You are on page 1of 15

metals

Article
Effect of Sub-Zero Treatment Temperatures on
Hardness, Flexural Strength, and Fracture Toughness
of Vanadis 6 Ledeburitic Die Steel
Peter Jurči 1, *, Ivo Dlouhý 2 , Petra Priknerová 3 and Zdeněk Mrštný 4
1 Faculty of Material Sciences and Technology of the STU in Trnava, J. Bottu 25, 917 24 Trnava, Slovakia
2 Institute of Physics of Materials, CEITEC-IPM, Czech Academy of Sciences, Zizkova 22, 616 62 Brno, Czech
Republic; idlouhy@ipm.cz
3 Prikner—Heat Treatment of Metals, Martínkovice 279, 550 01 Martínkovice, Czech Republic;
p.priknerova@prikner.cz
4 Department of Metal Processing, AIR PRODUCTS spol. s r.o., Ústecká 30, 405 02 Děčín, Czech Republic;
MRSTNYZ@airproducts.com
* Correspondence: p.jurci@seznam.cz; Tel.: +421908674097

Received: 7 November 2018; Accepted: 8 December 2018; Published: 11 December 2018 

Abstract: Any improvement on the service life of tools reduces the tooling costs, and assists
to increase labor productivity by decreasing the needs for either the tools’ re-grinding or their
replacement. This requires, among others, an enhancement of the key mechanical properties of the
tool materials, by newer treatment route development. The current paper describes the impact of
different heat treatment regimes, including austenitizing; sub-zero treatments; and tempering on the
hardness, flexural strength, and toughness of tool steel, which is demonstrated upon Vanadis 6 steel.
An improvement in the hardness due to the sub-zero treatment is reported, but it is also pointed out
that both the flexural strength and fracture toughness of the material cannot be inevitably deteriorated
by the application of this processing. Finally, it is demonstrated that both of these properties, despite
their conflicting relationship, in most cases, can be improved simultaneously when the material is
treated in the proper way.

Keywords: tool steel; sub-zero treatment; microstructure; hardness; flexural strength; fracture
toughness; fractography

1. Introduction
High-chromium high-vanadium ledeburitic cold work die steels have been extensively used in
industrial applications where high hardness and nondeforming qualities are required. In addition,
alloying with vanadium provides the materials with an outstanding wear resistance. To obtain these
beneficial characteristics, the steels must be subjected to a proper heat treatment procedure, which is
specific to each steel grade. This comprises austenitizing, which is followed by rapid quenching to the
room temperature (also called as “conventional heat treatment” (CHT)). As-quenched ledeburitic steels
contain martensite; retained austenite (10–20 vol. %); and undissolved carbides of a different chemistry,
crystallography, size, and shape. The subsequent tempering induces the precipitation of carbides,
converts the retained austenite into either martensite or bainite (in the case of high temperature
tempering), and thereby determines the final bulk microstructure and properties of a particular steel.
It has been demonstrated that the application of the sub-zero treatment (SZT) further extracts
the durability of the tools [1–4]. These improvements are based on the significant reduction of
the retained austenite [5–12], accelerated precipitation rate and more uniform transient precipitates

