A review: ferritic-martensitic steels – treatment, structure and mechanical properties
The constantly growing consumption of electricity requires the development and implementation of more powerful and energy-intensive systems of the new generation. Fusion and fission reactors of the 4th generation (Gen-IV) will make it possible to cover the growing demand for electricity. Since Gen-I...
Збережено в:
Дата: | 2022 |
---|---|
Автори: | , |
Формат: | Стаття |
Мова: | English |
Опубліковано: |
Національний науковий центр «Харківський фізико-технічний інститут» НАН України
2022
|
Назва видання: | Вопросы атомной науки и техники |
Теми: | |
Онлайн доступ: | http://dspace.nbuv.gov.ua/handle/123456789/195411 |
Теги: |
Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
|
Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Цитувати: | A review: ferritic-martensitic steels – treatment, structure and mechanical properties / H.Yu. Rostova, G.D. Tolstolutska // Problems of Atomic Science and Technology. — 2022. — № 4. — С. 66-84. — Бібліогр.: 134 назв. — англ. |
Репозитарії
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-195411 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1954112023-12-05T11:57:17Z A review: ferritic-martensitic steels – treatment, structure and mechanical properties Rostova, H.Yu. Tolstolutska, G.D. Thermal and fast reactor materials The constantly growing consumption of electricity requires the development and implementation of more powerful and energy-intensive systems of the new generation. Fusion and fission reactors of the 4th generation (Gen-IV) will make it possible to cover the growing demand for electricity. Since Gen-IV reactors will operate at higher temperatures and radiation doses, the problem of selecting scientifically based structural materials arises, since conventional reactor materials are not suitable for use in such severe operating conditions. Among the structural materials under consideration for future generations of reactors, special attention is paid to 9…12% Cr ferritic-martensitic steels due to their higher radiation tolerance and excellent mechanical properties compared to traditionally used austenitic steels. This review presents the main ferritic-martensitic steels that will be used as structural materials, their structure, mechanical properties and various thermal and thermomechanical treatments applied to them. Вживання електроенергії, що постійно зростає, вимагає розробки та впровадження більш потужних та енергоємних систем нового покоління. Ядерні та термоядерні установки четвертого покоління (Gen-IV) дадуть можливість покрити зростаючий попит на електроенергію. Оскільки реактори Gen-IV працюватимуть за вищих температур і доз опромінення, виникає проблема підбора науково обґрунтованих конструкційних матеріалів, так як реакторні матеріали, що нині використовуються, не придатні для використання в таких жорстких умовах експлуатації. Серед конструкційних матеріалів, що розглядаються, для майбутніх поколінь реакторів особлива увага приділяється 9…12% Cr феритно-мартенситним сталям через їх більш високу радіаційну толерантність і відмінні механічні властивості порівняно з традиційно використовуваними аустенітними сталями. У даному огляді розглянуті основні феритно-мартенситні сталі, які будуть використовуватися як конструкційні матеріали, їх структура, механічні властивості та різні термічні та термомеханічні обробки, що застосовуються до них. 2022 Article A review: ferritic-martensitic steels – treatment, structure and mechanical properties / H.Yu. Rostova, G.D. Tolstolutska // Problems of Atomic Science and Technology. — 2022. — № 4. — С. 66-84. — Бібліогр.: 134 назв. — англ. 1562-6016 DOI: https://doi.org/10.46813/2022-140-066 http://dspace.nbuv.gov.ua/handle/123456789/195411 621.039.1(075.8) en Вопросы атомной науки и техники Національний науковий центр «Харківський фізико-технічний інститут» НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
topic |
Thermal and fast reactor materials Thermal and fast reactor materials |
spellingShingle |
Thermal and fast reactor materials Thermal and fast reactor materials Rostova, H.Yu. Tolstolutska, G.D. A review: ferritic-martensitic steels – treatment, structure and mechanical properties Вопросы атомной науки и техники |
description |
The constantly growing consumption of electricity requires the development and implementation of more powerful and energy-intensive systems of the new generation. Fusion and fission reactors of the 4th generation (Gen-IV) will make it possible to cover the growing demand for electricity. Since Gen-IV reactors will operate at higher temperatures and radiation doses, the problem of selecting scientifically based structural materials arises, since conventional reactor materials are not suitable for use in such severe operating conditions. Among the structural materials under consideration for future generations of reactors, special attention is paid to 9…12% Cr ferritic-martensitic steels due to their higher radiation tolerance and excellent mechanical properties compared to traditionally used austenitic steels. This review presents the main ferritic-martensitic steels that will be used as structural materials, their structure, mechanical properties and various thermal and thermomechanical treatments applied to them. |
format |
Article |
author |
Rostova, H.Yu. Tolstolutska, G.D. |
author_facet |
Rostova, H.Yu. Tolstolutska, G.D. |
author_sort |
Rostova, H.Yu. |
title |
A review: ferritic-martensitic steels – treatment, structure and mechanical properties |
title_short |
A review: ferritic-martensitic steels – treatment, structure and mechanical properties |
title_full |
A review: ferritic-martensitic steels – treatment, structure and mechanical properties |
title_fullStr |
A review: ferritic-martensitic steels – treatment, structure and mechanical properties |
title_full_unstemmed |
A review: ferritic-martensitic steels – treatment, structure and mechanical properties |
title_sort |
review: ferritic-martensitic steels – treatment, structure and mechanical properties |
publisher |
Національний науковий центр «Харківський фізико-технічний інститут» НАН України |
publishDate |
2022 |
topic_facet |
Thermal and fast reactor materials |
url |
http://dspace.nbuv.gov.ua/handle/123456789/195411 |
citation_txt |
A review: ferritic-martensitic steels – treatment, structure and mechanical properties / H.Yu. Rostova, G.D. Tolstolutska // Problems of Atomic Science and Technology. — 2022. — № 4. — С. 66-84. — Бібліогр.: 134 назв. — англ. |
series |
Вопросы атомной науки и техники |
work_keys_str_mv |
AT rostovahyu areviewferriticmartensiticsteelstreatmentstructureandmechanicalproperties AT tolstolutskagd areviewferriticmartensiticsteelstreatmentstructureandmechanicalproperties AT rostovahyu reviewferriticmartensiticsteelstreatmentstructureandmechanicalproperties AT tolstolutskagd reviewferriticmartensiticsteelstreatmentstructureandmechanicalproperties |
first_indexed |
2025-07-16T23:34:51Z |
last_indexed |
2025-07-16T23:34:51Z |
_version_ |
1837848488033910784 |
fulltext |
66 ISSN 1562-6016. ВАНТ. 2022. №4(140)
SECTION 2
THERMAL AND FAST REACTOR MATERIALS
https://doi.org/10.46813/2022-140-066
UDC 621.039.1(075.8)
A REVIEW: FERRITIC-MARTENSITIC STEELS – TREATMENT,
STRUCTURE AND MECHANICAL PROPERTIES
H.Yu. Rostova, G.D. Tolstolutska
National Science Center “Kharkov Institute of Physics and Technology”, Kharkiv, Ukraine
E-mail: veg-annie@ukr.net
The constantly growing consumption of electricity requires the development and implementation of more powerful
and energy-intensive systems of the new generation. Fusion and fission reactors of the 4th generation (Gen-IV) will
make it possible to cover the growing demand for electricity. Since Gen-IV reactors will operate at higher temperatures
and radiation doses, the problem of selecting scientifically based structural materials arises, since conventional reactor
materials are not suitable for use in such severe operating conditions. Among the structural materials under
consideration for future generations of reactors, special attention is paid to 9…12% Cr ferritic-martensitic steels due
to their higher radiation tolerance and excellent mechanical properties compared to traditionally used austenitic steels.
This review presents the main ferritic-martensitic steels that will be used as structural materials, their structure,
mechanical properties and various thermal and thermomechanical treatments applied to them.
INTRODUCTION
Most of the industrial nuclear reactors operating in
the world are of the second generation category.
Generation III reactors have just begun to be used, and
Generation III+ reactors are in the commercialization
stage. Although the safety and reliability of these reactors
are very high, it is widely recognized that nuclear power
plays a critical role in meeting the world's ever-increasing
energy needs. In 2000, the Gen-IV initiative was created
– an international alliance between ten countries and the
European Union, and the number of participating
countries is growing [1]. This initiative calls for the
creation of new nuclear power systems that will
significantly improve safety and reliability,
sustainability, reactor useful life (60 years or more) and
profitability, making them stand out from existing
nuclear power reactors.
Six main types of reactors were selected for further
research and development and subsequent
implementation: Gas-cooled fast reactor (GFR), Lead-
cooled fast reactor (LFR), Molten salt reactor (MSR),
Sodium-cooled fast reactor (SFR), Very high
temperature reactor (VHTR), Super critical water-cooled
reactor (SCWR). Operating temperatures for each type of
reactor are: GFR – 850 °C, LFR – 480…800 °C, MSR –
700…800 °C, SFR – 550 °C, VHTR – >900 °C, SCWR
– 510…625 °C [2]. It is important to note that the
majority of operating industrial reactors practically do
not have a coolant temperature above 350 °C. Therefore,
the predicted operating conditions for Gen-IV systems
pose significant challenges in the choice of construction
materials. Structural components will be exposed to
different operating conditions: exposure to higher
temperatures, higher doses of neutrons and extremely
aggressive environments. However, it is also important
to note that one material suitable for one Gen-IV design
may not be suitable for similar applications in other
reactors, depending on the operating conditions of the
particular reactor.
Among the desired characteristics of structural
materials of Gen-IV, the following can be noted: the in-
core materials need to exhibit dimensional stability under
irradiation, whether under stress (irradiation creep or
relaxation) or without stress (swelling, growth); the
mechanical properties of all structural materials (tensile
strength, ductility, creep resistance, fracture toughness,
resilience) have to remain acceptable after ageing; the
materials have to retain their properties in corrosive
environments (reactor coolant or process fluid);
acceptable resistance to radiation damage (radiation
hardening and embrittlement) at high neutron doses
(10…150 dpa), helium embrittlement, etc.; high degree
of chemical compatibility of structural materials and
coolant, as well as fuel. Finally, processability,
weldability, cost, and any other important aspects should
be considered in the material selection process. All these
requirements are related to the fundamental mechanisms
of high-temperature degradation, such as phase
instability, oxidation, radiation-induced segregation, etc.
Ferritic-martensitic (FM) steels are one of the
candidate materials for the future generation of reactors
[3–7]. Before the finding the irradiation-induced void
swelling phenomenon [8] high chromium
(9…12 wt.% Cr) FM steels were used in the
petrochemical industry, and later for gas turbines and
conventional fossil fuel power plant applying [9, 10].
Firstly FM steels were not considered as structural and
cladding materials in the early nuclear reactors and
conventional austenitic stainless steels were used as
structural components in nuclear service, but when it was
understood that swelling behavior is a key parameter for
fast reactor core materials selection [11–13] it became
obvious that a scientifically based selection of new
radiation-resistant materials is needed.
mailto:veg-annie@ukr.net
ISSN 1562-6016. ВАНТ. 2022. №4(140) 67
FM steels, especially containing 9…12 wt.% Cr
(conventional Cr-Mo steels and the reduced activation
ferritic-martensitic (RAFM) steels that differ from Cr-
Mo steels due to W presence in substitution of Mo),
exhibit excellent mechanical properties, thermal
properties, irradiation tolerance, corrosion, and oxidation
resistance [14]. These superior properties make the high
Cr FM steels excellent candidates for Gen-IV reactors
with several concepts proposed including GFR, SCWR,
MSR, liquid metal cooled sodium fast reactors (Na-
LMR) and lead fast reactors (Pb-LMR) [15, 16].
The reference FM steel considered for several reactor
systems is the modified 9Cr-1Mo T91 steel. However,
other FM steels, such as e.g. the Eurofer 97 and T92
steels, will be also included in GETMAT work
programme [17]. Regarding to Europinean conception
[18], FM steels are also considered as possible ATF
materials as a cladding and core components where
swelling must be low.
It was established that FM steels have higher swelling
resistance [19] in comparison with the conventional
austenitic stainless steels and that’s why they were
selected as materials for core components. In addition,
it’s well known that swelling of bcc-materials (Fe-Cr
steels) typically is much lower than widely used
austenitic base alloys. Moreover, swelling of steels with
9% Cr and which have mainly martensitic structure is
lower than in 12% steels, structure of which consisted of
both martensite and ferrite [20–22].
12Cr-1MoVW (HT9 grade) in the USA, unstabilized
9Cr-1Mo (EM10) and duplex 9Cr-2MoVNb (EM12) in
France, 11Cr-MoVNbW (PNCFMS) in Japan, 12Cr-
MoVNb (FV448, FV607, 1.4914) in the UK and
Germany [23–29] were used for the fabrication of fuel
assemblies irradiated in various fast reactors worldwide,
among which EBRII (Experimental Breeder Reactor II)
and the Fast Flux Test Facility (FFTF) in the USA,
Phoenix and SuperPhoenix in France, Prototype Fast
Reactor (PFR) in the UK. Because of their confined
strength and creep efficiency at high temperatures
(around 560 °C [9, 30]), this type of steels were mainly
applied as wrappers (or ducts) in liquid metal cooled fast
reactors where temperature did not exceed about 550 °C.
