196
CN107 Rp0,2 [MPa] 604 644 693 775 878 926 1393
Rm [MPa] 1075 1097 1161 1236 1325 1386 1613
R [MPa] 2545 2648 2754 2797 2704 2674 1702
Au [%] 62 58 58 55 52 52 6
A [%] 74 69 70 69 65 66 6
Z [%] 65 63 64 61 61 58 2
Ws [J/cm3] 694 687 736 749 776 811 75

CN96 Rp0,2 [MPa] 600 652 733 783 857 945 1419
Rm [MPa] 1020 1057 1122 1204 1258 1350 1679
R [MPa] 2547 2267 2418 2782 2831 2902 1854
Au [%] 61 57 54 56 54 50 6.5
A [%] 74 71 69 70 67 63 6.5
Z [%] 69 70 66 61 62 63 6.5
Ws [J/cm3] 676 670 706 781 745 821 96

CN85 Rp0,2 [MPa] 552 592 675 719 824 891 –


Rm [MPa] 1000 1030 1100 1160 1250 1310 –
R [MPa] 2136 2313 2412 2915 2704 3026 –
Au [%] 53 59 57 54 48 46 –
A [%] 67 75 72 69 63 61 –
Z [%] 64 66 64 65 62 62 –
Ws [J/cm3] 603 698 723 736 730 739 –

CN85x Rp0,2 [MPa] 561 596 648 733 829 855 –


Rm [MPa] 1002 1086 1117 1197 1279 1339 –
R [MPa] 2143 1923 2259 2519 2546 2393 –
Au [%] 52 52 54 51 54 54 –
A [%] 67 65 69 66 72 67 –
Z [%] 61 51 61 60 58 47 –
Ws [J/cm3] 607 667 706 733 856 836 –

GCN85 Rp0,2 [MPa] 477 516 553 614 669 718 –


Rm [MPa] 831 878 925 977 1052 1122 –
R [MPa] 1278 1315 1554 1687 1623 2311 –
Au [%] 44 48 43 38 45 43 –
A [%] 46 54 48 42 47 43 –
Z [%] 41 42 48 47 40 52 –
Ws [J/cm3] 329 411 394 400 431 421 –

CrMnN Rp0,2 [MPa] 626 694 767 801 917 1012 –


Rm [MPa] 1014 1065 1149 1223 1295 1409 –
R [MPa] 2679 – – – – – –
Au [%] 44 44 47 47 48 45 –
A [%] 63 64 64 64 65 60 –
Z [%] 77 73 73 63 64 62 –
Ws [J/cm3] 570 609 672 726 778 765 –
144 Appendix A

Table A.7 Tensile properties of CN85 at elevated temperatures, after heating the specimens
were soaked for 1 h at the testing temperature TT
TT (°C) Rp0,2 [MPa] Rm [MPa] Au (%) A (%)
20 561 1002 52 67
300 312 760 54 70
400 319 679 39 57
500 259 637 38 47
550 235 607 33 42
600 213 549 30 37
625 216/2071) 482 29 33
650 203 479 24 27
700 212 413 19 26
1)
Upper/ lower yield point

Table A.8 Loss of strength at slightly elevated testing temperature TT of specimens cold
stretched by 20 % and aged at TA
Steel CN85 CN96
TA [C°] – 360, 2 h – 360, 2 h
TT [°C] 20 100 130 150 22 100 130 150
Rp0,2 [MPa] 1017 956 933 866 1092 964 920 891
Rp1 [MPa] 1101 969 946 884 1092 950 912 886
Rm [MPa] 1235 1110 1084 1053 1247 1138 1109 1096
R [MPa] 2532 2148 2198 1918 2676 2687 2735 2712
Ag [%] 26 23 23 27 29 33 31 31
A [%] 37 35 33 38 46 46 44 44
Z [%] 62 60 58 55 66 67 68 67
Ws [J/cm3] 441 371 344 383 541 498 461 449