Metals 2018, 8, 1047; doi:10.3390/met8121047 www.mdpi.com/journal/metals


Metals 2018, 8, 1047 2 of 15

distribution [13,14], martensite refinement [11,12], and substantially enhanced number and population
density of small globular carbides [5–11,14–18].
The mentioned variations in the microstructure provide the steels with a better wear resistance
and high resistivity to undesirable dimensional changes, which is the prerequisite for the enhancement
of the service life time of tools and components. On the other hand, the resistance against crack
initiation/propagation (being quantified by impact toughness, flexural strength, and/or fracture
toughness) is generally lowered by the presence of carbides, as they are brittle in nature. Moreover,
the materials manufactured by “classical” ingot metallurgy suffer from the anisotropy of carbide
particle distribution, as well as from carbide stringers and clusters, which often act as the nuclei of
premature fracture of tools. Despite the fact that it is important to keep at least an acceptable level
of the steel toughness, little attention has been paid to date to its experimental determination for
ledeburitic steels. Moreover, the obtained results manifest an evident inconsistency. For instance,
Surberg et al. [19] and Molinari et al. [20] reported almost no effect of SZT on impact toughness, while
Collins and Dormer [16] and Rhyim et al. [21] recorded a considerable reduction of impact toughness
using the application of SZT for the treatment of the same steel grade. Also, the obtained results
on the fracture toughness, KIC , are mostly inconsistent. Das et al. [17], for instance, reported on the
deterioration of the fracture toughness because of SZT, but this deterioration was most pronounced
after SZTs at −75 and −125 ◦ C, while the SZT at −196 ◦ C led only to a minimal KIC reduction. In our
recent papers [15,22,23], on the other hand, almost no effect, or rather fracture toughness, has been
reported for the improvement of Vanadis 6 steel when SZT is at −196 and −140 ◦ C. Also, it has been
demonstrated that the flexural strength is slightly (and rather positively) affected by this treatment.
The question of the optimal SZT temperature and duration at this temperature is still under
debate. In the 1950s, it was believed that temperatures down to approximately −75 ◦ C were sufficient
to transform a substantial portion of the retained austenite into the martensite, and that lower
temperatures had no practical effect in the treatment of steels. Much later, a boiling temperature
of liquid nitrogen (−196 ◦ C) was suggested for treatment, and most of the experimental works have
been conducted using this temperature. Alternatively, only few studies suggested other temperatures,
for instance −140 ◦ C, for the treatment of Cr-ledeburitic tool steels [2,24]. Stratton [1], on the other
hand, recommended using a boiling helium temperature (−269 ◦ C) for treatment in selected cases.
The current paper summarizes the findings of the investigations that are focused on the impact of
different SZT conditions (temperature and processing time) on the mechanical performance of Cr-V
ledeburitic cold work die steel Vanadis 6. Contrary to our previous papers in this field, here, we firstly
focused our interest on the phenomena that are affected by different SZT temperatures, including their
crucial impact on the mechanical properties.

2. Material and Experimental Methods

2.1. Material and Processing


The experimental material was powder metallurgy (PM) ledeburitic tool steel Vanadis 6, with a
nominal composition (in wt %) of 2.1% C, 1.0% Si, 0.4% Mn, 6.8% Cr, 1.5% Mo, 5.4% V, and Fe as the
balance. The steel was supplied as soft-annealed flat bars, with a hardness of 270 HV10. The choice of
the PM Cr-V ledeburitic steel was motivated by the fact that it manifests a high degree of microstructure
isotropy, which enables us to disregard the orientation in the sampling [25].
The samples were net-shape machined and were subjected to pre-determined heat treatment
schedules (Figure 1). Conventional heat treatment (CHT) consisted of the following steps:
Gradual heating up to the desired austenitizing temperature TA (1050 ◦ C) in a vacuum furnace
(1), holding at that temperature for 30 min to homogenize the austenite (2), followed by quenching
by nitrogen gas (five bar) (3). One set of specimens was tempered after quenching, without SZT.
Another three sets of specimens were, immediately after quenching to room temperature, moved to the
cryogenic system, where they were cooled down at a cooling rate of 1 ◦ C/min to the SZT temperature
Metals 2018, 8, 1047 3 of 15
Metals 2018, 8, x FOR PEER REVIEW 3 of 16

(4).
SZTs The SZTs
have have
been been carried
carried out atdifferent
out at three three different temperatures,
temperatures, namelynamely −140,and
−140, −196, −196, and
−269 269 ◦for
°C,−and C,
a duration of 17 h (5). After that, the material was re-heated to the room temperature, using a
and for a duration of 17 h (5). After that, the material was re-heated to the room temperature, using
◦ C/min (6). Double tempering (2 + 2 h) was performed immediately ((7) and (8)),
heating rate of 1 °C/min (6). Double tempering (2 + 2 h) was performed immediately ((7) and (8)),
namely at temperatures of 170 or 530 ◦°C. C. In addition, the tempering temperatures of 330 and 450 ◦°C C
were used for the construction of the tempering charts.
construction of the tempering charts.