Though, 12%Cr alloys HT9, FV448 and 9% Cr EM12
were used for cladding. Due to their excellent
dimensional stability under irradiation, the use of FM
steels as core materials in fast neutron reactors has been
successful. Fuel assemblies made from FM steels have
achieved high burnups with corresponding displacement
doses on steels up to about 150…160 dpa [31, 32]. In
particular, EM10 steel was chosen as wrappers material
of Na-cooled fast reactors and for core components of
other types of reactors: in ASTRID (Advanced Sodium
Technological Reactor for Industrial Demonstration)
[33], PNC-FMS steel for JSFR (Japan Sodium-cooled
Fast Reactor) [34], CFR1000 (China Fast Reactor 1000)
[35] and PGSFR (Prototype Generation-IV Sodium-
cooled Fast Reactor) [36]. Another 12% Cr steel
15Cr12MoVWN was proposed for wrapper tubes in
Chinese sodium-cooled fast reactors based on the
requirements of tensile and creep strengths [37].
HT9 steel has been considered for both in-core and
out-of-core applications of fast breeder reactors [38, 39],
and for the first wall and blanket structures of fusion
systems [40, 41]. Moreover, the HT9 steel has been
applied fruitfully in the FFTF as fuel cladding and ducts
[42]. HT9 steel was also considered as cladding material
in LFR systems [43, 44].
9% Cr RAFM steel Eurofer 97 and its ODS versions
is considered mainly for fusion reactors like DEMO as
blanket, divertor cassette, structural material for the
Water Cooled Lead Lithium concept of a fusion reactor
and wall material [45–50]. Some other reduce activation
steels were chosen for fusion systems: F82H and JLF-1
(Japan), OPTIFER Ia and OPTIFER II (Europe) and
ORNL 9Cr-2WVTa (USA) [15].
FM steels are considered primary candidates for SFR
cladding and duct materials of several Gen-IV SFR
designs. The US fast reactor program adopted HT9.
Similar types of steel have been chosen in Europe and
Japan (EM12, DIN 1.4914, and PNC-FMS). In addition,
T91 and T92 have also been chosen for similar
application [7].
Worth a separate mention the modified 9Cr-1Mo steel
with additions of Nb, V, and N (i.e. mod. 9Cr-1Mo or
Grade 91/T91) which was developed in the 1970s [51,
52], in the first instance for SFR steam generator
application and then as sole material for the sodium
transport system [53]. T91 showed higher creep strength
than standard commercial alloys in the range 2.25Cr-
1Mo to 12Cr-1Mo [51] due to presence of fine V, Nb
carbonitrides that arise during normalization and
tempering treatments. From 1980s T91 was used for
boiler tubing applications [54] and in the power and
petrochemical industries [54, 55] and for conventional
power plants that were designed for higher temperatures
than the preceding plants (up to 590 °C) [9]. Additionally
it should be noted that FM steels have smaller thermal
expansion coefficient and a better thermal conductivity
than austenitic ones and this leading to lower thermally
induced stresses, better thermal fatigue properties and
therefore increased component lifetime.
In nuclear power application, T91 has been provided
for the Steam Generator of the European Fast Reactor
(EFR) project. Following the EFR project, T91 has been
recently considered as a candidate for the steam generator
of the ASTRID prototype [56]. Moreover, FM steels are
possible candidate materials for a power conversion
system based on the super-critical CO2 Brayton cycle
[57].
T91 has also been chosen to manufacture the steam
generator in the Prototype Fast Breeder Reactor [58] and
Commercial Fast Breeder Reactors. Also, T91 was
chosen as structural material for the primary and
secondary heat transport system components (piping,
intermediate heat exchanger and steam generator) of
JSFR [59, 60].
LFR are fast spectrum reactors cooled by molten lead
(or lead-based alloys) operating at high temperatures
[61]. Liquid Lead or liquid Lead-Bismuth eutectic (LBE)
were also chosen as coolant for several concepts of
Accelerator Driven Systems (ADS). ADS are subcritical
nuclear systems developed for different applications, in
particular the transmutation of nuclear waste [62, 63].
FM steels have been selected, or considered as possible
candidate materials for structural and core materials of
68 ISSN 1562-6016. ВАНТ. 2022. №4(140)
various LFR and ADS concepts. The materials
performance assessment in heavy liquid metals has
focused on FM and austenitic steels and the experiments
have been conducted mainly in LBE. In particular, the
FM steel T91 was chosen as one of the candidate
materials for internal structural components of future
LFR and ADS, because of its very good behaviour under
irradiation and better corrosion resistance than austenitic
steels. The proton beam window of the MEGAPIE
(MEGAwatt Pilot Experiment) LBE spallation target, the
first liquid metal target was made of T91 [64, 65]. T91
was also considered as wrapper and possibly cladding
material for the lead-cooled ELFR (European Lead Fast
Reactor) and ALFRED (Advanced Lead Fast Reactor
European Demonstrator, a smaller size LFR
demonstrator) [66–68] as well as for the wrapper, core
support plate and spallation target window of MYRHHA
(Multi-purpose hybrid research reactor for high-tech
applications), an accelerator-driven LBE cooled system
designed to operate both in subcritical and critical modes
[69, 70].
The primary reference material for the hot vessel of
reactor pressure vessel option is Grade 91. A substantial
database on the baseline mechanical properties of the
Grade 91 steel is currently available. Sufficient data are
also available on the long-term thermal aging effects on
the mechanical properties for this steel. However,
additional data are needed for the mechanical properties
of thick sections, where there is the possibility of retained
ferrite in this martensitic steel that can lead to
embrittlement. As with the other alloys under
consideration, properties in impure helium must also be
explored. Grade 91 is a relatively mature material, as
indicated by its inclusion in Section III of the ASME
Boiler and Pressure Vessel Code (BPVC), including in
subsection NH on high-temperature materials. Code
qualification applies for operation to 300000 h, whereas
the current design concept of 60 years would require over
420000 h, if operated at 80% efficiency [71].
Gas-cooled systems studied in the framework of Gen-
IV are the GFR and the VHTR. However, European
materials research programmes for the gas cooled
reactors have focussed on the VHTR concept [72–74].
T91 steel is a candidate material for the hot reactor vessel
[75] of the gas-cooled Gen-IV systems, which should
operate at temperatures close to ~ 450 °C on condition if
no separate cooling circuit is provided in the reactor
design. So far, no reactor vessel using 9Cr FM steels has
ever been built.
The concept of SCWR is also considered FM steels
as structural materials. A number of the FM steels was
found to have better stress corrosion cracking resistance
in SCW than other classes of metallic alloys, however
due to their poorer corrosion behavior, they are probably
less suitable for use in a SCW environment [76, 77].
Finally FM steels such as HT9 are foreseen to be used
as wrapper and cladding materials for the Travelling
Wave Reactor (TWR), a sodium cooled “breed and burn”
fast reactor concept [78]. But it is important to note that
TWR have much higher peak doses (~ 600 dpa) in
comparison with Gen-IV systems (~ 200 dpa). This push
towards very high damage levels in revolutionary nuclear
systems has sparked in recent years numerous irradiation
experiments of FM steels to several hundreds of dpa
using ion beams, with the aim to investigate the
microstructural evolutions occurring at very high damage
doses, in terms of swelling, intergranular segregation and
precipitation behaviors [79–82].
A particular class of materials should also be noted –
oxide dispersion strengthened (ODS) FM steels. ODS
FM steels are considered for cladding materials for high
burn-up fast neutron reactor fuels (especially MA956 and
PM 2000). The nanosized dispersoids of yttrium oxide
give these alloys a good creep resistance at high
temperatures [83]. The ODS grades currently developed
in the frame of the SFR or fusion contain 9…12% Cr.
However, these alloys could show some limitations in
terms of internal corrosion (oxide clad reaction) and
temperature (phase transition around 1075 K). Therefore,
ferritic steels with ~ 14% Cr and more could be used up
to 1175 K.
As noted above, FM steels have high-temperature
mechanical characteristics limited to ~ 500 °C (heat
resistance, thermal stability, creep, etc.), while the
operating temperatures of new generation reactors are
much higher [7]. Moreover, in 12% Cr steels, there is a
problem of δ-ferrite formation, which can degrade the
mechanical properties, especially after irradiation. It
should be noted that microstructure, not crystal structure,
is the dominant characteristic that determines swelling of
FM steels [84]. But when using thermomechanical
(TMT) and heat (HT) treatment, these disadvantages can
be avoided. Furthermore, namely structure and
mechanical properties determine the safe and economical
functioning of nuclear systems.
This review discusses the methods of thermal and
thermomechanical treatment, microstructure features and
mechanical characteristics of FM steels that will be used
in future generations reactors.
1. TREATMENT METHODS OF FM STEELS
The standard HT of FM steels is normalization in the
austenite stability region (≥ 1040 °C) and tempering at
temperatures below the AC1 point (> 700 °C) [85–87].
The received state after the standard HT is usually called
“N&T”.
Thermal and thermomechanical treatment of steels
and alloys are used for structure modernization and
improving of mechanical, physical, radiation and other
properties [88–91].
Among the main methods of HT austenization, partial
and short-term tempering, additional quenching, etc. are
noted. The most common TMT methods are ausforming,
rolling (cold and hot), equal channel angular pressing
(ECAP), high pressure torsion, overall forging, etc.
While HT mainly leads to an improvement in the
structural features of materials due to phase
transformations, TMT makes it possible to qualitatively
improve both the structural and mechanical
characteristics of metals and alloys. TMT increases the
strength properties of materials while maintaining a
sufficient level of ductility. Thus, after TMT, the strength
of the material increases both as a result of cold-
hardening formed during plastic deformation and as a
result of quenching. Due to this, during TMT it is
ISSN 1562-6016. ВАНТ. 2022. №4(140) 69
possible to achieve higher hardening than during
conventional quenching.
At the same time, 9% Cr steels are recommended to
be subjected to TMT, mainly to improve the
characteristics of high-temperature strength and stability,
while for 12% Cr FM steels, it is necessary to apply HT
that should be used to prevent the formation of δ-ferrite,
which negatively affects the mechanical characteristics
of this type of steels. But for 9% Cr FM steels, HT is also
provided for the modernization of the structure before the
following TMT.
Moreover, the mechanical properties of FM steels are
critical and important for the practical applications at
elevated temperatures, especially in nuclear application.
Various strengthening approaches have been used to
improve the strength of FM steel, among which
precipitation hardening and strain hardening have much
influence on mechanical properties at elevated
temperatures. Normalizing and tempering temperature
strongly influences the shape and size of second phase
precipitate particles (i.e. M23C6, MX). Additionally,
severe plastic deformation (SPD) technique, which
causes refinement of grain size up to nano-scale, increase
in dislocation density and increase in number of
precipitates with several order of magnitudes. It also
imparts significant hardening of martensite and fine
distribution of second phase particles. These second
phase particles can stabilize the microstructure by
pinning the grain boundaries and restricting the motion
of dislocations to assure high temperature strength. The
main shortcomings of actual 9…12% Cr high-chromium
steels are that the creep resistance is not enough to fulfill
the engineering requirements at temperatures higher than
600 °C and the material undergoes a cyclic softening.
Creep strength at high temperature could be improved by
a microstructural optimization through nano-
precipitation, guided by computational thermodynamics,
and thermomechanical control process optimization [3,
15].
Let us consider data on different modes and methods
of thermal and thermomechanical treatment of FM steels.
In the following sections, the effect of these treatments
on the mechanical properties and structure of 9…12% Cr
steels will be presented.
To study the effect of HT on microstructure and
hardness of Grade 91 steel [92] normalizing was first
done on each sample within a temperature range of
1020…1100 °C for 2, 4, and 8 h. Normalizing
temperature and time were so chosen that austenitization
had been complete before tempering was carried out.
Normalized samples were air cooled (AC) down to room
temperature before tempering at various temperatures
and times. Samples that have been normalized at 1040 °C
for 2, 4, and 8 h were tempered at 690, 725, 745, and
790 °C for 2, 8, and 20 h. Tempering of samples
normalized at 1040 °C for 2 h was expanded to include
temperatures in the range of 635…850 °C. This created a
matrix of close to 50 possible combinations of
normalizing and tempering scenarios, which provided an
adequate sequence to reveal the microstructural changes
during HT.
Heterogeneous grain growth during austenitization in
ASTM A213-T91 steel has been studied in [93]. The
material was received in its standard condition, that is,
normalized at 1060°C and tempered 40 min at 780 °C.
The thermal cycles were carried as follows: heating at a
rate of 1 and 50 °C/s, austenite holding at 1060 and
1080 °C for 1, 15, and 30 min and quenching at 50 °C/s.
The microstructural parameters (dislocation density,
martensite lath width, precipitate diameters and volume
fractions) have been investigated for the 9% Cr steel P92
(NF616) after 2 h at 1070 °C of austenitising and 2 h at
715, 775, 835 °C of tempering (to increase degrees of
martensite recovery and decrease dislocation density),
2 h at 970 and 1145 °C of austenitising and without or
with tempering at 775 °C for 2 h (to increase prior
austenite grain size) and 2 h at 1070 °C of austenitising
with furnace cooling (FC) to 780 °C and subsequent
tempering at 780 °C for 8 h with FC to room temperature
(to receive ferritic matrix with no martensitic
transformation) [94].
Very interesting method of thermal treatment
(namely – partial tempering) is proposed in [95]. This
modified HT method is used instead of TMT to achieve
high mechanical characteristics. All the partially
tempered (PT) samples were processed by normalizing
the as-received (AR) T91 steel at 1000 °C for 15 min
with AC and a subsequent partial tempering treatment at
300, 400, 500, 600, and 700 °C for only 3 min followed
by water cooling. A detailed scheme of thermal treatment
is shown in Fig. 1.
Fig. 1. The heat treatment procedures for T91 steel.
AC – air cooling, WC – water cooling [95]
The influence of HT on structure and tensile
properties of 15Cr12MoVWN (HT9 type) FM steel was
determined in [96]. HT regeimes were as follows:
Austenitizing at 1000, 1050, and 1080 °C, tempering at
650, 700, and 760 °C with next water quenching (WQ),
oil quenching or AC.