Table A.9 Results of creep tests under constant temperature and stress of steel CN85,
SA = solution annealed, CW = additionally cold worked
Temperature Stress e_ min e(_emin ) t(_emin ) ef tf
[°C] [MPa] [% / h] [%] [h] [%] [h]
SA 600 350 3.3510-03 1.78 47.99 6.48 239.39
370 5.2510-03 1.64 24.21 6.06 161.4
400 4.3310-02 3.42 4.69 9.37 27.61
650 300 1.1410-06 2.54 91.78 6.32 420.15
350 6.1410-02 5.39 5.84 8.42 35.23
400 1.5110-01 5.58 1.80 9.12 14.49
700 270 8.1210-02 1.76 5.40 19.95 74.15
300 1.4710-01 1.95 4.84 12.2 29.63
(continued)
Appendix A 145

Table A.9 (continued)


Temperature Stress e_ min e(_emin ) t(_emin ) ef tf
[°C] [MPa] [% / h] [%] [h] [%] [h]
CW 600 370 2.7410-03 0.18 18.48 0.80 173.42
400 3.4410-03 0.32 14.79 1.04 86.88
650 320 6.8710-03 0.27 16.67 1.75 106.58
350 8.4110-03 0.13 5.45 0.52 44.20
370 7.0610-03 0.31 3.60 1.54 18.10
700 250 2.7710-02 0.41 15.13 9.57 105.84
270 2.2010-02 0.12 8.64 6.94 64.19

Table A.10 Vickers hardness measured at room temperature: macrohardness HV30, micro-
hardness HV0.1 on fracture face after tensile test at RT
Steel HV30 HV0.1 HV30
CN85 237 619 after tensile test at °C
500 550 600 650 700
head1) 257 263 255 250 256
gauge2) 476 441 415 387 371
CN96 270 590 after % elongation by cold drawing
16 20 28 37
360 363 400 423
CN107 270 630
CN94Mo1 277
CN103Mo1 276
CN96Cu2 254
GCN65 264
GCN88 265
GCN98 271
GCN115 278
GCN85 244
CrNi 141 460
MnC 211 741
CrMnN 262 526
MnCr82 248 566
MnCr77 212 547
1) 2)
undeformed specimen head, within deformed gauge length

Table A.11 ISO-V notch impact toughness KV at RT and ductile-to-brittle transition temper-
ature DBTT at KV = 100 J, in brackets = specimens of transverse taking
Steel KV [J] DBTT [°C]
CN85x 272 -88
CN85 325 (233) -94 (-112)
CN96 364 (296) -91 (-49)
(continued)
146 Appendix A

Table A.11 (continued)


Steel KV [J] DBTT [°C]
CN107 333 (297) -49 (-53)
CN94Mo1 271 -
CN103Mo1 271 -
CN96Cu2 207 -
GCN65 250 -93
GCN88 265 -87
GCN98 317 -96
GCN115 249 -47
GCN85 290 -94
CrMnN 313 -21
MnCr82 187 -77
CrNi 2881) -1961)
MnC 280 -59
1)
After G. Kalla and G. Uhlig, Bänder, Bleche, Rohre (1992) 136-142

Table A.12 Results of impact wear tests: wear rate between 3000 and 12000 impacts, surface
hardness and hardness penetration measured in a section normal to the wear surface
Steel Total mass loss Wear rate Surface hardness Hardness penetration
[mg] [lg/impact] [HV0.05] [mm]
CN85 48.5 4.4 – –
CN96 61.5 4.7 613 0.4
CN107 60 5.3 726 0.4
CrMnN 53 4.6 538 0.4
MnC 58 5.2 761 0.9
MnCr77 63 5.7 547 0.3
CrNi 116 9.6 524 1.1

Table A.13 Effect of molybdenum and copper on mass loss Dm, corrosion rate v and removal w
during uninterrupted submersion tests in aqueous solutions lasting for 120 h
CN96 CN94Mo1 CN96Cu2 CN107 CN103Mo1
10 %H2SO4 Dm [g] 16.7670 8.7784 0.6995 2.2984 0.0922
v [g/m2h] 131.2 36.1 5.7 17.9 0.76
w [mm/a] 150.3 41.2 6.5 20.7 0.88
1 %HCl Dm [g] 1.9125 4.0429 3.7070 0.9131 2.0997
v [g/m2h] 19.4 42.6 38 8.9 21.5
w [mm/a] 22.3 48.7 43.4 10.3 24.7
3 %HCl Dm [g] 16.7284 13.7921 13.5929 16.1793 13.2903
v [g/m2h] 164.4 115.5 115.4 157.7 113.9
w [mm/a] 188.3 132 131.9 181.7 130.9
density [g/cm3] 7.65 7.67 7.66 7.60 7.62
Appendix A 147