1. A
Figure 1.
Figure A schematic
schematicofofheat
heattreatment schedules
treatment applied
schedules forfor
applied conventional heatheat
conventional treatment (CHT)
treatment and
(CHT)
sub-zero treatment (SZT).
and sub-zero treatment (SZT).

The SZT temperatures were generated as follows: For treatment at −140 ◦ C, cold nitrogen gas
The SZT temperatures were generated as follows: For treatment at −140 °C, cold nitrogen gas
was introduced into the cryogenic system. The treatment at −196 ◦ C was carried out in liquid nitrogen;
was introduced into the cryogenic system. The treatment at −196 °C was carried out in liquid
however, first, only cold nitrogen gas was supplied into the processing chamber, and only after the
nitrogen; however, first, only cold nitrogen gas was supplied into the processing chamber, and only
specimens were cooled down to approximately −190 ◦ C did the liquid nitrogen come into touch with
after the specimens were cooled down to approximately −190 °C did the liquid nitrogen come into
the steel specimens. The temperature of −269 ◦ C was generated by liquid helium. Also, in this case,
touch with the steel specimens. The temperature of −269 °C was generated by liquid helium. Also,
the samples were cooled down to the temperature of the liquid nitrogen, by introducing cold nitrogen
in this case, the samples were cooled down to the temperature of the liquid nitrogen, by
gas into the cryogenic system, and were held there for 15 min, in order to equalize the temperature
introducing cold nitrogen gas into the cryogenic system, and were held there for 15 min, in order to
through the specimens. Afterwards, the specimens were moved into a Dewar’s container filled with
equalize the temperature through the specimens. Afterwards, the specimens were moved into a
liquid helium, cooled down to the temperature of −269 ◦ C, and held there for 17 h.
Dewar’s container filled with liquid helium, cooled down to the temperature of −269 °C, and held
there for 17 h. Methods
2.2. Experimental
strength was measured on bar test specimens with dimensions of 10 mm × 10 mm ×
The flexural Methods
2.2. Experimental
100 mm, and with a defined final surface roughness of Ra = 0.05–0.07 µm. It should be noted that it is
The flexural strength was measured on bar test specimens with dimensions of 10 mm × 10 mm
very important to define the surface roughness before testing, as it was initially demonstrated that the
× 100 mm, and with a defined final surface roughness of Ra = 0.05–0.07 μm. It should be noted that it
flexural strength, as a measure of the resistance of the material against the fracture initiation, is strongly
is very important to define the surface roughness before testing, as it was initially demonstrated
dependent on the Ra parameter [26]. The specimens were loaded in three-point bending at a loading
that the flexural strength, as a measure of the resistance of the material against the fracture
rate of 1 mm/min, up to the moment of fracture. The distance between the loading roller supports
initiation, is strongly dependent on the Ra parameter [26]. The specimens were loaded in three-point
was 80 mm. The flexural strength was determined from the maximum (fracture) load according to
bending at a loading rate of 1 mm/min, up to the moment of fracture. The distance between the
standard approach. In addition, the work of fracture, W of , was determined as the deformation energy
loading roller supports was 80 mm. The flexural strength was determined from the maximum
evaluated from the corresponding area below the measured load–displacement curve.
(fracture) load according to standard approach. In addition, the work of fracture, Wof, was
For the fracture toughness determination, pre-cracked specimens with dimensions of
determined as the deformation energy evaluated from the corresponding area below the measured
10 mm × 10 mm × 55 mm were used. Both the pre-crack preparation and the tests were carried out
load–displacement curve.
at room temperature according to the ISO 12135 standard [27]. For the test, an Instron 8862 machine
For the fracture toughness determination, pre-cracked specimens with dimensions of 10 mm ×
was used; the samples were loaded in three-point bending, with a roller span of 40 mm, and a
10 mm × 55 mm were used. Both the pre-crack preparation and the tests were carried out at room
loading rate of 0.1 mm/min was applied. The specimen deflection was measured by means of an
temperature according to the ISO 12135 standard [27]. For the test, an Instron 8862 machine was
inductive transducer integrated directly into the loading axis. In total, five samples were tested for
used; the samples were loaded in three-point bending, with a roller span of 40 mm, and a loading
each investigated condition.
rate of 0.1 mm/min was applied. The specimen deflection was measured by means of an inductive
transducer integrated directly into the loading axis. In total, five samples were tested for each
investigated condition.
Metals 2018, 8, 1047 4 of 15

Macro-hardness measurements were completed using the Vickers (HV10) method.