In [97] to assess the failure resistance of HT9
specimens were heat treated to twenty-five combinations
of austenitization temperature. Each HT consisted of
austenitizing at one of five temperatures for one hour,
followed by AC and then tempering at one of five
tempering conditions followed by subsequent AC. The
austenitizing temperatures (Tγ) were 950, 1000, 1050,
1100, and 1200 °C. The tempering conditions were
650 °C – 1 h, 650 °C – 56 h, 715 °C – 1 h,
735 °C – 0.5 h, and 780 °C – 1 h; the corresponding
70 ISSN 1562-6016. ВАНТ. 2022. №4(140)
tempering parameters (PT = T(K) (logt(h) + 20) × 10-3)
ranged from 18.5 to 21.
The effects of the HT conditions on the
microstructure and mechanical properties of HT9 steel
were studied in [98]. The HT9 steel was normalized at
1000 to 1100 C for 30 min followed by AC to room
temperature. The tempering treatment of the normalized
specimens was carried out at 700 to 780 C in 1 h
followed by AC to room temperature.
Effects of HTs, with varying normalizing
temperature, temper temperature and cooling rate, on the
microstructure and mechanical properties were studied in
[99]. To investigate the effect of normalizing temperature
on grain size, hardness and grain boundary
misorientation, Eurofer ODS steels were subjected to
solution treatment at 980, 1040, 1100, 1150, 1300, and
1350 °C, followed by cooling and tempering at 700 °C;
T92 and Eurofer 97 steels were subjected to solution
treatment at 980, 1040, 1100, and 1150 °C, followed by
WQ and AC, respectively, and tempering at 700 °C.
Different cooling rates, WQ, AC, and FC are applied to
Eurofer ODS steels to observe the effect on grain
boundary misorientation. Eurofer ODS and T92 were
subjected to solution treatment at 980 °C, followed by
WQ, and tempering at 550, 650, 750, and 850 °C.
Regarding TMT – in [100] the Eurofer 97 plate was
hot deformed using two reheating conditions (1075 and
1175 °C). Two finish rolling temperatures (750 and
650 °C) and two different total reductions (30 and 40%)
are also analyzed. The tempering effect is also studied (in
the temperature range 720…760 °C).
Another study of the effect of ausforming of Eurofer
97 steel on structure and mechanical properties described
in details in [101]. Here the TMT consisted of a solution
and austenitization treatment at 1250 °C in air followed
by cooling in a second furnace. The temperature of the
second furnace was set to the desired rolling temperature
(namely 600, 700, 800, and 900 °C).
To study microstructural and mechanical
characteristics of Eurofer 97 specimens in the AR
condition were ECAP processed at 550 °С through a die
with an intersection angle of 105° [102]. The billet was
rotated around its longitudinal axis before inserting in the
die for the following ECAP pass, either 180° (route C) or
+90° and -90° alternatively (route BA). The billets were
subjected to 1, 2, and 4 ECAP passes following route C,
as well as 8 passes following route BA.
Guidelines for tailoring the microstructure and
mechanical properties of T91 steel via ECAE (equal
channel angular extrusion) enabled TMT for an improved
combination of strength and ductility are presented in
[103]. TMT processes include: (1) austenitizing at
800…1200 °C (with an interval of 100 °C) for 1 h
followed by quenching in ice water; (2) ECAE process of
austenite at 1000 °C followed by WQ, and then
tempering the thus- deformed T91 steel for 1 h at
300…800 °C; (3) tempering the water quenched
specimens at 500 °C for 10 h, followed by ECAE of the
tempered specimens. ECAE was carried out with number
of extrusions from 1 to 3 at room temperature, 300 and
700 °С.
Effect of normalizing (950…1050 ºC) and tempering
(650…700 ºC) has been investigated to study the
microstructure and mechanical properties of modified
9Cr-1Mo steel (G91) [104]. To improve the high
temperature mechanical properties, hot rolling has been
done at optimized normalizing and tempering conditions.
The hot rolling is carried out in three passes at 1050 and
550 ºC.
In [105] to refine the 9Cr-1Mo martensitic steel (T91)
it was processed by ECAP method at room temperature
and annealed at 500 °C for 2 h subsequently. ECAP
processing was carried out in a self-designed die with the
intersecting angle of 120° with an amount of extrusions
from 1 to 6.
A brand new method of deformation of 9…12% Cr
steels is proposed in [106]. The microstructure and
mechanical properties of the FM steel T91 after SPD and
subsequent thermal treatment was investigated. SPD was
realized by the original method, developed in NSC KIPT
of multiple “upsetting-extrusion” (MUE) [107], in
temperature range of stability of ferritic phase
(750…575 °C), thermal treatment of SPD specimens was
carried out at temperatures 550…730 °C during
1…100 h.
Effects of ausforming through upsetting and
following tempering on microstructure and mechanical
properties under uniaxial tension of 12CrWMoNbVB
FM steel are studied in [108]. TMT included heating in
air to ~ 890 °C, holding at this temperature for 15 min,
and following upsetting (i.e. ausforming). After that the
samples from the ausformed steel were tempered in an
argon atmosphere at three modes: 720 °C for 3 h; 665 °C
for 3 h; 550 °С for 25 h, with further cooling of the
samples in the air.
2. MICROSTRUCTURE OF FM STEELS
After standard HT, all 9…12% Cr FM steels have a
tempered martensite structure. It necessarily contains
prior austenite grain boundaries (PAGB) decorated with
carbides of the type М23С6 (where M – Cr, Fe, W). In the
body of the prior austenite grains (PAG), carbides and/or
carbonitrides of the MX type are uniformly distributed
(where M – V, Nb, Ta and X – C, N). Generally, the sizes
of М23С6 precipitates are 60…200 nm, while МХ –
≤ 50 nm [109–113]. The mean grain size according to
[86, 96, 114] is ~ 20 µm. But if the Eurofer 97 and T91
steels are supplied as a hot rolled plates, they present a
fine structure with a PAG size in the ASTM range
10…11.5 (6.7…11 µm) for the 14 mm plate and ASTM
10.5…11 (8…9.4 µm) for the 25 mm thickness plate [92,
115, 116].
The typical view of 9…12% Cr steels structure is
presented on Fig. 2. It can be seen that 9% Cr steels have
a pure martensitic structure, while 12% Cr contains δ-
ferrite in the form of white islets and streaks. Because of
its higher chromium (a ferrite stabilizer) content, HT9
can contain 0…2% δ-ferrite, depending on composition
variations within the alloy specifications (see Fig. 2,c). It
is generally believed that the effect of δ-ferrite in
9…12% Cr FM steel on properties depends on its volume
fraction. Hu et al. [117] and Hu [118] show that, in the
steel E911, δ-ferrite has little effect on the strength of
steel. When the volume fraction of δ-ferrite is smaller
than 0.3%, it can slightly increase the strength of steel. A
small amount of δ-ferrite (less than 1%) does not affect
ISSN 1562-6016. ВАНТ. 2022. №4(140) 71
the properties of steel, and can even improve the ductility
and toughness of steel to a certain extent. Only when it is
over 1% is a bigger impact produced in the fracture
toughness of the steel.
Also, the ferrite phase inhibits austenite grain growth,
but it adversely influences the strength and, directly or
indirectly, the toughness [119–121], particularly if
present as films between the grains of the austenite,
which is subsequently transformed to martensite and
tempered. That’s why steels with 9% Cr were favored
over those with 12% Cr because of the difficulty of
eliminating δ-ferrite in a 12% Cr steel without increasing
carbon or manganese for austenite stabilization and
specific heat treatment, but manganese promotes chi-
phase precipitation during irradiation,
which can cause embrittlement [15].
Fig. 2. Microstructure of initial Т91 (a); Eurofer 97 (b) [122], and НТ9 (c) [84] steels
After HT and TMT, the structure of 9…12% Cr steels
changing. Thus, after normalizing time of 2 h, PAG size
and martensite lath size increased with increasing
normalizing temperatures of 1020…1100 °C for steel
T91. The as-normalized (1040 °C for 2 h) microstructure
had hard martensitic lath structures with high dislocation
density and for constant normalizing temperature of
1040 °C, martensite lath size and PAG size increased
with increasing normalizing time of 2, 4, and 8 h. Grain
size increased with increasing normalizing time, and
tempering time. The sample normalized at 1040 °C for
2 h and tempered at 790 °C for 2 h has grain size of
~ 11 μm, while sample normalized at 1040 °C for 8 h,
and tempered at 790 °C for 20 h resulted in a grain size
of ~ 17 μm. The grain size of initial steel was ~ 7 µm
[92]. The obtained structures are given on Fig. 3.
Fig. 3. Microstructure of T91 steel: a – initial; b – normalized at 1040 °C for 8 h
and tempered at 790 °C for 20 h [92]
Heterogeneous grain growth during austenitization in
T91 steel has been investigated in [93]. Starting from a
uniform, fine austenite grain size distribution after 1 min
of holding time, a heterogeneous austenite grain size
distribution was observed after 15 min of holding time at
1060 and 1080 °C and it was relatively insensitive to the
change in the heating rate. This heterogeneous growth is
probably related to the evolution of MX precipitates
during the austenite treatment. For the 1 min of
annealing, precipitates were found to be Nb-rich MX
(major) and V-rich MX (minor), these last with sizes
higher than 50 nm. Very few M23C6 particles were also
detected. For the 15` and 30` of annealing times V-rich
precipitates were not found, and the chemical
composition of the Nb-rich precipitates displayed values
slightly higher than those measured in the AR state. M3C
precipitation was not observed to be related to MX
precipitation; on the contrary, zones with MX presence
and no M3C particles were marked. On the other hand,
M3C precipitates are mainly found in coarse martensite
laths, a similar observation has already been reported in
[123] for the T92 steel.
Some interesting results were obtained in [94]. Thus,
increasing of tempering temperature for P92 steel leads
to enhanced martensite recovery processes, which
decrease the dislocation density. These processes were
accelerated at higher tempering temperatures, so that
tempering at 715 °C led to a slightly higher dislocation
density than the standard tempering at 775 °C.
Tempering at 835 °C caused a sharp decrease in the
dislocation density of about 75%. With increasing
austenitising temperature the austenite grain size and the
martensite lath width increase (lath – from 0.38 nm at
970 °C to 0.42 nm at 1070 °C, and 0.58 nm at 1145 °C;
a b
72 ISSN 1562-6016. ВАНТ. 2022. №4(140)
grain size – from 10 µm at 970 °C to 20 µm at 1070 °C,
and 60 µm at 1145 °C). The dimensions of the larger
M23C6 and the fine MX precipitates were similar for all
three HT conditions.
The untypical, for initial state, phases are observed in
T91 steel structure after partially tempering [95]. The AR
sample is shows a typical fully tempered martensite (α’T)
microstructure with a large number of coarsened
precipitates distributed along PAGBs and martensitic
lath boundaries, indicating a significant release of carbon
atoms from the matrix. The microstructure of 1000WQ
sample (normalizing at 1000 °C for 1 h, followed by WQ)
consisting of a majority of quenched martensite (α’) and
a small amount of auto-tempered martensite (α’AT)
formed during the cooling process. Relatively clean
(precipitate-free) PAGBs and lath boundaries indicate
that most carbon atoms are dissolved into the matrix
forming supersaturated quenched martensite. A certain
amount of coalesced bainitic ferrite (αB) (labelled by
yellow dash line on Fig. 4) and partially tempered
martensite (α’PT) were discovered in the PT T91 samples,
as shown in Fig. 4. The microstructure of the PT T91 steel
consists of partially tempered martensite and bainitic
ferrite with transition carbides and ultra-fine martensite.
The transition carbides within the bainitic ferrite grains
in the PT samples are identified as ε- and θ-carbides, and
ε-θ-carbide complexes.
The microstructures of 15Cr12MoVWN steel after
austenitizing at 1000, 1050, and 1080 °C, tempering at
650, 700, and 760 °C with next WQ, oil quenching or AC
is investigated in [96]. It is observed that
15Cr12MoVWN steel has a fully martensitic structure
without formation of δ-ferrite after all HT regimes. The
average grain size varies little with the different cooling
methods and tempering temperatures. On the contrary,
the PAG size changes obviously with different
normalizing temperatures, especially for normalizing at
1080 °C. The higher the normalizing temperature, the
greater the grain size. The normalizing temperature is
therefore the most important factor affecting the PAG
size among the influencing factors in the orthogonal
design experiment.
Fine martensitic laths dominated in the specimens
tempered at 650 and 700 °C irrespective of the
normalizing temperature and cooling method. When the
tempering temperature increased to 760 °C, the
dislocation recovery accelerated, and the dislocations in
the martensitic lath passed through the precipitates by
climbing or bypassing to form a dislocation wall, which
promoted the formation of subgrains and accelerated the
crushing of the lath. The increasing tendency of the
martensitic lath width is low from 650 to 700 °C, but
significant at 760 °C during tempering. The lath width
increased with increasing grain size and normalizing
temperature. The tempering temperature was more
effective in adjusting the lath width than the normalizing
temperature. The present results of lath width variation in
the 15Cr12MoVWN steel specimens are in good
agreement with the literature [124].
Two types of precipitates were observed in
15Cr12MoVWN steel. The elongated precipitates M23C6
distributed in the chain along the grain boundaries,
whereas fine spheroid particles MX distribute
predominantly in the laths.
The HT9 cladding tube has a typical martensite
structure, and the PAG size increased with an increase of
normalizing temperature [98]. The microstructure of the
normalized specimens consisted of lath martensite with a
high dislocation density due to austenite to martensite
transformation. The tempered specimens showed a
tempered lath martensite structure with a low dislocation
density because of the dislocation recovery. The lath
width is not greatly affected by the tempering
temperature change. A precipitate, unlike a lath structure,
varies depending on the tempering temperature.