Table A.14 Effect of carbon on mass loss Dm, corrosion rate v and removal w during unin-
terrupted submersion tests in aqueous solutions lasting for 120 and 600 h, respectively
120 h GCN65 GCN88 GCN98 GCN115
Dm [g] 0.1285 0.3901 0.1131 0.1138
10 %H2SO4 v [g/m2h] 1.22 3.25 1 1
w [mm/a] 1.4 3.7 1.1 1.1
1 %HCl Dm [g] 0.9632 0.8582 0.6578 0.4318
v [g/m2h] 9.2 7.1 5.5 3.6
w [mm/a] 10.5 8.2 6.3 4.1
3 %HCl Dm[g] 12.7642 14.2557 13.5281 12.7844
v [g/m2h] 121.5 118.9 112.8 106.8
w [mm/a] 139.2 136.2 129.2 122.2
600 h
10 %H2SO4 Dm [g] 0.1248 0.2421 0.1489 0.1177
v [g/m2h] 0.24 0.40 0.25 0.2
w [mm/a] 0.27 0.46 0.29 0.22
density [g/cm3] 7.65 7.65 7.65 7.65

Table A.15 Characteristic data of current density/SHE potential tests at room temperature in
(a) 0.5 m H2SO4, b 3 % NaCl (see Fig. 4.21)
(a) UR UPP UP UA UB URP UB’ iPP iP i0 iRP iB’
[V] [lA/cm2]
CN85x -0.30 -0.25 0.21 0.29 1.08 1.40 1.63 30 11000 2.1 2650 40
CN96 -0.24 -0.22 0.21 0.40 1.05 1.32 1.64 54 2000 6.2 3100 50
CN107 -0.24 -0.22 0.22 0.44 1.10 1.31 1.63 69 460 3.2 4200 70
CN94Mo1 0.11 - 0.31 0.45 1.10 1.33 1.63 16 - 6.8 3400 110
CN103Mo1 -0.025 - 0.29 0.40 1.09 1.31 1.61 10 - 2.8 4000 77
CN96Cu2 -0.11 -0.10 0.22 0.45 1.15 1.33 1.65 45 54 2.8 2800 29
GCN65 -0.07 - 0.22 0.29 1.11 1.32 1.63 - 49 1.5 930 20
GCN88 -0.07 - 0.22 0.36 1.11 1.32 1.61 - 100 1.5 4300 65
GCN98 -0.07 - 0.22 0.36 1.11 1.32 1.61 - 71 1.5 4940 40
GCN15 -0.07 - 0.22 0.36 1.11 1.32 1.61 - 63 1.5 4980 100
GCN85 -0.31 -0.25 0.21 0.28 1.12 1.32 1.62 - 8900 2.2 2950 92
CrMnN -0.04 - 0.19 0.47 1.04 1.40 1.61 17 - 2 1950 20
CrNi -0.13 -0.08 0.28 0.4 1.09 - - 2 70 1.8 - -
(b) UR [V] UP [V] UA [V] UB [V] ip [lA/cm2] i0 [lA/cm2]
2
CN85x -0.23 - - -0.19 (0.69) - 80
CN96 -0.39 -0.19 -0.05 0.20 3 2.1
CN107 -0.31 -0.16 0.05 0.91 3.2 2
CN94Mo1 -0.37 -0.10 -0.05 0.44 0.55 0.43
CN103Mo1 -0.27 -0.16 - 0.88 0.38 1
CN96Cu2 -0.29 -0.16 - -0.04 3.1 4
(continued)
148 Appendix A

Table A.15 (continued)


(b) UR [V] UP [V] UA [V] UB [V] ip [lA/cm2] i0 [lA/cm2]
GCN65 -0.34 -0.11 -0.05 0.18 7.6 4
GCN88 -0.34 -0.09 -0.07 1.24 4.4 1.4
GCN98 -0.43 -0.14 -0.02 1.25 5.6 1.4
GCN115 -0.51 -0.28 -0.13 1.28 1.3 1
GCN85 -0.20 - - 0.02 (0.65)2) - 22
MnCr821) -0.53 - - -0.03 - -
CrMnN -0.18 0.03 0.10 0.74 3.1 3
CrNi -0.10 0.06 0.08 0.26 0.9 0.8
1) 2)
Taken from [1.5] In brackets: derived from a linear plot