Ten measurements were made for the metallographic specimens processed with any heat-treatment
parameters, and both the mean values and standard deviations were then calculated.
The material microstructure was evaluated using optical and scanning electron microscopy.
The morphology of the broken test specimens was examined using a JEOL JSM 7600F scanning
electron microscope (SEM, Jeol Ltd., Tokyo, Japan). The analysis was coupled with energy dispersive
spectrometry (EDS, Oxford Instruments, plc., High Wycombe, UK), in order to make a clear
distinction between the eutectic (ECs) and secondary carbides (SCs). For further details regarding
the categorization of carbides, check the literature [11,28]. In brief, the ECs contain mainly vanadium
(more than 50 mass%), and much lower amounts of chromium and iron. Alternatively, chromium
and iron (at a ratio of around 1:1) are the most represented elements in SCs. Also, it has been
established recently that the EC-particles are MC-carbides with the SCs, namely M7 C3 [11,14,18].
The population densities of each carbide type have been determined on twenty-five randomly acquired
SEM micrographs, at a magnification of 3000×. The mean values and the standard deviations of the
obtained data were calculated.
X-ray diffraction (XRD) line profiles have been recorded using a Phillips PW 1710 diffractometer
(Philips Analytical B.V., Almelo, The Netherlands), by filtered Coα1,2 characteristic radiation. A range
of two-theta angles of 30–127◦ was considered for the recording of the diffraction peaks. The retained
austenite content, γR , was determined in accordance with the appropriate ASTM standard [29]. Here,
it should be noted that the characteristic peaks of the major solid solutions, which are considered for
the determination of the γR amounts, are often superimposed by peaks of the carbides. Therefore, the
analysis was coupled with the Rietveld refinement of the obtained X-ray spectra, before computing the
γR content.

3. Results and Discussion


Figure 2a is an optical micrograph of the initial (as-received) state of the experimental steel.
The microstructure is composed of matrix and fine carbides, which are uniformly distributed
throughout the matrix. The SEM micrograph (Figure 2b) and corresponding EDS maps of chromium
and vanadium (Figure 2c,d) provide better insight into the as-annealed microstructure of the Vanadis 6
steel. It is shown that three modifications of carbides are presented in the steel. The first type of particle
is presented as fine globular carbides in the ferritic matrix, and is denoted as spheroidized pearlite
(SP). Also, secondary and eutectic carbides (SCs and ECs) appear in the steel microstructure. EDS
maps (Figure 2c,d) clearly delineate that the carbides involved in SP, as well as the coarser particles
(size 1–3 µm), contain a high chromium amount, while the carbides with a size of around 1.5 µm are
vanadium-rich. According to the recently published results [28,30], one can conclude that both the SCs
and particles in SP are M7 C3 -carbides, while the vanadium-rich particles belong to the group of ECs,
and they are of a MC nature.
Typical SEM micrographs showing the microstructure of CHT, and differently SZT specimens
after low-temperature tempering, are presented in Figure 3. The steel contains the martensite, the
retained austenite, eutectic carbides (ECs), secondary carbides (SCs), and small globular carbides after
CHT (Figure 3a). The application of SZTs reduces the retained austenite amount and increases the
number of SGCs, while the numbers of both the ECs and SCs are unaffected. The increase of the SGC
numbers is the maximum for the specimen that was subjected to the SZT at −140 ◦ C, while the other
SZTs gave lower SGCs counts (Figure 3b–d).
Metals 2018, 8, 1047 5 of 15
Metals 2018, 8, x FOR PEER REVIEW 5 of 16

Figure 2. Microstructure of as-received steel: (a) optical micrograph, (b) SEM micrograph, and
corresponding energy dispersive spectrometry (EDS) maps of (c) chromium and (d) vanadium.