Fig. 4. SEM micrographs of the AR, WQ, and PT T91
specimens heat treated at 300…600 °C:
a – AR T91 steel; b – WQ sample;
c–f – PT samples [95]
Fig. 5 shows the precipitates and PAG size, as well as
the lath width according to the tempering parameter (i.e.
Hollomon-Jaffe parameter). The PAG size indicates a
similar value by increasing the tempering parameter. The
PAG size was not affected by the tempering parameter.
It is known that PAG size varies with the normalizing
temperature. The lath width showed similar values by
increasing the tempering parameter. The lath width was
not affected by the tempering parameter. However, the
size of the precipitate was increased by increasing the
tempering parameter. The precipitate behavior was
affected by the tempering parameter.
ISSN 1562-6016. ВАНТ. 2022. №4(140) 73
Fig. 5. Tempering parameter with the lath width, prior
austenite grain size and precipitation mean size
at 1050 °C [98]
Effect of HT on microstructure of Eurofer 97 steel,
two Eurofer ODS variants (HXX and FZK, the latter
having a very low carbon concentration) and T92 (or
NF616) is studied in [99]. Thus, for T92 and Eurofer 97
steels, the grain size increases with the increase of
solution temperature. The grain coarsening rate of T92 is
faster than that of Eurofer 97. This may be because there
is higher Ta concentration in Eurofer 97 than Nb in T92
(both Ta and Nb play a role of grain refinement). In
Eurofer ODS steels (HXX and FZK), the grain size
remains almost unchanged up to 1300 °C. After that,
grain size increases with the increase of normalizing
temperature. This indicates that the pinning of grain
boundaries by yttria (Y2O3) particles is effective below
1300 °C. The dependence of PAG size on normalizing
temperature is shown in Fig. 6.
Fig. 6. Prior grain size vs. normalizing
temperature [99]
A strong effect of reheating temperature before TMT
on austenite grain growth is detected [100]. Hence, after
heating at 1075 °C, subsequent hot rolling at 750 °C (hot
reduction was 40%) and next tempering at 720 °C the
structure of the Eurofer 97 sample is similar to tempered
martensite with fine prior austenite grains (Fig. 7,a). But
after heating at 1175 °C, hot rolling at 650 °C (hot
reduction was 40%) and tempering at 720 °C the
dramatically grain growth is observed and the form of the
structure changes significantly (see Fig. 7,b).
Another variation of ausforming of Eurofer 97 was
performed in [101]. The microstructural investigations
showed large and elongated primary boundaries. Neither
of the materials exhibited a full martensitic structure.
Despite the annealing at 750 °C for 2 h, a high density of
dislocations still observed. With increasing of rolling
temperature, the martensite laths and secondary phases
were refined. After rolling at 900 °C a larger amount of
fine precipitates (10…20 nm diameter) inside the
subgrains was observed. Also a reduction of the particle
agglomeration with increasing of TMT temperature was
detected. Agglomerations of fine MX precipitates were
only observed in the materials rolled at 600 °C. Coarse
M23C6 carbides a representinall material due to the
carbon content of 0.1 wt.%. Their size and shape were
not affected by the rolling conditions. M23C6 particles
with a typical sizes range between 100 and 300 nm were
measured in all cases.
Fig. 7. Effect of reheating temperature on
microstructure of Eurofer 97 steel: a – after reheating
at 1075 °C; b – after reheating at 1175 °C [100]
The effect of ECAP on the microstructure of
tempered Eurofer was studied in [102]. The
microstructure of the AR samples exhibited traces of the
PAGs and martensite laths. Furthermore, carbide
particles were found homogeneously distributed in the
matrix, as well as located at PAGBs and along the lath
boundaries. After the first ECAP pass the microstructure
exhibited large elongated grains fragmented into
subgrains, as expected. This fragmentation may be due to
dislocation arrangement into walls and cells induced by
the dynamic recovery mechanism during the ECAP
process at 823 K. Note that some subgrains are separated
by a dislocation wall and other subgrains seem to have
developed from a dislocation cell. After the fourth pass,
the microstructure was no longer elongated, martensite
laths were not observed and an equiaxed structure of
a
b
74 ISSN 1562-6016. ВАНТ. 2022. №4(140)
submicron grains appeared, but still the number low
angle grain boundaries was abundant. Eight passes via
route BA generated a more refined structure, with grain
sizes less ~ 500 nm and a high number of high-angle
grain boundaries.
TMT by ECAP method of T91 steel was carried out
in [103]. The as-annealed (so-called AA, annealing of
AR material at 800 °C for 1 h, followed by FC) T91 steel
has typical tempered martensitic lath structure (α`T) with
minor second phases, including retained austenite or
martensite (M/A constituent), and/or transformed
carbides. The fully tempered T91 is mainly composed of
ferrite phase with packet and block boundaries. PAGBs
and packet boundaries are often decorated by “white”
carbide particles. Tempered martensite laths in the AA
specimen dominate the microstructures inside a packet.
WQ of T91 from 800 °C (specimen called 800WQ)
leads to fine microstructure and it is shown that this fine
structure is a mixture of grain boundary ferrite (αA),
martensite (α`M), and polygonal ferrite (αP). Some
martensitic blocks within PAGs were etched. A PAG in
the 800WQ specimen is surrounded by adjacent αA
containing undissolved carbides. The ferrite phase (αP)
and fine martensite are occasionally separated by a
packet boundary.
The microstructures of the water quenched 900 and
1000 °C specimens (900WQ and 1000WQ, respectively)
have primarily martensite α`M and polygonal ferrite (αP),
αA is absent in the 900WQ specimens. PAGBs are also
observed in the 900WQ specimen and a triple junction of
PAGs in the 900WQ T91 is observed. Retained austenite
γ is identified as the thin layers between martensite laths
with dark contrast. The αP ferrite with martensite
precipitates is also observed in the 900WQ specimen.
Significant coarsening of austenite grains occurs after
HT at 1100 °C or higher prior to WQ as was detected in
the 1100WQ and 1200WQ specimens. Blocks are clearly
observed in these two specimens. Except for packets
around the grain boundaries, no packet boundaries are
identified due to the randomly oriented blocks. Bainitic
ferrite (αB) within the prior austenite grains in the
1100WQ and 1200WQ specimens was preferentially
etched away. Auto-tempered martensite (α`AT) is
frequently observed in the 1200WQ specimen. Carbides
formed within the martensite laths.
The grain sizes of the PAGs in the WQ T91 steel are
similar ~ 13 µm in the AA, 900WQ and 1000WQ
specimens. For specimens heated to 1100 °C or greater
(1100WQ and 1200WQ), the average grain size
increased by an order of magnitude from ~ 13 to 150 µm.
Carbides in auto-tempered martensite and bainitic
ferrite were also studied. Black needle-shape precipitates
are identified in the specimen 1200WQ as η-carbide and
cementite. Precipitates in bainitic ferrite align along
nearly the same direction, in comparison to randomly
oriented precipitates in auto-tempered martensite in the
1200WQ material and were identified to be ε-carbide.
It can be observed that such phases as auto-tempered
martensite, bainitic ferrite etc. formed in 9% Cr steels
after WQ [95, 103]. However, the effect of these phases
on radiation resistance, mechanical properties after long-
term servicing in environments with elevated
temperatures, corrosion and erosion resistance is almost
not studied. That is why a more detailed study of the
properties of FM steels after this type of quenching is
required.
The influence of HT and TMT by hot rolling on
dislocation density, grain and precipitates size of T91
steel studied in details in [104]. The average packet size
of initial steel is found ~ 15 μm, which is reduced to 5 μm
in a sample, normalized at 950 °C and tempered
subsequently. This indicates that the packet size is
strongly dependent on the normalizing temperature. Hot
rolling with 27% deformation causes reduction in the
packet size up to 10.7 μm.
The variation of particles size at the grain boundary
and grain interior under various HT and hot rolled
conditions was presented. In AR state, the average size
of particles at the grain and within grain boundary is
found to be (250 ± 34) and (180 ± 20) nm, respectively.
The particle size is observed to decrease with the increase
in normalizing temperature (950…1050 ºC). Particle size
increases at the PAGBs and within grains with the
increase in tempering temperature. In order to optimize
the size of precipitate particles to achieve high strength
and ductility, tempering is performed between 650 to
750 ºC. There is a decrease in precipitate size with the
low tempering temperature, which results into high
stability of lath boundaries by M23C6 particles at elevated
temperatures and less dislocation recovery. The particle
size is found to be minimum for R550-T650 (where R –
rolling, T – tempering), which is (95 ± 20) nm along
grain boundary and (77 ± 19) nm within grain. The small
particle size provides better pinning effect and hinders
the movement of dislocations, which leads to an increase
in the mechanical strength at room temperature and
650 ºC. At elevated temperature, high rate of self-
diffusion of iron atoms and gliding of dislocations cause
deformation and vacancies formation. In order to prevent
the movement of dislocations, sufficient number of
precipitates are required for the improvement in high
temperature mechanical strength.
After normalization at 1050 °C and tempering at
750 °C the structure is typical tempered martensite with
martensitic lath of 100…400 nm width, where M23C6
particles decorates the lath boundaries and have average
diameter of about 150…300 nm. Additionally, average
diameter of fine carbides is observed as 15…40 nm
within the matrix. If normalized at 1050 °C and tempered
at 750, 700, 650 °C the precipitate size changes
significantly with the decrease in tempering temperature.
Additionally, lath width and precipitates (M23C6/MX)
increases with the increase in tempering temperature.
Similar behavior of precipitate size can be observed
after hot rolling at 550 °C and tempering at 650 °C (i.e.
R550-T650). It shows the formation of dislocation cells
and decrease in lath size (~ 150…300 nm) as compared
to other heat-treated conditions. High dislocation activity
causes formation of dislocation forests. The higher extent
of decrease in precipitate size and lath width significantly
improved the mechanical properties at room temperature
and 650 °C. The precipitate encircled in R550-T650
condition confirms the Cr-rich M23C6 carbides.
The microstructure of AR T91 steel and after ECAP
is investigated in [105]. Compared with the initial
material (average grain size ~ 20 µm), after one extrusion
ISSN 1562-6016. ВАНТ. 2022. №4(140) 75
pass, martensitic laths are destroyed and grains are
broken. The grain elongation gradually becomes evident
and the grain boundaries become more and more hazy
until invisible with the increase of extrusion passes. The
results indicate that due to plastic deformation the
microstructure is severely distorted.
The microstructure of the initial sample contains
arrays of martensitic laths with a width of ~ 0.4 µm and
there are a certain amount of dislocations distributed in
the matrix. Furthermore, some second-phase particles are
found to disperse in the martensite borders and grain
boundaries. For the sample after one pass, it is difficult
to observe complete martensitic laths. However, many
deformation bands in the martensitic laths, as well as lots
of dislocations can be observed, which may be regarded
as the formation of dislocation cells [125–128]. With the
increase of extrusion passes, a number of dislocations
will be introduced and pile up and eventually form new
low-angle boundaries. The equiaxed and refined grains
are formed after six passes, where most of the grain
boundaries are of high angle type. In addition, there are
high density dislocations in the grains and get tangled
together in different places. After six passes the grain size
is distributed in the range of 0…700 nm with the mean
grain size of about 200 nm. So it can be concluded that
the SPD such as ECAP can effectively refine the grains
of T91 steel down to the nanometer scales.
To evaluate the thermal stability of the T91 steel
refined by ECAP processing, the annealing was
performed at 500 °C for 2 h for the ECAPed samples. The
average grain sizes of the ECAPed samples after
annealing are slightly larger than those before the
annealing. For example, after annealing the average grain
size for the sample after six ECAP passes has slightly
grown up from about 200 nm to about 300 nm. In
addition, the dislocation density in the samples after
annealing is much lower than in the samples before the
annealing. Therefore, the annealing at 500 °C mainly
decreases the density of dislocations and results in very
little grain growth.
The investigation of the microstructure after different
number of “upsetting-extrusion” (UE) cycles showed that
it is similar – well-formed grained microstructure with
contrasting boundaries [106]. Increase in the number of
cycles and decrease of UE temperature leads to the
considerable decrease in subgrain size (from 5 μm to
145 nm) and to the improving of their size distribution
uniformity. Decrease in mean subgrain size d with the
increase in number of UE cycles occurs irregularly.
Especially significant microstructure refinement is
observed during deformation up to 3 cycles (e = 4.5).
Herewith d decreases more than 10 times – from 5 μm in
AR state to ~ 475 nm. With a subsequent increase in
number of cycles from 3 to 5 (total deformation e = 8.0)
value of d decreases approximately in 3 times – from
~ 475 to ~ 145 nm. Studies of precipitates size and
density of deformed samples indicated that M23C6
precipitates density is practically independent on thermal
treatment temperature and their mean size increases a few
with the temperature increase, which is in agreement with
the literature [92, 129]. Herewith density of MX
precipitates is maximal (2.0·1021 1/m3) and size of
precipitates is minimal (11 nm) after thermal treatment at
550 C during 25 h.
Finally, after upsetting of a 12CrWMoNbVB steel at
temperature of 890 °C with the true deformation of 0.7, a
banding microstructure with a pronounced direction of
deformation and a transverse size of subgrains of 260 nm
is observed (vs. 220 nm in initial state) [108]. In some
areas, there is no clearly pronounced direction, the grains
are oriented chaotically. After tempering deformed
specimens at 720, 665, and 550 °C a subgrain structure
with an initial stage of recrystallization is observed,
when, along with the formed larger grains with clear
boundaries, practically without dislocations in the grain
field, there are also subgrains of the initial state 5–7 times
smaller in size. There are noticeably more areas with a
smaller subgrain after tempering at 550 °С for 25 h than
after tempering at 720 °С for 3 h. Regarding density of
M23C6 carbides it was found that it varies slightly in all
modes of HT and their size is in the range of
100…140 nm in initial state and after tempering at 720
and 665 °С, but reach 86 nm after tempering at 550 °С.