Table A.16 Temperature TSA and time tSA of solution annealing applied to the new HIS, pN2 is
the nitrogen pressure calculated by Thermo-Calc to retain the level of nitrogen at the surface
Steel TSA [°C] tSA [min] pN2 [mbar]
CN96 1100 30 62.2
CN107 1175 30 106.5
CN85 1100 30 59.8
CN94Mo1 1125 30 72.8
CN103Mo1 1175 30 114.6
CN96Cu2 1200 30 342.7
GCN65 1100 45 42.5
GCN88 1150 45 76.8
GCN98 1175 45 76.6
GCN115 1200 45 150.3
MnCr82 1100 30 -
CrMnN 1050 45 -
CrNi 1100 30 -
MnC 1050 30 -

Table A.17 Steel CN96 cold drawn up to 37 % reduction of cross-section. a Results of tensile
tests at RT, b Costants of the Ludwigson fit
(a) Reduction [%]
0 16 20 29 37
Rp0.2 [MPa] 600 810 988 1103 1190
Rp1 [MPa] 606 914 1103 1241 1379
Rm [MPa] 1020 1164 1243 1320 1410
R [MPa] 2547 2699 2624 2825 2766
Au [%] 61 32 26 30 12
A [%] 74 45 39 35 27
Z [%] 69 63 59 58 58
Ws [J/cm3] 676 501 469 453 368
Rp0.2/Rm [-] 0.59 0.70 0.79 0.84 0.84
(continued)
Appendix A 149

Table A.17 (continued)


(a) Reduction [%]
(b) Constants
Reduction [%] K1 n1 K2 eK2 -n2
0 2411 0.518 6.10 445 14.94
16 2224 0.283 6.06 428 21.90
20 2216 0.233 6.03 415 24.23
29 2166 0.183 5.93 377 27.97
37 2038 0.117 5.65 284 39.80

Table A.18 Liquidus temperature TL, solidus temperature TS and the begin of precipitation at TP
of HIS listed in Table 2.1 (see Figs. B1 to B10)
Steel TL [°C] TS [°C] TP [°C]
CN85x 1364 1307 1006
CN96 1362 1292 1047
CN107 1351 1256 1146
CN94Mo1 1362 1289 1031
CN103Mo1 1351 1258 1119
CN96Cu2 1352 1269 1139
GCN65 1385 1344 981
GCN88 1364 1305 1006
GCN98 1355 1270 1118
GCN115 1349 1251 1167
Appendix B
Figures

H. Berns et al., High Interstitial Stainless Austenitic Steels, 151


Engineering Materials, DOI: 10.1007/978-3-642-33701-7,
Ó Springer-Verlag Berlin Heidelberg 2013
152 Appendix B

CN85x
(a) 1500 0.8
L 1.2 p N2 [bar]
1400 L+G
L+F
L+F+A
1300 F

temperature [°C]
1200
A
F+A
1100 A+M2 N
1000

900
A+M2N+M23 C6
800
A+M2N+M 23 C6 + σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]
N [mass%]
0.8 0.6 0.4 0.2 0
(b) 1500 1p N2 [bar]
L+F+G L+G L
1400

1300 L+A F+A


temperature [°C]

1200 A
CN107
1100 A+M 23 C 6
CN96
1000 CN85
A+M2N
900
A+M2N+M23 C6
800
A+M 2N+M 23 C6 + σ
700
0 0.2 0.4 0.6 0.8
C [mass%]
0 0.11 0.250.430.67 1 1.5 2.3 4 9 ∞
C/N

(c) 1.0
A A L
0.8
mole fraction [ - ]