Typical SEM micrographs showing the microstructure of CHT, and differently SZT specimens
after low-temperature tempering, are presented in Figure 3. The steel contains the martensite, the
retained austenite, eutectic carbides (ECs), secondary carbides (SCs), and small globular carbides
after CHT (Figure 3a). The application of SZTs reduces the retained austenite amount and increases
the number of SGCs, while the numbers of both the ECs and SCs are unaffected. The increase of the
SGC numbers is the maximum for the specimen that was subjected to the SZT at −140 °C, while the
other SZTs gave lower SGCs counts (Figure 3b–d).
The SEM micrographs of the microstructures after tempering at 530 °C are shown in Figure 4.
This tempering temperature is high enough to evoke the retained austenite transformation, thus, all
of the specimens were almost free of this phase. The CHT steel and the steel after different SZTs
contained tempered martensite, ECs, SCs, and SGCs. The increased tempering temperature
promotes
Figurea2.
Figure 2.decrease in theof
Microstructure
Microstructure ofSGC numbers,
as-received
as-received steel:while
steel: (a) the numbers
(a) optical
optical micrograph,
micrograph, of (b)
both
(b) SEM
SEMthemicrograph,
ECs and and
micrograph, SCs are
and
unaffected.
corresponding
corresponding energy dispersive spectrometry
energy dispersive spectrometry (EDS)
(EDS) maps
maps ofof (c)
(c) chromium
chromium andand (d)
(d) vanadium.
vanadium.

Typical SEM micrographs showing the microstructure of CHT, and differently SZT specimens
after low-temperature tempering, are presented in Figure 3. The steel contains the martensite, the
retained austenite, eutectic carbides (ECs), secondary carbides (SCs), and small globular carbides
after CHT (Figure 3a). The application of SZTs reduces the retained austenite amount and increases
the number of SGCs, while the numbers of both the ECs and SCs are unaffected. The increase of the
SGC numbers is the maximum for the specimen that was subjected to the SZT at −140 °C, while the
other SZTs gave lower SGCs counts (Figure 3b–d).
The SEM micrographs of the microstructures after tempering at 530 °C are shown in Figure 4.
This tempering temperature is high enough to evoke the retained austenite transformation, thus, all
of the specimens were almost free of this phase. The CHT steel and the steel after different SZTs
contained tempered martensite, ECs, SCs, and SGCs. The increased tempering temperature
promotes a decrease in the SGC numbers, while the numbers of both the ECs and SCs are
unaffected.

Figure 3. Typical SEM micrographs showing the microstructure of specimens after tempering at 170 ◦ C:
(a) CHT, (b) SZT at −140 ◦ C, (c) SZT at −196 ◦ C, and (d) SZT at −269 ◦ C. γR —retained austenite;
ECs—eutectic carbides; SCs—secondary carbides; SGCs—small globular carbides.

The SEM micrographs of the microstructures after tempering at 530 ◦ C are shown in Figure 4.
This tempering temperature is high enough to evoke the retained austenite transformation, thus, all of
the specimens were almost free of this phase. The CHT steel and the steel after different SZTs contained
tempered martensite, ECs, SCs, and SGCs. The increased tempering temperature promotes a decrease
in the SGC numbers, while the numbers of both the ECs and SCs are unaffected.
Metals 2018, 8, x FOR PEER REVIEW 6 of 16
Metals 2018, 8, x FOR PEER REVIEW 6 of 16
Figure 3. Typical SEM micrographs showing the microstructure of specimens after tempering at 170
Figure
°C: 3. Typical
(a) CHT, SEMatmicrographs
(b) SZT −140 °C, (c) showing the °C,
SZT at −196 microstructure
and (d) SZTofatspecimens
−269 °C. γafter tempering
R—retained at 170
austenite;
°C: (a)8,CHT,
1047 (b)
MetalsECs—eutectic
2018, SZT at −140 °C, (c) SZT at −196 °C, and (d) SZT at −269 °C.
carbides; SCs—secondary carbides; SGCs—small globular carbides. γR—retained austenite;6 of 15
ECs—eutectic carbides; SCs—secondary carbides; SGCs—small globular carbides.