3. MECHANICAL PROPERTIES
OF FM STEELS
As mentioned in the previous section, various types
of HT and TMT lead to different changes in the
microstructure of materials. The grain size, the present
phases and the types of precipitates will affect the
mechanical properties of the steels.
The main parameters of mechanical characteristics
are yield strength (σ0.2), ultimate tensile stress (UTS or
σВ), ultimate elongation (δ) and microhardness (HV).
Below we consider the effect of HT and TMT on
conventional and reduced activation FM steels.
Effect of normalizing treatment on Grade 91 hardness
is studied in [92]. Hardness measurements were
performed on the alloy samples normalized at 1020,
1040, 1050, 1060, 1080, and 1100 °C while keeping the
time constant at 2 h. The hardness decreased with
increasing normalizing temperature from 4000 MPa at
1020 °C to ~ 3500 MPa at 1100 °C. The hardness value
was about the same for normalizing carried out at 1020
and 1040 °C, but after that, the hardness gradually
decreased until reaching 1100 °C. The effect of change
in normalizing time was studied by heat treating Grade
91 steel at 1040 °C for 2, 4, and 8 h. The hardness
decreases from 4000 MPa after 2 h to ~ 3400 MPa after
8 h.
Alloy normalized at 1040 °C for 2 h was tempered at
690, 725, 745, and 790 °C for 2, 8, and 20 h. Hardness of
the alloy decreased with increasing tempering
temperature and time. Alloy tempered at 690 and 725 °C
showed significant drop in hardness with increasing
tempering time, but for 745 and 790 °C the decrease in
hardness was gradual. Similar profiles were observed for
samples normalized at 1040 °C for 4 h then tempered,
and samples normalized at 1040 °C for 8 h then tempered
at aforementioned temperatures. Among all samples
normalized at 1040 °C and tempered, samples
normalized for 2 h and tempered at 690 °C for 2 h had
the highest hardness, while sample normalized for 8 h
and tempered at 790 °C for 20 h had the lowest hardness.
The tempering temperature of the sample normalized at
76 ISSN 1562-6016. ВАНТ. 2022. №4(140)
1040 °C for 2 h was expanded to temperatures slightly
above AC1 temperatures. Hardness of the alloy
normalized at 1040 °C for 2 h and tempered for 2 h
decreased with increasing tempering temperature up to
745 °C, stabilized till 820 °C.
The σ0.2, σB, and δ at room temperature, 600 and
650 °C for P92 steel in different heat treated conditions
are determined in [94]. Except for the specimen given the
usual austenization treatment (2 h at 1070 °C) but a low
temperature tempering treatment (2 h at 715 °C), the
yield and tensile strengths for different HTs were similar
at each of the test temperatures. The low temperature
tempering led to an increase in strength, especially at
room temperature; this is attributed to the somewhat
higher dislocation density. It should be noted that the
minimum yield stress and tensile strength measured at
room temperature (440 and 750 MPa, respectively)
specified for P92 were met by all the HTs applied.
The high values of yield and tensile strengths are
reached after partial tempering at 500 °C (500PT): 1370
and 1506 MPa, respectively [95]. After mentioned
treatment the microstructure of T91 steel is dual and
consists of 15% of bainitic ferrite (αB) and 85% of partial
tempered martensite (α`PT). Also, the 500PT specimen
has the highest number of ε and θ carbides density
~ 2.32·1013 m-2. The best uniform elongation (8.2%) was
established in AR condition. In fact, the strengthening
mechanisms in the PT T91 are mainly attributed to
precipitation hardening from transition carbides, solid
solution strengthening from solute carbon atoms, and
subgrain boundary strengthening from martensitic lath
and ultra-fine martensite. The large uniform elongation
arises from bainitic ferrite induced by a reduced cooling
rate.
The martensitic lath width and dislocations were the
main microstructural factors influencing the tensile
strength of 15Cr12MoVWN steel [96]. The
strengthening contribution from M23C6 and MX was
higher than that from grain boundaries of PAG and was
the other important factor for strengthening. The average
sizes of PAG and M23C6 particles were the main factors
influencing the tensile ductility. Tempering temperature
had the most significant influence on the evolution of
precipitates and dislocation densities. Normalizing
temperature had the most significant influence on the
coarsening of PAG and M23C6. Cooling methods had less
influence in both microstructure and tensile properties
compared with the other two HT parameters. Also, based
on the tensile strength and elongation at 23 and 550 °C,
the optimized HT parameter was determined:
normalizing at 1050 °C, AC to room temperature and
tempering at 700 °C. The tensile properties of specimens
in optimized HT condition were 1014 MPa (σB),
810 MPa (σ0.2), and 18.8% (elongation) tested at 23 °C,
and the values were 577.5 MPa (σB), 469 MPa (σ0.2), and
39.8% (elongation) tested at 550 °C.
The mechanical properties of HT9 steel specimens
heat treated to twenty-five combinations consisting of
five values of austenitization temperature Tγ
(950…1200 °C) and five tempering parameters PT
(18.5 < PT < 21) were investigated in [97]. Hardness and
strength parameters showed a slight maximum (35 HRC
and 1250 MPa, respectively) and dynamic (lower shelf)
toughness a slight minimum at Tγ = 1050 °C, but in
general these properties were relatively insensitive to Tγ.
This is despite large changes in the PAG size and lath
packet size with increasing Tγ. With increasing PT the
hardness and yield strength decreased, and the dynamic
(lower shelf) toughness increased slightly. The observed
variation in yield strength is most likely the result of
primary dependence on dislocation density, matrix
carbide dispersion and interstitial solute (carbon) content;
these parameters are little affected by Tγ, but are reduced
with increasing PT.
Fig. 8 shows the hardness change of HT9 with the HT
conditions obtained in [98]. The hardness was shown to
increase with increase in normalizing temperature. A
tensile test was carried out at room temperature and high
temperature (650 ℃) to evaluate the tensile properties of
the heat treated specimen under the same conditions. The
tensile test was carried out at room temperature and high
temperature. The yield stress shows a tendency to
decrease proportionally. Yield stress was shown to
increase with an increase in normalizing temperature.
These tendencies were also represented at high
temperature.
Fig. 8. Hardness change of HT9 with HT condition [98]
For T92 and Eurofer 97 steels, hardness increases
with the increase of normalizing temperature [99]. From
Fig. 6, we know that the grain size of T92 and Eurofer 97
increase with the increase of normalizing temperature.
According to the Hall-Petch relationship, the hardness
should decrease with the increase of grain size.
Therefore, this indicates that there is another contribution
to the hardness in T92 and Eurofer 97. The hardness of
Eurofer ODS steels decreases with increased normalizing
temperature. ODS FZK steel shows lower hardness than
ODS HXX, possibly due to its very low carbon
concentration. For T92, hardness decreases with
increasing of tempering temperature. For Eurofer ODS
steels, the hardness decreases with increase of tempering
temperature from 550 to 750 °C, then increases at
850 °C. This suggests that some reversion to martensite
takes place above 750 °C in Eurofer ODS. The HXX
steel shows higher hardness than FZK. With the increase
of normalizing temperature, the grain size increases in
T92 and Eurofer 97 steels and remains almost unchanged
in Eurofer ODS steels, the hardness increases in T92 and
Eurofer 97 and decreases in Eurofer ODS steels. The
formation of Nb- and Ta-rich carbides is related to the
increase of hardness. Ta enrichment in yittria particles is
ISSN 1562-6016. ВАНТ. 2022. №4(140) 77
observed, which may be the reason for the lack of Ta-rich
carbides and resulting reduction in hardness with
normalizing temperature in Eurofer ODS steels. With the
increase of tempering temperature, the hardness
decreases in T92 while in Eurofer ODS steels, the
hardness decreases until tempering temperature up to
750 °C, then increases slightly.
Higher hardness results are shown following higher
temperature re-heating at 1175 °C (HV10 = 2840…3060 MPa)
[100]. This is achievable following a hardenability
improvement after austenite grain size growth. A strong
effect of reheating temperature on austenite grain growth
is detected. Tensile tests have been performed on at
T = 20, 550, and 650 °C. It was established that σ0.2 and
σB after re-heating at 1175 °C + rolling at 650 °C +
tempering at 720 °C are higher than that after re-heating
at 1075 °C + rolling at 750 °C + tempering at 720 °C in
the case of testing at room temperature and 650 °C, but
vice versa at testing temperature 550 °C. Furthermore, at
all 3 testing temperatures the elongation is better after
reheating at 1075 °C. It seems that such behaviour is
connected with a big difference in structures of
specimens (see Fig. 7).
The tensile tests of the materials showed only minor
variations between the different materials [101]. While
the material treated at 900 °C it shows lower strength at
room temperature, in the range of the operation window
(550…700 °C) within the error range, no differences can
be observed. The TMT brought a general increase in
strength to the materials. The yield strength is
approximately 50 MPa higher than the AR condition
throughout the whole tested temperature range
Elongations of the TMT alloys are orders of magnitude
lower (0.5…2% vs. 10%) than in the initial state under
the same creep conditions. Since none of the specimens
have failed so far, creep-to-rupture times are expected to
be greatly prolonged.
ECAP processing enhances both yield strength and
ultimate tensile stress up to temperatures around 773 K
[102]. Above this temperature, which is near the
processing temperature, the material processed four
times via route C or eight times via BA became softer than
the tempered material and the materials processed by 1
pass, or 2 passes via route C. The effect of ECAP on the
temperature dependence of the hardening ratio
(UTS/yield strength) is studied. In contrast to the
expected result for steels, and other metals, processed by
SPD, which do not exhibit significant work hardening
during tensile testing, Eurofer processed by ECAP does.
The hardening ratio for ECAP processed Eurofer was
somewhat enhanced compared with the initial Eurofer in
the investigated temperature range, except for the
materials ECAP processed by 8 passes via route BA that
had a lower hardening ratio at room temperature. It is
should be noted that the steep increase in the hardening
ratio for the materials processed by 4 or 8 passes is
accompanied by an abrupt change in the temperature
trend of their uniform and fracture elongations. However,
tensile behavior of materials processed by 1 pass, or 2
passes via route C, is very similar to that for the AR
material.
Effect of the preliminary HT and ECAP on
mechanical properties of T91 steel is described in [103].
In general, the WQ T91 steels show a combination of
high strength and ductility. The yield strength of 800WQ
material is 970 MPa, ~ 150 MPa lower than those heat
treated above 900 °C. The uniform elongation is ~ 7% for
all of the WQ T91 steel, which is insensitive to
austenitizing temperature prior to WQ.
The hardness of AA specimen changes slightly and
remains at ~ 2 GPa, while for 2B300 (2 passes ECAE at
300 °C) and 1000ECAE (one pass ECAE at 1000 °C and
followed by WQ) the hardness value almost constant at
annealing temperatures of 500 °C, but as the temperature
increases, it begins to decrease sharply. Also it is worth
mentioning that the softening in the 1000ECAE and
2B300 specimens after annealing may have different
underlying mechanisms. As the 1000ECAE specimens
have primarily martensite, its softening after tempering is
due to the decomposition of martensite. In contrast, the
2B300 specimen has mostly ferrite, and the softening
after annealing is primarily due to grain coarsening.
The AA specimen has a high work hardening rate and
large uniform elongation as the large grains can
accommodate a rapid increase of the density of mobile
dislocations during deformation. In contrast, the
1000WQ specimen has less work hardening capability
and uniform strain as its smaller martensite laths already
contain a high density of dislocations.
The yield strength-uniform elongation paradox in
T91 steels was examined [103]. Amajority of metallic
materials, such as those processed by cold working,
follow the general syndrome: the strengthening of a
material is often accompanied by a reduction in its
ductility.
As shown in Fig. 9, the TMT processes enable to
tailor the mechanical properties of T91 steel over a broad
range. Three distinct groups are highlighted on this plot.
First, the light blue band outlines T91 steel with primarily
ferrite phases. These steels were processed via ECAE at
low-to-intermediate temperatures (room temperature to
700 °C). The higher strength of ferritic T91 steels
originates mostly from grain refinement. The highest
strength achieved in ferritic T91 steel in [103] is
1000 MPa with a mere 2% uniform elongation. In the
second category (dominated by martensite shown as red
band in Fig. 9), TMTs via austenitizing at 900 °C or
greater (followed by WQ) and ECAE at 1000 °C induce
primarily martensite. The yield strength of 1000WQ/500
10h/1A300 T91 exceeds 1600 MPa with, however, only
~ 1.5% uniform elongation. HT of the ECAE specimens
leads to an improvement of uniform elongation, to ~ 4%,
with a yield strength of 1400 MPa. The third category
includes T91 steels processed by WQ at 800 °C or WQ
at 900…1200 °C followed by a 600 °C tempering
procedure as shown in the purple band. T91 processed by
this route is composed of both ferrite and martensite. The
formation of ferrite reduces the yield strength by
~ 150 MPa. It is noteworthy that when the austenitizing
process is performed above 900 °C, WQ leads to the
production of FM T91 steel composed primarily of
martensite. The as-quenched martensite results in high
yield strength, ~ 1200 MPa, as well as a reasonable
uniform elongation, ~ 6…8% as shown by the yellow
triangles. The retention of good ductility may be related
to the low C concentration in T91 steel as well as auto-
78 ISSN 1562-6016. ВАНТ. 2022. №4(140)
tempering of martensite during the cooling process. The
accomplishment of high strength with good ductility via
such a simple WQ process is a very encouraging
observation. While the strength and ductility of
martensitic T91 (red band) and ferritic T91 (light blue
band) may still follow the paradoxical relationship,
ductile martensite can bring the coordinates of the
strength and ductility to a greater level beyond the
ferrites. The FM T91 (in the purple band) bridges the
mechanical behavior between the two distinct groups
(red and light blue band). This TMT roadmap thus
permits the tailoring of T91 steels with various
combinations of strength and ductility for different
industrial and nuclear applications.