0.6

0.4

0.2
M2N
M23 C6
G
0 σ
700 900 1100 1300 1500
temperature [°C]
Fig. B.1
Figs. B.1–B.10 Calculated phase diagrams for the actual chemical composition of HIS listed in
Table 2.1, a Plot over C+N in which the dotted line represents the steel content, b Plot over C/N
in which the dotted line represents the C/N ratio of the steel and (+) in Fig. B.1b the TP of three
steels for comparison, c Mole fraction in dependence of temperature. The target phase field of
homogeneous austenite extends from the solidus temperature TS to the begin of precipitation at
TP, which are listed together with the liquidus temperature TL in Table A.18
Appendix B 153

CN96
(a) 1500 0.8
1.2 [bar]
L L+G
1400 L+F
F L+F+A L+A
1300

temperature [°C]
1200
A
1100 F+A

1000
A+M2N
900 A+M2N+M23C6

800
A+M2N+M23C6 + σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]
N [mass%]
0.8 0.6 0.4 0.2 0
(b) 1500 1 pN2 [bar]
L+G
1400 L L+F

1300 L+A
temperature [°C]

1200
A
1100
A+M23C 6
1000 A+M2N
900 A+M2N+M23 C6

800 A+M2N+M23 C6 + σ

700
0 0.2 0.4 0.6 0.8
C [mass%]
0 0.11 0.250.430.67 1 1.5 2.3 4 9 ∞
C/N

(c) 1.0
A A
L
0.8
mole fraction [ - ]

0.6

0.4

0.2
M2N M23C 6 G
σ
0
700 900 1100 1300 1500
temperature [°C]

Fig. B.2
154 Appendix B

CN107
(a) 1500 0.8 1.2 pN2 [bar]
L
1400 L+G
L+F
1300 F L+F+A L+A

temperature [°C]
1200 A
F+A
1100
A+M23 C6
1000
F+A
900 +M23 C 6
A+M2N+M23 C6
800
A+M2N+M23 C6 + σ
700
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
(C+N) [mass%]
N [mass%]
1 0.8 0.6 0.4 0.2 0
(b) 1500 1 pN2 [bar]
L+G
1400 L

1300
temperature [°C]

L+A
1200 A

1100
A+M23 C6
1000 A+M2N

900 A+M2N+M23 C6

800 A+M2N+M23 C6 + σ

700
0 0.2 0.4 0.6 0.8 1
C [mass%]
0 0.11 0.250.430.67 1 1.5 2.3 4 9 ∞
C/N

(c) 1.0
A A L
0.8
mole fraction [ - ]

0.6

0.4

0.2
M23 C6
M2N G
σ
0
700 900 1100 1300 1500
temperature [°C]

Fig. B.3
Appendix B 155

CN94Mo1
(a) 1500 L+G
L 0.8
1400 1.2 p N2 [bar]
L+F
L+F+A L+A
1300

temperature [°C]
1200
F+A A
1100

1000
C6
A+ M 23
900
A+M2N+M23C6
800
A+M2N+M 23C6+ σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]
N [mass%]
0.8 0.6 0.4 0.2 0
(b) 1500 1 pN2 [bar]
L+G L
1400 F+L

1300 L+A
temperature [°C]

1200 A

1100
A+M23C6
1000
A+M2N
900
A+M2N+M23C6
800
A+M2N+M23 C6 + σ
700
0 0.2 0.4 0.6 0.8
C [mass%]
0 0.11 0.250.430.67 1 1.5 2.3 4 9 ∞
C/N

(c) 1.0
L
A
A
0.8
mole fraction [ - ]

0.6

0.4

0.2
M2N M 23C 6 G
0 σ
700 900 1100 1300 1500
temperature [°C]

Fig. B.4
156 Appendix B

CN103Mo1
(a) 1500
L+G
L 0.8
1400 1.2 pN2 [bar]
L+F
1300 L+F+A
F L+A

temperature [°C]
1200 A
F+A
1100
A+M23C 6
1000

900 A+M23C6 A+M2N+M23 C 6



800
A+M2N+M23C6+ σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]

N [mass%]
1 0.8 0.6 0.4 0.2 0
(b) 1500 1 p [bar]
L+G N2
1400 L

1300
temperature [°C]

L+A
1200 A

1100 A+M23 C 6

1000 A+M2N

900
A+M2N+M23C 6
800 A+M2N+M23 C6 +σ

700
0 0.2 0.4 0.6 0.8 1
C [mass%]