Fig. 9. Summary of yield strength-uniform
elongation (a) and yield strength-elongation to
failure (b) maps for T91 steel subjected to various types
of TMTs [103]
The hardness of G91 steel at various HT and hot
rolled conditions is investigated in [104]. The initial steel
exhibits the hardness about 2770 MPa. With the
reduction in tempering temperature (650 ºC), the
hardness increases to ~ 3710 MPa while keeping the
normalizing temperature (1050 ºC) same AR condition.
At high tempering temperature, dissolved solute
elements are re-precipitated in the form of M23C6 and
MX, which results into a loss in solid solution
strengthening. The loss in solid solution strengthening
indicates the decrease in hardness value. In addition,
hardness value increases with the increase in normalizing
temperature while keeping tempering temperature same.
As majority of precipitates dissolve at high normalizing
tempering temperature and leads to solid solution
strengthening. Previous studies show that the hardness
mainly depends on primary/secondary phases of MX
particles, presence of N and C in the solution, dislocation
density and grain size [124]. Hardness of 9Cr-1Mo steel
is less affected by prior austenite grain size rather than
packet size [130]. Specimen treated at R550-T650 and
R1050-T650 conditions have hardness value equal to
3680 and 3580 MPa, respectively. This might be due to
high residual stress generated during hot rolling
treatment.
For tempering temperatures (650 and 700 ºC), σ0.2 and
UTS is found to increase with the increase in normalizing
temperature. Specifically, a significant increase in σ0.2
and UTS is observed at tempering temperature of 650 ºC,
whereas there is a reduction in σ0.2 and UTS for
tempering at 700 °C as compared to tempering at 650 °C.
In case of N1050-T650 (N – normalization), the yield
strength increases by 37.3% at 20 ºC and 73.2% at 650 ºC
as compared to AR condition i.e. N1050-T750.
Correspondingly, the UTS increases by 34.2% at 20 ºC
and 71.5% at 650 ºC and the elongation reduces by 10.9%
at 20 ºC and by 22.7% at 650 ºC. The improvement in
mechanical strength at low tempering temperature (i.e.
650 ºC) suggests the contribution of precipitate size and
dislocation density. The high tempering temperature (i.e.
750 ºC) causes the growth of M23C6 particles, subgrain
boundary formation and recovery.
Specimen processed at R550-T650 and R1050-T650
has got a significant improvement in σ0.2 and UTS at
650 ºC as compared to AR condition. In case of R550-
T650, the σ0.2 increases by 36.1% at 20 ºC and 98.0% at
650 ºC as compared to AR condition whereas the UTS
increases by 54.9% at 20 ºC and 107.0% at 650 ºC.
However, the elongation decreases by 32.9% at 20 °C
and 37.5% at 650 ºC. The enhancement in mechanical
properties followed by rolling is more as compared to
heat-treated conditions. This may be due to high extent
of decrease in lath width, highly dense second phase
precipitates and increase in dislocation density.
For specimen treated at R1050-T650 conditions, the
yield strength increases by 29.3% at 20 ºC and 76.3% at
650 ºC as compared to AR condition. Correspondingly,
the UTS increases by 35.6% at 20 ºC and 74.2% at
650 ºC, and the elongation decreases by 27.7% at 20 ºC
and 29.2% at 650 ºC.
The evolutions of microhardness of T91 steel with
number of extrusion passes (from 1 to 6) before and after
the annealing are presented in [105]. It is noted that the
microhardness exhibits a relatively steep increases from
initial 2700 to 3300 MPa after one pass, then gradually
increases up to about 3800 MPa until six passes, implying
that the grain refinement is most appreciable at the first
pass, which is consistent with other researches of ECAP
processing for different materials [126, 131, 132]. After
the 2 h annealing at 500 °C, the microhardness drops
slightly, which may result from the slight grain growth
and vanishment of a number of dislocations.
The initial T91 steel exhibits an elongation of 23%
and tensile strength of 730 MPa, which is higher than the
elongation of 16% but is lower than the tensile strength
of 848 MPa of T91 steel in Ref. [133]. After single and 6
extrusion passes, the tensile strength increases to 920 and
1160 MPa while the elongation drastically drops to 16
and 10%, respectively. With the increasing number of
a
b
ISSN 1562-6016. ВАНТ. 2022. №4(140) 79
extrusion passes the tensile strength increases while the
elongation decreases. This suggests that the improvement
in strength accompanies with the loss of the ductility in
the ECAPed T91 steel. After the annealing at 500 °C for
2 h, the tensile strength of the samples extruded one and
six passes decreases to 880 and 1140 MPa, while the
elongation increases to 21 and 12%, respectively.
Similarly, an evident drop in tensile strength had been
observed when annealing temperatures were elevated to
600 and 700 °C [134]. In general, the ductility of the
ECAPed samples can be improved by annealing at the
expense of a little decrease of tensile strength. However,
the slight increase of tensile strength after annealing for
the two passes and four passes ECAPed samples is
abnormal, which may be attributed to the experimental
errors caused by the small sample dimension
(16 × 1.7 × 1 mm) and the estimation of the tensile
strength.
The effect of UE number and subsequent HT of T91
steel on mechanical properties is studied in [106]. From
Fig. 10,a it can be seen that with a grain size decreasing
(consequently, with an increase in the number of UE
cycles) the hardness of the samples increases and reaches
the value of ~ 2900 MPa after 5 cycles of UE
(vs. ~ 1900 MPa without deformation). It was
established that optimal temperature of HT is 550 °C at
which a high level of hardness is maintained up to 100 h
of exposure. What is more for thermal treatment at this
temperature during 25 h low increase in microhardness is
observed (approximately by 100 MPa) despite the
increase in grain size from 145 to 245 nm, decrease in
dislocation density from 5.4·1010 to 2.88·1010 cm-2 and
decrease in concentration of carbon dissolved into matrix
from 0.053 to 0.0124 wt.%. The reason for this is,
apparently, the formation of high number of fine MX-
type precipitates during thermal treatment, which induce
the increase in dispersion hardening. At T = 600 °C and
higher the intensive growth of grains occurs also as
abrupt decrease in microhardness with the increase in
time of thermal treatment (see Fig. 10,b). The effect of
dispersion hardening occurs to be insufficient for
compensation of strength loss induced by the grain size
increase.
Strength characteristics of steel T91 (σ0.2 and σB) with
submicron ferritic microstructure, obtained by SPD and
by SPD with subsequent thermal treatment exceed
characteristics of steel with martensitic microstructure
obtained by standard processing “normalization +
tempering” in temperature range -196…+550 °C.
Herewith their ductility remains on sufficient level. It
was revealed for the first time that in specimens of steel
T91 with martensitic microstructure with a decrease of
testing temperature in range of liquid nitrogen not only
strength increases but also ductility improves (rupture
elongation δ). For specimens with ferritic microstructure
such behavior was not observed. This difference is
probably related to the structural features of the
martensitic laths boundaries. But further investigation is
necessary to explain this effect.
Fig. 10. Dependence of the mean grain size (dmg) and microhardness (HV) of T91 steel with the number of cycles
(zero cycle corresponds to the specimen exposed at 750 °C during 15 min) (a); dependence of microhardness of
specimens subjected to 5 cycles of upsetting-extrusion on time of HT at different temperatures (b).
Dashed line defines the level of microhardness in initial state (tempered martensite) [106]
Finally, the characteristics of strength (σ0, the
proportionality limit; σB, the ultimate tensile strength)
and plasticity (δ, the rupture elongation) of
12CrWMoNbVB after upsetting and the next HT was
studied in [108].
The strength properties of materials decrease
monotonically as the test temperature rises. The highest
values of hardness are obtained after upsetting
(~ 3700 MPa), the lowest – after upsetting and tempering
at 720 °C for hour (~ 2800 MPa), the hardness of initial
sample was ~ 2620 MPa.
Upsetting led to an increase in steel strength
characteristics relative to their values for initial state in
the range from ~ 10 to ~ 60%, depending on the test
temperature. Tempering after ausforming led to a drop in
the values of σ0 and σB in the entire temperature range of
research, but nevertheless these characteristics remained
predominantly higher than in initial sample. The relative
elongation of the sample of 12CrWMoNbVB steel after
ausforming decreases relative to the values in standard
steel. Tempering increases the δ values, and at high test
temperatures (550 and 650 °C) they exceed those in the
standard steel (~ 10 %).
a b
80 ISSN 1562-6016. ВАНТ. 2022. №4(140)
CONCLUSIONS
This review considers the main traditional and
reduced activation 9…12% Cr FM steels chosen as
structural materials for future generation reactors due to
their excellent radiation tolerance, acceptable mechanical
characteristics, high thermal conductivity and low
thermal expansion in comparison with conventional
austenitic steels. The effect of TMT and HT on the
structure and mechanical properties of these steels has
been studied. Based on the above, the following
conclusions can be drawn:
1. Normalization temperature affects grain size,
while cooling rate and tempering temperature have little
effect on grain growth. At the same time, with an increase
in the normalization temperature, despite grain growth,
the hardness of 9…12% Cr steels in most cases also
increases. This is mainly related to the formation of Nb-
and Ta-rich carbides, that is precipitation strengthening.
2. It was defined, that the mass fraction of M23C6
precipitates remained constant up to the AC1 temperature
but decreased drastically as austenite phase started
forming. Given that austenite has higher affinity for C
than α-ferrite, the M23C6 precipitates exist up to 865 °C,
and then dissolve into the austenitic matrix. The mass
fraction of MX type precipitates is constant till AC3
temperature, but then gradually decreased with increased
temperature and totally dissolved at ~ 1200 °C.
3. Various types of thermal (normalizing,
austenitizing, partial tempering, water and oil quenching)
and thermomechanical (ausforming, equal channel
angular pressing or extrusion, rolling, multiple
“upsetting-extrusion”, upsetting) treatment of
9…12% Cr steels are presented. It was found that 9% Cr
FM steels became better subjected to TMT, and 12% Cr
– to HT. But for 9% Cr FM steels, HT is also provided to
improve structural factors with the following TMT
procession.
4. Atypical thermal and thermomechanical
treatments of 9% Cr FM steels are considered. It has been
established that preliminary HT by the partial tempering
method makes it possible to modernize the structure
before subsequent mechanical treatment. MUE allows to
achieve an ultrafine grained structure and qualitatively
improve the characteristics of T91 steel.
5. It is necessary to carry out comprehensive studies
of the effect of phases formed in FM steels after water
quenching on high-temperature characteristics and
radiation resistance.
ACKNOWLEDGEMENTS
The work was financially supported by the National
Academy of Science of Ukraine (program “Support of
the development of main lines of scientific
investigations” (KPKVK 6541230)).
REFERENCES
1. L.K. Mansur et al. Materials needs for fusion,
Generation IV fission reactors and spallation neutron
sources – similarities and differences // J. Nucl. Mater.
2004, v. 329-333, p. 166-172.
2. S.M. Goldberg, R. Rosner. Nuclear reactors:
generation to generation. Cambridge, MA: American
Academy of Arts and Sciences, 2011, 29 p.
3. R. Klueh, D. Harries. High-Chromium Ferritic
and Martensitic Steels for Nuclear Applications. West
Conshohocken: ASTM International, 2001, 221 p.
4. R.L. Klueh, A.T. Nelson. Ferritic/martensitic
steels for next-generation reactors // J. Nucl. Mater. 2007,
v. 371, issue 1-3, p. 37-52.
5. C. Cabet, F. Dalle, E. Gaganidze, J. Henry,
H. Tanigawa. Ferritic-martensitic steels for fission and
fusion applications // J. Nucl. Mater. 2019, v. 523,
p. 510-537.
6. T. Jayakumar, M.D. Mathew, K. Laha. High
temperature materials for nuclear fast fission and fusion
reactors and advanced fossil power plants // Procedia
Eng. 2013, v. 55, p. 259-270.
7. J.S. Cheon et al. Sodium fast reactor evaluation:
Core materials // J. Nucl. Mater. 2009, v. 392, issue 2,
p. 324-330.
8. C. Cawthorne, E. Fulton. Voids in irradiated
stainless steel // Nature. 1967, v. 216, p. 575-576.
9. F. Masuyama. New developments in steels for
power generation boilers // Advanced Heat Resistant
Steels for Power Generation. London: IOM
Communications Ltd, 1999, p. 33.
10. J.W. Davis, D.J. Michel. Topical Conference on
Ferritic Alloys for Use in Nuclear Energy Technologies.
Warrendale, PA: The Metallurgical Society of AIME,
1984, p. 261.
11. J. Bishop. Metallurgy in relation to reactor
design and development // Dimensional stability and
mechanical behavior of irradiated metals and alloys.
London: Thomas Telford Publishing, 1984, p. 115.
12. J.W. Bennett, K.E. Horton. Materials requi-
rements for liquid metal fast breeder reactors // Metall.
Mater. Trans. A. 1978, v. 9, issue 2, p. 143-149.
13. H. Brager, J. Perrin. Effects of Radiation on
Materials: Eleventh Conference. Philadelphia PA:
American Society for Testing and Materials, 1982,
p. 1197.
14. R.L. Klueh, J.P. Shingledecker, R.W. Swin-
deman, D.T. Hoelzer. Oxide dispersion-strengthened
steels: A comparison of some commercial and
experimental alloys // J. Nucl. Mater. 2005, v. 341,
issue 2-3, p. 103-114.
15. R.L. Klueh. Elevated temperature ferritic and
martensitic steels and their application to future nuclear
reactors // Int. Mater. Rev. 2005, v. 50, issue 5, p. 287-
310.
16. S.J. Zinkle, J.T. Busby. Structural materials for
fission and fusion energy // Mater. Today. 2009, v. 12,
issue 11, p. 12-19.
17. C. Fazio et al. European cross-cutting research
on structural materials for Generation IV and
transmutation systems // J. Nucl. Mater. 2009, v. 392,
issue 2, p. 316-323.
18. L. Malerba et al. Materials for Sustainable
Nuclear Energy: A European Strategic Research and
Innovation Agenda for All Reactor Generations //
Energies. 2022, v. 15, issue 5, p. 1845.
19. E.A. Little, D.A. Stow. Void-swelling in irons
and ferritic steels: II. An experimental survey of
materials irradiated in a fast reactor // J. Nucl. Mater.
1979, v. 87, issue 1, p. 25-39.
ISSN 1562-6016. ВАНТ. 2022. №4(140) 81
20. V.N. Voyevodin, G.D. Tolstolutskaya,
M.A. Tikhonovsky, A.S. Kuprin, A.S. Kalchenko.
Mechanisms of radiation damage and development of
structural materials for operating and advanced nuclear
reactors // Problems of Atomic Science and Technology
(PAST). 2021, N 5(135), p. 3-20.
21. E. Getto et al. Void swelling and microstructure
evolution at very high damage level in self-ion irradiated
ferritic-martensitic steels // J. Nucl. Mater. 2016, v. 480,
p. 159-176.
22. V. Voyevodin et al. Structural features and
operational characteristics of steel T91 // East Eur. J.
Phys. 2020, v. 3, p. 93-98.
23. S. Nomura et al. Development of long life FBR
core materials // Proceedings of the International
Conference on Fast Reactors and Related Fuel Cycles.
Tokyo: Atomic Energy Society of Japan, 1991, Paper 7.4.
24. A. Uehira, S. Ukai, T. Mizuno, T. Asaga,
E. Yoshida. Tensile Properties of llCr-0.5Mo-2W, V, Nb
Stainless Steel in LMFBR Environment // J. Nucl. Sci.
Technol. 2000, v. 37, p. 780-786.
25. K. Anderko. Zur eignung warmfester
vergütungsstähle mit 9 bis 12% chrom für komponenten
im kern schneller reaktoren – ein überblick // J. Nucl.
Mater. 1980, v. 95, p. 31-43.
26. Technical reports. Structural materials for
liquid metal cooled fast reactor fuel assemblies
operational behaviour. Vienna: International Atomic
Energy Agency, 2012, p. 19.
27. C. Brown et al. Cladding and wrapper
development for fast breeder reactor high performance //
Proceedings of the International Conference on Fast
Reactors and Related Fuel Cycles. Tokyo: Atomic
Energy Society of Japan, 1991, Paper 7.5.
28. J.W. Davis, D.J. Michel. Topical Conference on
Ferritic Alloys for Use in Nuclear Energy Technologies.
Warrendale, PA: The Metallurgical Society of AIME,
1984, p. 299.
29. A. Lovell, A. Fox, W. Sutherland, S. Hecht.
Demonstration of a high burnup heterogeneous core
using ferritic/martensitic materials // Proceedings of the
International Conference on Reliable Fuels for Liquid
Metal Reactors. Richland, WA: American Nuclear
Society, 1987, p. 3.
30. R. Viswanathan. Materials technology for coal
fired power plants // Adv. Mater. Process. 2004, v. 162,
p. 73-76.
31. B.H. Sencer, J.R. Kennedy, J.I. Cole, S.A. Ma-
loy, F.A. Garner. Microstructural analysis of an HT9 fuel
assembly duct irradiated in FFTF to 155 dpa at 443° C //
J. Nucl. Mater. 2009, v. 393, p. 235-241.
32. P. Gavoille et al. Mechanical Properties of
Cladding and Wrapper Materials for the ASTRID Fast-
Reactor Project // Proceedings of the International
Conference on Fast Reactors and Related Fuel Cycles:
Safe Technologies and Sustainable Scenarios. Vienna:
International Atomic Energy Agency, 2013, Paper T4.2.
33. C. Fazio, P. Dubuisson. Achievement and new
challenges for high performance materials in Europe //
Proceedings of the International Conference on Fast
Reactors and Related Fuel Cycles: Safe Technologies
and Sustainable Scenarios. Vienna: International Atomic
Energy Agency, 2013, p. 487.
34. T. Asayama et al. Development of structural
materials for JSFR – overview and current status //
International Conference on Fast Reactors and Related
Fuel Cycles. Vienna: International Atomic Energy
Agency, 2009, p. 417-431.
35. D. Zhang // 49th Meeting of the Technical
Working Group on Fast Reactors (TWG-FR), Buenos
Aires, Argentina, 2016.
36. J. Yoo. Korean SFR development program and
technical activity for improving economical
competitiveness // Technical Meeting on Fast Reactors
and Related Fuel Cycle Facilities with Improved
Characteristics, IAEA. Vienna, Austria, 2013.
37. T. Ma, X. Hao, P. Wang. Effect of heat
treatments on Charpy impact properties of
15Cr12MoVWN ferritic/martensitic steel // J. Iron Steel
Res. Int. 2022, v. 29, p. 512-518.
38. R.L. Klueh, D.J. Alexander. Heat treatment
effects on impact toughness of 9Cr-1MoVNb and 12Cr-
1MoVW steels irradiated to 100 dpa // J. Nucl. Mater.
1998, v. 253-258, p. 1269-1274.
39. D. S. Gelles. Development of martensitic steels
for high neutron damage applications // J. Nucl. Mater.
1996, v. 239, p. 99-106.
40. N.M. Ghoniem, J. Blink, N. Hoffman. Selection
of alloy steel type for fusion power plant applications in
the 350-500 °C range // Proc. of the Topical Conf. on
Ferritic Alloys for Use in Nuclear Technology,
Snowbird, Utah, 1983.
41. D. Gilbon, C. Rivera. Behavior of different
ferritic steels under ion, electron and fast neutron
irradiation // J. Nucl. Mater. 1988, v. 155-157, p. 1268-
1273.
42. T.R. Allen, J.T. Busby, R.L. Klueh, S.A. Maloy,
M.B. Toloczko. Cladding and duct materials for
advanced nuclear recycle reactors // JOM. 2008, v. 60,
N 1, p. 15-23.
43. D. Frazer et al. Degradation of HT9 under
simultaneous ion beam irradiation and liquid metal
corrosion // J. Nucl. Mater. 2016, v. 479, p. 382-389.
44. Y. Kurata, M. Futakawa, S. Saito. Comparison
of the corrosion behavior of austenitic and
ferritic/martensitic steels exposed to static liquid Pb-Bi at
450 and 500 °C // J. Nucl. Mater. 2005, v. 343, p. 333-
340.
45. R. Lindau et al. Present development status of
EUROFER and ODS-EUROFER for application in
blanket concepts // Fus. Eng. Des. 2005, v.75-79, p. 989-
996.
46. D. Kumar et al. The effects of fusion reactor
thermal transients on the microstructure of Eurofer-97
steel // J. Nucl. Mater. 2021, v. 554, p. 153084.
47. G. Mazzone et al. Choice of a low operating
temperature for the DEMO EUROFER97 divertor
cassette // Fusion Eng. Des. 2017, v. 124, p. 655-658.
48. J. Roth et al. EUROFER as wall material:
Reduced sputtering yields due to W surface enrichment
// J. Nucl. Mater. 2014, v. 454, p. 1-6.
49. J.H. You et al. Divertor of the European DEMO:
Engineering and technologies for power exhaust //
Fusion Eng. Des. 2022, v. 175, p. 113010.
82 ISSN 1562-6016. ВАНТ. 2022. №4(140)
50. G. Benamati et al. Mechanical and corrosion
behaviour of EUROFER 97 steel exposed to Pb-17Li //
J. Nucl. Mater. 2002, v. 307-311, p. 1391-1395.
51. V. Sikka, C. Ward, K. Thomas. Modified 9Cr-
1Mo steel – an improved alloy for steam generator
applications // Proc. International Conf. on Ferritic
Steels for High Temperature Applications. Metals Park,
OH: American Society for Metals, 1983, p. 65.
52. G. Bodine, R. McDonald. Laboratory and pilot
commercial process/product development of modified
9Cr-1Mo ferritic alloy // Proc. International Conf. on
Ferritic Steels for High Temperature Applications.
Metals Park, OH: American Society for Metals, 1983,
p. 9-20.
53. G. Cunningham, P. Patriarca, E. Hoffman.
Ferritic steels as alternate structural materials for high-
temperature applications // Proc. International Conf. on
Ferritic Steels for High Temperature Applications.
Metals Park, OH: American Society for Metals, 1983,
p. 3-6.
54. B. Roberts, R. Swindeman. TVA experience in
the application of 9Cr-1Mo-0.2V-CB (Grade 91) steel //
Proceedings of the EPRI Conference on 9Cr Materials
Fabrication and Joining Technologies. Charlotte, NC:
Fossil Repair Applications Center (FRAC), 2001, Paper
20-1.
55. R. Swindeman, M. Santella, P. Maziasz,
B. Roberts, K. Coleman. Issues in replacing Cr-Mo steels
and stainless steels with 9Cr-1Mo-V steel // Int. J. Press.
Vessel. Pip. 2004, v. 81, p. 507-512.
56. S. Dubiez-Legoff et al. Selection of materials
for sodium fast reactor steam generators // Proceedings
of the 2012 International Congress on Advances in
Nuclear Power Plants – ICAPP '12. La Grange Park, IL:
American Nuclear Society, 2012, Paper 12053.
57. C. Cabet, F. Rouillard. Corrosion phenomena
induced by gases in Generation IV nuclear reactors //
Structural materials for generation IV nuclear reactors.
Sawston: Woodhead Publishing, 2017, p. 75-104.
58. S.C. Chetal et al. Current Status of Fast Reactors
and Future Plans in India // Energy Procedia. 2011, v. 7,
p. 64-73.
59. K. Aoto et al. Design study and R&D progress
on Japan sodium-cooled fast reactor // J. Nucl. Sci.
Technol. 2011, v. 48, p. 463-471.
60. Y. Chikazawa et al. Evaluation of JSFR key
technologies // Nucl. Technol. 2012, v. 179, p. 360-373.
61. https://www.westinghousenuclear.com/new-
plants/lead-cooled-fast-reactor
62. G. Bauer, M. Salvatores, G. Heusener.
MEGAPIE, a 1 MW pilot experiment for a liquid metal
spallation target // J. Nucl. Mater. 2001, v. 296, p. 17-33.
63. F. Bianchi et al. Status and trend of core design
activities for heavy metal cooled accelerator driven
system // Energy Convers. Manag. 2006, v. 47, p. 2698-
2709.
64. Y. Dai et al. Assessment of the lifetime of the
beam window of MEGAPIE target liquid metal container
// J. Nucl. Mater. 2006, v. 356, p. 308-320.
65. W. Wagner, F. Gröschel, K. Thomsen,
H. Heyck. MEGAPIE at SINQ – The first liquid metal
target driven by a megawatt class proton beam // J. Nucl.
Mater. 2008, v. 377, p. 12-16.
66. A. Alemberti et al. European lead fast reactor–
ELSY // Nucl. Eng. Des. 2011, v. 241, p. 3470-3480.
67. A. Alemberti. European Strategy for LFRs
Development // Proceedings of the International
Workshop on Innovative Nuclear Reactors Cooled by
Heavy Liquid Metals: Status and Perspectives, Pisa,
Italy, 2012.
68. A. Alemberti, V. Smirnov, C. Smith,
M. Takahashi. Overview of lead-cooled fast reactor
activities // Prog. Nucl. Energy. 2014, v. 77, p. 300-307.
69. H. Abderrahim, P. Baeten, D. Bruyn,
R. Fernandez. MYRRHA – A multi-purpose fast
spectrum research reactor // Energy Convers. Manag.
2012, v. 63, p. 4-10.
70. P. Schuurmans. Technical Meeting on Liquid
Metal Reactor Concepts: Core Design and Structural
Materials, Jun 2013, Vienna (Austria).
71. W.R. Corwin et al. Generation IV reactors
integrated materials technology program plan: focus on
very high temperature reactor materials // Technical
Report. United States, 2008; doi:10.2172/951084.
72. D. Hittner et al. RAPHAEL: Development of
advanced technologies for high and very high
temperature reactors // Proceedings of the FISA.
Luxembourg: Office for Official Publ. of the European
Communities, 2006, p. 196-209.
73. C. Poette et al. Status of the ETDR Design //
Proceedings of the ICAPP 2007 Conference, 13-18 May
2007, Nice, France, Paper 7208.
74. H. Burlet et al. Evaluation of Nickel-Based
Materials for VHTR Heat Exchanger // Proceedings of
the OECD NEA NSC Workshop on Structural Materials
for Innovative Nuclear Systems (SMINS). London:
OECD Publishing, 2008, p. 79.
75. F. Carre, P. Yvon, P. Anzieu, N. Chauvin, J.-
Y. Malo. Update of the French R&D strategy on gas-
cooled reactors // Nucl. Eng. Des. 2010, v. 240, p. 2401-
2408.