0 0.11 0.25 0.43 0.67 1 1.5 2.3 4 9 ∞


C/N
1.0
(c) A
L
A
0.8
mole fraction [ - ]

0.6

0.4

0.2
M2N M23C 6
G
σ
0
700 900 1100 1300 1500
temperature [°C]

Fig. B.5
Appendix B 157

CN96Cu2
(a) 1500 0.8 1.2 pN2[bar]
L
1400 L+G
L+F
L+F+A L+A
1300
F

temperature [°C]
1200
F+A A
1100 A+M23C 6

1000
A+M2N+M23C6
900

800
A+M2N+M23C6+σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]

N [mass%]
0.8 0.6 0.4 0.2 0
(b) 1500 1p
N2 [bar]
L+G
1400 L

1300
temperature [°C]

L+A
A
1200
A+M23C 6
1100 A+M2N

1000

900 A+M2N+M23C6

800

700
0 0.2 0.4 0.6 0.8
C [mass%]

0 0.11 0.25 0.43 0.67 1 1.5 2.3 4 9 ∞


C/N
(c) 1.0
A L
A
0.8
mole fraction [ - ]

0.6

0.4

0.2
M23C 6
M2N G
σ
0
700 900 1100 1300 1500
temperature [°C]

Fig. B.6
158 Appendix B

GCN65
(a) 1500
L 1.2 p N2 [bar] L+G
1400 0.8
L+F

1300

temperature [°C]
1200 A
F+A
1100

1000
A+M2N
900

800 A+M2N+ σ
A+M2N+M 23C 6+σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]
N [mass%]
0. 6 0. 5 0. 4 0. 3 0. 2 0. 1 0
(b) 1500
1 [bar]
L
1400
L+F
1300 L+F+A
temperature [°C]

F+A
1200
F+A+M23 C 6
1100
A
1000
A+M23 C 6
A+M2N
900
A+M2N+M23 C6
800 A+σ
A+M2N+M23C6+σ
700
0 0.1 0.2 0.3 0.4 0.5 0.6
C [mass%]
0 0.11 0.25 0.43 0.67 1 1.5 2.3 4 9 ∞
C/N
(c) 1.0
A L
A
0.8
mole fraction [ - ]

0.6

0.4

0.2 F F
σ
M2N G
M 23C 6 A
0
700 900 1100 1300 1500
temperature [°C]

Fig. B.7
Appendix B 159

GCN88
(a) 1500
0.8
L 1.2 p N2 [bar]
1400 L+G
L+F
L+F+A L+A
1300 F

temperature [°C]
1200 A
F+A
1100
A+M2N
1000

900 A+M2N+M23C6

800
A+M2N+M23 C6 +σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]
N [mass%]
0.8 0.7 0. 6 0. 5 0.4 0.3 0. 2 0. 1 0
(b) 1500
1 p N2 [bar]
L+G
L
1400
L+F
1300 L+F+A
temperature [°C]

1200
A
1100
A+M23C6
1000
A+M2N
900
A+M2N+M23 C6
800
A+M2N+M23 C6 +σ
700
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
C [mass%]
0 0.11 0.25 0.43 0.67 1 1.5 2.3 4 9 ∞
C/N
(c) 1.0
L
A
0.8 A
mole fraction [ - ]

0.6

0.4

0.2
σ M2N M C F
23 6 G
0
700 900 1100 1300 1500
temperature [°C]

Fig. B.8
160 Appendix B

GCN98
(a) 1500 0.8 1.2 p [bar]
N2
L
1400 L+G
L+F
1300 F L+F+A L+A

temperature [°C]
1200 F+A A

1100
A+M23 C6
1000
F+A
900 +M23 C6
A+M2 N+M23 C6
800
A+M2N+M23 C6 +σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]

N [mass%]
0.8 0.6 0.4 0.2 0
(b) 1500
1 pN2 [bar]
L+G L
1400
L+F
1300 L+A
temperature [°C]

L+F+A
1200 A
1100
A+M23C6
1000 A+M2N
900 A+M2N+M23C6

800
A+M2N+M23C6 + σ
700
0 0.2 0.4 0.6 0.8
C [mass%]

0 0.11 0.25 0.43 0.67 1 1.5


C/N
2.3 4 9 ∞
(c) 1.0
L
A A
0.8
mole fraction [ - ]