76. G. Was et al. Corrosion and stress corrosion
cracking in supercritical water // J. Nucl. Mater. 2007,
v. 371, p. 176-201.
77. C. Sun, R. Hui, W. Qu, S. Yick. Progress in
Corrosion Resistant Materials for Supercritical Water
Reactors // Corros. Sci. 2009, v. 51, p. 2508-2523.
78. P. Hejzlar et al. TERRAPOWER, LLC traveling
wave reactor development program overview // Nucl.
Eng. Tech. 2013, v. 45, p. 731-744.
79. G. Was et al. Emulation of reactor irradiation
damage using ion beams // Scripta Mater. 2014, v. 88,
p. 33-36.
80. J. Wharry, G. Was. A systematic study of
radiation-induced segregation in ferritic–martensitic
alloys // J. Nucl. Mater. 2013, v. 442, p. 7-16.
81. J. Gigax et al. Radiation response of alloy T91
at damage levels up to 1000 peak dpa // J. Nucl. Mater.
2016, v. 482, p. 257-265.
82. E. Getto, G. Vancoevering, G. Was. The co-
evolution of microstructure features in self-ion irradiated
HT9 at very high damage levels // J. Nucl. Mater. 2017,
v. 484, p. 193-208.
83. A. Alamo, V. Lambard, X. Averty, M.H. Mat-
hon. Assessment of ODS-14%Cr ferritic alloy for high
https://www.westinghousenuclear.com/new-plants/lead-cooled-fast-reactor
https://www.westinghousenuclear.com/new-plants/lead-cooled-fast-reactor
ISSN 1562-6016. ВАНТ. 2022. №4(140) 83
temperature applications // J. Nucl. Mater. 2004, v. 329-
333, p. 333-337.
84. R. L. Klueh. Analysis of swelling behaviour of
ferritic/martensitic steels // Philos. Mag. 2008, v. 98,
issue 28, p. 2618-2636.
85. M. Rieth et al. EUROFER 97 Tensile, charpy,
creep and structural tests (FZKA – 6911). Germany,
2003.
86. https://www.astm.org/a0213_a0213m-18.html
87. H. Zhang, B. Long, Y. Dai. Metallography
studies and hardness measurements on
ferritic/martensitic steels irradiated in STIP // J. Nucl.
Mater. 2008, v. 377, issue 1, p. 122-131.
88. E.B. Kula, M. Azrin. Thermomechanical
Processing of Ferrous Alloys // Advances in Deformation
Processing. Boston, MA: Springer, 1978, p. 245-300.
89. V.F. Zackay. Thermomechanical processing //
Mater. Sci. Eng. 1976, v. 25, p. 247-261.
90. S. Yue. Thermomechanical processing of
ferrous alloys // Metalworking: Bulk Forming: ASM
Handbook. Materials Park, OH: ASM International,
2005, v. 14A, p. 286-296.
91. G.E. Totten. Steel heat treatment: metallurgy
and technologies (2nd ed.). Boca Raton, FL: CRC Press,
2007, 1576 p.
92. T. Shrestha et al. Effect of heat treatment on
microstructure and hardness of Grade 91 steel // Metals.
2015, v. 5, p. 131-149.
93. N.Z. Gutiérrez et al. Heterogeneous austenite
grain growth in ASTM A213–T91 Steel // ISIJ Int. 2007,
v. 47, N 8, p. 1178-1187.
94. P.J. Ennis et al. Influence of heat treatments on
microstructural parameters and mechanical properties of
P92 steel // Mater. Sc. and Techn. 2000, v. 16, issue 10,
p. 1226-1233.
95. Z. Shang et al. Tailoring the strength and
ductility of T91 steel by partial tempering treatment //
Acta Mater. 2019, v. 169, p. 209-224.
96. T. Ma et al. Effect of heat treatments on
microstructural evolution and tensile properties of
15Cr12MoVWN ferritic/martensitic steel // Metals.
2020, v. 10, p. 1271.
97. R. Maiti et al. Mechanical properties of HT-9 as
a function of heat treatment // J. Nucl. Mater. 1986,
v. 141-143, p. 527-531.
98. H.M. Heo et al. Effect of heat treatment and
aging conditions on the microstructure and mechanical
properties of HT9 steel for fuel cladding tube // Korean
J. Met. Mater. 2015, v. 53, № 3, p. 177-186.
99. Z. Lu et al. Effect of heat treatment on
microstructure and hardness of Eurofer 97, Eurofer ODS
and T92 steels // J. Nucl. Mater. 2009, v. 386-388, p. 445-
448.
100. C. Testani et al. Austenitization and tempering
temperatures effects on EUROFER 97 steel // Materials
Science Forum. 2018, v. 941, p. 711-716.
101. J. Hoffmann et al. Improvement of reduced
activation 9%Cr steels by ausforming // Nucl. Mater.
Energy. 2016, v. 6, p. 12-17.
102. P. Fernández et al. Microstructural and
mechanical characteristics of EUROFER’97 processed
by equal channel angular pressing // J. Nucl. Mater. 2011,
v. 417, p. 20-24.
103. M. Song et al. A roadmap for tailoring the
strength and ductility of ferritic/martensitic T91 steel via
thermo-mechanical treatment // Acta Mater. 2016,
v. 112, p. 361-377.
104. S.S. Samant et al. Influence of intermediate
rolling on mechanical behavior of modified 9Cr-1Mo
steel // Mater. Sci. Eng. A. 2018, v. 738, p. 135-152.
105. Z.Q. Fan et al. The microstructure and
mechanical properties of T91 steel processed by ECAP
at room temperature // J. Nucl. Mater. 2013, v. 434,
p. 417-421.
106. V.N. Voyevodin et al. A new approach to
thermo-mechanical treatment of steel T91 by multiple
upsetting-extrusion in a ferritic range // Mater. Sci. Eng.
A. 2021, v. 822, 141686;
https://doi.org/10.1016/j.msea.2021.141686
107. Патент 42487 A Україна, МПК 7 H01B12/00.
Спосіб виготовлення ніобій-титанового
надпровідник / О.В. Чорний, Я.Д. Стародубов,
О.І. Волчок, Г.Є. Сторожилов. Бюл. №9. Заявл.
12.03.2001; Опубл. 15.10.2001.
108. I.F. Kislyak et al. Effect of ausforming on
mechanical properties of 12%Cr ferritic/martensitic steel
// Problems of Atomic Science and Technology. 2022,
N 1(137), p. 76-84.
109. Y.Z. Shen et al. Identification of precipitate
phases in a 11Cr ferritic/martensitic steel using electron
microdiffraction // J. Nucl. Mater. 2010, v. 400, p. 64-68.
110. P. Anderson, T. Bellgard, F.L. Jones. Creep
deformation in a modified 9Cr-1Mo steel // Mater. Sci.
Technol. 2003, v. 19, p. 207-213.
111. F. Abe, S. Nakazawa, H. Araki, T. Noda. The
role of microstructural instability on creep behaviour of
a low radioactivation martensitic 9Cr-2 W steel // Metal.
Trans. A. 1992, v. 23, p. 469-477.
112. M. Tamura et al. Development of potential low
activation ferritic and austenitic steels // J. Nucl. Mater.
1986, v. 141-143, p. 1067-1073.
113. M. Tamura, K. Ikeda, H. Esaka, K. Shinozuka.
Precipitation behavior of NbC in 9%Cr1%Mo0.2%VNb
steel // ISIJ Int. 2001, v. 41, p. 908-914.
114. R. C. Wilcox, B. A. Chin. Austenitizing and
microstructure of a HT-9 steel // Metallography. 1984,
v. 17, issue 3, p. 285-298.
115. P. Fernandez et al. Reduced Activation
Ferritic/Martensitic Steel Eurofer 97 as Possible
Structural Material for Fusion Devices Metallurgical
Characterization on As-Received Condition and after
Simulated Services Conditions // CIEMAT-1048, Spain,
2004.
116. H. Rostova et al. Cavitation wear of Eurofer 97,
Cr18Ni10Ti and 42HNM alloys // Acta Polytechnica.
2021, v. 61(6), p. 762-767.
117. X. Hu, N. Xiao, X. Luo, D. Li. Effects of δ-
ferrite on the microstructure and mechanical properties in
a tungsten-alloyed 10 %Cr ultra-supercritical steel // Acta
Metall. Sin. 2009, v. 45, p. 553-558.
118. P. Hu et al. Nitride-strengthened reduced
activation ferritic/martensitic steels // Fusion Eng. Des.
2010, v. 85, p. 1632-1637.
119. J.W. Davis, D.J. Michel. Topical Conference on
Ferritic Alloys for Use in Nuclear Energy Technologies.
https://www.astm.org/a0213_a0213m-18.html
https://doi.org/10.1016/j.msea.2021.141686
84 ISSN 1562-6016. ВАНТ. 2022. №4(140)
Warrendale, PA: The Metallurgical Society of AIME,
1984, p. 299.
120. J.W. Davis, D.J. Michel. Topical Conference on
Ferritic Alloys for Use in Nuclear Energy Technologies.
Warrendale, PA: The Metallurgical Society of AIME,
1984, p. 347.
121. K. Anderko, L. Schäfer, E. Materna-Morris.
Effect of the δ-ferrite phase on the impact properties of
martensitic chromium steels // J. Nucl. Mater. 1991,
v. 179-181, p. 492-495.
122. Г.Ю. Ростова та ін. Кавітаційне зношування
ферритно-мартенситної сталі Т91 // Фізико-хімічна
механіка матеріалів (ФХММ). 2022, №3, с. 79-83.
123. A. Zielinska-Lipiec, A. Czyrska-Filemonowicz,
P.J. Ennis, O. Wachter. The influence of heat treatments
on the microstructure of 9% chromium steels containing
tungsten // J. Mater. Process. Technol. 1997, v. 64,
p. 397-405.
124. D.R. Barbadikar et al. Effect of normalizing and
tempering temperatures on microstructure and
mechanical properties of P92 steel // Int. J. Pres. Ves. Pip.
2015, v. 132, p. 97-105.
125. Z.G. Liu et al. Electron-microscopy investi-
gation on nanocrystal formation in pure Fe and carbon
steel during ball milling // Mater. Sci. Eng. A. 2003,
v. 362, p. 322-326.
126. D.H. Shin, I. Kim, J. Kim, K.T. Park. Grain
refinement mechanism during equal-channel angular
pressing of a low-carbon steel // Acta Mater. 2001, v. 49,
p. 1285-1292.
127. J.Y. Huang, Y.T. Zhu, H. Jiang, T.C. Lowe.
Microstructures and dislocation configurations in
nanostructured Cu processed by repetitive corrugation
and straightening // Acta Mater. 2001, v. 49, p. 1497-
1505.
128. G.J. Raab, R.Z. Valiev, T.C. Lowe, Y.T. Zhu.
Continuous processing of ultrafine grained Al by ECAP-
Conform // Mater. Sci. Eng. A. 2004, v. 382, p. 30-34.
129. J. Vivas et al. Effect of ausforming temperature
on the microstructure of G91 steel // Metals. 2017, v. 7,
p. 236.
130. A. Chatterjee, D. Chakrabarti, A. Moitra,
R. Mitra, A.K. Bhaduri. Effect of deformation
temperature on the ductile-brittle transition behavior of a
modified 9Cr-1Mo steel // Mater. Sci. Eng. A. 2015,
v. 630, p. 58-70.
131. B.Q. Han, S. Yue. Processing of ultrafine ferrite
steels // J. Mater. Process Technol. 2003, v. 136, p. 100-
104.
132. W.Q. Cao et al. Strain path effects on the
microstructure evolution and mechanical properties of
Zr702 // Mater. Sci. Eng. A. 2005, v. 395, p. 77-86.
133. J. Henry, X. Averty, Y. Dai, J.P. Pizzanelli.
Tensile behaviour of 9Cr-1Mo tempered martensitic
steels irradiated up to 20 dpa in a spallation environment
// J. Nucl. Mater. 2008, v. 377, p. 80-93.
134. D.C. Foley, K.T. Hartwig, S.A. Maloy,
P. Hosemann, X. Zhang. Grain refinement of T91 alloy
by equal channel angular pressing // J. Nucl. Mater. 2009,
v. 389, p. 221-224.
Article received 22.07.2022
ЛІТЕРАТУРНИЙ ОГЛЯД: ФЕРИТНО-МАРТЕНСИТНІ СТАЛІ – ОБРОБКА,
СТРУКТУРА ТА МЕХАНІЧНІ ВЛАСТИВОСТІ
Г.Ю. Ростова, Г.Д. Толстолуцька
Вживання електроенергії, що постійно зростає, вимагає розробки та впровадження більш потужних та
енергоємних систем нового покоління. Ядерні та термоядерні установки четвертого покоління (Gen-IV)
дадуть можливість покрити зростаючий попит на електроенергію. Оскільки реактори Gen-IV працюватимуть
за вищих температур і доз опромінення, виникає проблема підбора науково обґрунтованих конструкційних
матеріалів, так як реакторні матеріали, що нині використовуються, не придатні для використання в таких
жорстких умовах експлуатації. Серед конструкційних матеріалів, що розглядаються, для майбутніх поколінь
реакторів особлива увага приділяється 9…12% Cr феритно-мартенситним сталям через їх більш високу
радіаційну толерантність і відмінні механічні властивості порівняно з традиційно використовуваними
аустенітними сталями. У даному огляді розглянуті основні феритно-мартенситні сталі, які будуть
використовуватися як конструкційні матеріали, їх структура, механічні властивості та різні термічні та
термомеханічні обробки, що застосовуються до них.
.
|