0.6

0.4

0.2
M2N
M23 C6 G
0 σ
700 900 1100 1300 1500
temperature [°C]

Fig. B.9
Appendix B 161

GCN115
(a) 1500
0.8 1.2 p [bar]
L N2
1400 L+G
L+F
1300 F L+F+A L+A

temperature [°C]
1200 A
F+A
1100
A+M23 C6
1000
F+A
+M 23C6
900 A+M2N+M23 C6

800
A+M2N+M23 C6 +σ
700
0 0.3 0.6 0.9 1.2 1.5
(C+N) [mass%]
N [mass%]
1 0.8 0.6 0.4 0.2 0
(b) 1500
1 p N2 [bar]
L+G
1400 L L+F
1300
temperature [°C]

L+A
1200 A

1100
A+M23 C6
1000 A+M2N

900 A+M2N+M23 C6

800 A+M 2N+M 23 C6 +σ

700
0 0.2 0.4 0.6 0.8 1
C [mass%]

0 0.11 0.25 0.43 0.67 1 1.5


C/N
2.3 4 9 ∞
(c) 1.0
L
A A
0.8
mole fraction [ - ]

0.6

0.4

0.2
M2N
M23 C6 G
σ
0
700 900 1100 1300 1500
temperature [°C]

Fig. B.10
162 Appendix B

Fig. B.11 Content of pseudo-pearlite after solution annealing and isothermal holding at tem-
perature Tis for 1 h and subsequent quenching in water
About the Authors

Hans Berns born 1935, studied ferrous metallurgy, worked in the speciality steel
industry from 1959–1979, doctorate TH Aachen l964, habilitation TU Berlin l975,
held the Chair of Materials Technology at the Ruhr University, Germany, from
1979–2000, since then working as emeritus.
Valentin Gavriljuk born 1938, studied metal physics, worked at the Institute for
Metal Physics, Kiev, Ukraine as post-graduate (l962), senior scientific researcher
(1973), leading scientific researcher (1986), head of department of physical
principles for design of steel and alloys since l989, doctorate l965, habilitation
l986.
Sascha Riedner born l975, studied mechanical engineering at the Ruhr University
Bochum, Germany, majoring in engineering materials, diploma 2004, doctorate
2009, since then working in R+D and customer services at the German Specialty
Steel Company (DEW) in Hagen.

H. Berns et al., High Interstitial Stainless Austenitic Steels, 163


Engineering Materials, DOI: 10.1007/978-3-642-33701-7,
Ó Springer-Verlag Berlin Heidelberg 2013
Index

A Chemical homogeneity. See Homogeneity


Ab initio calculation, 22 Cleanness, 116
Adiabatic shear bands, 66, 135 Cleavage-like fracture, 61
Aging, 123, 134 Clustering, 28, 36, 127
Amorphous structure, 71 Cluster size, 41, 43, 127
Application, 134 Cold drawing, 123
Atomic Cold manufacture, 133
bonds, 21 Cold worked structure, 51
configuration, 22, 30 Cold work hardening, 56
distribution, 28, 46, 127 Constitution, 7
interaction, 30 C/N ratio, 9, 111, 152
Austenite interstitial content, 8
homogeneous, 7, 22, 116 steels investigated, 149
retained, 34 substitutional content, 11
stability, 128 tramp elements, 16
Contra HIS, 137
Cooling time, 120, 132
B Coordination spheres, 30
Ballistic impact, 65, 135 Corrosion, 99, 130, 134
Bearing ring, 134 current density/potential, 102, 147
Body friendly, 137 intercrystalline, 106, 118, 120
Boiling, 9 submersion, 100, 146
Brittleness. See Embrittlement Creep, 91
Cross-section, 111, 120, 132
Cyclic loading, 68
C
C/N ratio, 9, 45
C+N influence D
constitution, 8, 118 DBTT, 62, 93, 129, 138, 145
electron structure, 24, 29, 128 Degassing, 9
stacking fault energy, 45 Dilatometry, 35
strength, 88 Dislocation
Carbides, 8 barrier, 64
Casting, 113 density, 74
Chemical composition, 18 pile up, 48

H. Berns et al., High Interstitial Stainless Austenitic Steels, 165


Engineering Materials, DOI: 10.1007/978-3-642-33701-7,
Ó Springer-Verlag Berlin Heidelberg 2013
166 Index

D (cont.) Interatomic bonds, 21, 24, 127


pinning, 48 Interstitial clouds, 49
splitting, 44, 51 content, 8
Iron structure, 24

E
Electron spin resonance, 26, 38 L
Electron structure, 4, 22, 127 Liquidus temperature, 112, 149
Elevated temperature, 88, 137 Ludwigson equation, 86, 141
Embrittlement, 2, 129, 136 Ludwik equation, 51
cleavage-like, 61
covalent, 21
precipitation, 120 M
ESR, 112, 116 Machining, 125
Magnetic measurement, 36, 41, 127
Magnetic properties, 106
F Manufacture, 111, 131
Fatigue Martensite (a), 32, 34, 70
push/pull, 73 Martensite (e), 50, 54, 57, 66, 74
repeated impacts, 68 Melting, 112
rotating bending, 94, 129 Microstructure. See Structure
Fermi level, 23 Mössbauer spectroscopy, 29
Ferrite-d, 9 Multiscale approach, 4
Foaming, 9
Forging, 115
Fractography, 90, 95 N
Fracture energy, 87, 128, 134, 136 Nanocrystals, 71
Free electron, 21, 38, 127 Nitrides, 8, 66, 75, 118
concentration, 27 Nitrogen
distribution, 24 dislocation, 128
DOS, 24
grainboundary, 49, 75
G pressure, 8
Grain size, 49, 116, 142 solubility, 9
yielding, 48
Nonmagnetic, 131, 134
H
Hadfield steel, 17, 36, 47, 52
Hall-Petch equation, 49 O
Hardness, 58, 93, 135, 145 Ordering, 29, 63, 127
Heat treatment, 116
High intersticial steels (HIS), 2, 17
High nitrogen steels (HNS), 1 P
Homogeneity Paramagnetic, 41
nano scale, 36, 128 Phase diagrams, 7, 152
segregation, 28, 114 Phases encountered, 7
Hot manufacture, 132 Planar slip, 54
Hot working, 115 Precipitation
dissolution, 116
during fatigue, 75
I during impact, 66
Impact toughness, 93, 123, 145 grainboundary, 91, 120
Ingot, 112 range, 132
Interaction energy, 33 temperature, 112, 149
Index 167

Pro HIS, 137 Steels investigated, 18


Properties, 85 Strain rate, 61
corrosion, 99 Stress-strain curves, 86, 124, 136
magnatic, 106 Structure, 21, 127
mechanical, 85, 128 after loading, 47
wear, 95 as-quenched, 22
Pseudo pearlite, 115, 118 Subzero temperature, 59, 88, 89, 129
Superparamagnetic clusters, 41, 43
Surface, 114, 117
Q
Quenching, 132, 133
continuous, 120 T
interrupted, 118 Tensile properties, 47, 85, 128, 141
comparison, 87
creep, 91, 144
R elevated temperature, 88, 129, 144
Reference steels, 17 strain rate, 61, 129
Refractories, 114 subzero temperature, 59, 88, 129, 143
Remelting, 112, 116 Thermo-Calc, 7
Retained austenite, 34 Thermodynamic stability, 23, 29, 34, 46
Retaining rings, 137 Tramp elements, 16
Ring rolling, 115 TRIP, 4
Rolling, 115 Twinning, 47, 53, 58
TWIP, 3, 4, 58

S
Scheil simulation, 113 W
Segregation, 28, 111, 114 Wear, 95, 136
Selection of steel, 17 abrasion, 96
Sensitization, 120 cavitation, 99
Serration, 88, 129 impact, 68, 96, 146
Short range decomposition, 28, 36 Welding, 124
Short range ordering, 28, 36, 63, 127 Work hardening, 47, 52, 86, 133
Sigma phase, 7, 12, 15 Work hardening mechanism, 56
Solidification, 7, 113
Solidus temperature, 112, 149
Solution annealing, 116, 148 Y
Stacking fault energy, 44, 50, 127 Yielding, 48

You might also like