Figure 4. SEM micrographs showing the microstructure of the specimens after tempering at 530 °C:
Figure
Figure
(a) 4. SEM
4.
CHT, micrographs
(b) SZT at −140 °C,showing
(c) SZTtheat microstructure
−196 °C, and (d)of the
SZTspecimens
at −269 °C. after tempering
tempering at
γR—retained ◦ C:
530 °C:
austenite;
(a) CHT, (b) SZT
CHT, (b) SZT at −140
at −140 ◦
°C, (c) SZT at
C, (c) SZT at − −196 ◦°C, and
196 C, and (d) SZT at −269
−269 C.
(d) SZT atglobular ◦
°C. γ —retained
γR —retained austenite;
austenite;
ECs—eutectic carbides; SCs—secondary carbides; SGCs—small carbides.
R The decrease in
ECs—eutectic
the carbides;
carbides;
SGC numbers, SCs—secondary
SCs—secondary
in comparison carbides;
carbides;
with the SGCs—small
SGCs—small
specimens atglobular
temperedglobular
170 °C, iscarbides.
carbides. TheThe
obvious. decrease
decrease in
in the
the
SGCSGC numbers,
numbers, in comparison
in comparison withwith the specimens
the specimens tempered
tempered at ◦170
at 170 °C,obvious.
C, is is obvious.
The results of the quantitative analysis of the SGCs are summarized in Figure 5, where it is
obvious results
The that
results ofofthe
treatmentthequantitative
at −140 °Canalysis
quantitative of the
analysis
produces of SGCs
the the are population
SGCs
highest summarized
are summarizedin Figure
density 5,these
where
inofFigure 5, itwhere
is obvious
particles, it
andis
that treatment at − 140 ◦ C produces the highest population density of these particles, and that the
obvious that treatment at −140 °C produces the highest population density
that the enhanced population density of the SGCs (compared to the state after CHT) is retained of these particles, and
enhanced
that
after the population
enhanced
tempering. density of density
population the SGCsof(compared
the SGCsto(compared
the state after CHT)
to the is retained
state after CHT) afteristempering.
retained
Figure
after Figure 66 shows
tempering. the tempering diagram of the CHT steel as well as of the steel
shows the tempering diagram of the CHT steel as well as of the steel after different after different SZTs.
It is visible
SZTs.Figure that the
6 shows
It is visible SZTs improve
thethe
that temperingthe hardness
diagram
SZTs improve over the
theofhardness conventional
the CHTover steelthe heat
as well treatment.
as of theheat
conventional The hardness
steeltreatment.
after different
Theof
the CHT
SZTs.
hardness material
It is
ofvisible
the CHT was
that 875
the HV
material 10,improve
SZTs
was and
875 it
HVwas
10,933,
the 942,
hardness
and and
it was 904the
over
933, HVconventional
942, 10 for
and the
904 HVspecimens
heat
10 theafter
fortreatment. SZTThe
specimens at
− 140, − 196, and − 269 ◦ C, respectively.
hardness of the CHT material was 875
after SZT at −140, −196, and −269 °C, respectively. HV 10, and it was 933, 942, and 904 HV 10 for the specimens
after SZT at −140, −196, and −269 °C, respectively.

Figure 5.
5. Population
Population density
densityof ofSGCs
SGCsininCHT
CHTandand different
different SZT
SZT specimens
specimens after
after tempering
tempering ◦ C:
at 170
at 170
Figure
°C:CHT,
(a) 5. Population
(a) CHT, (b) SZT
(b) SZT density
at −at
140 ◦
−140 of
C,°C, SGCs
(c) (c)
SZTSZT in
at − CHT
at196 ◦ and
−196C,°C,
and different
and
(d)(d)
SZT SZT
SZT
at at specimens

−269 C.
−269 °C. after tempering at 170
°C: (a) CHT, (b) SZT at −140 °C, (c) SZT at −196 °C, and (d) SZT at −269 °C.
Metals 2018, 8, 1047 7 of 15
Metals 2018, 8, x FOR PEER REVIEW 7 of 16

Figure 6. Hardness
Figure 6. Hardness of
ofCHT
CHTand
anddifferently
differentlySZT
SZTspecimens
specimensmade
madeofof
Vanadis 6 steel
Vanadis in dependence
6 steel on
in dependence
tempering temperature.
on tempering temperature.

Conventionally heat-treated steel


Conventionally heat-treated steel manifests
manifests an an evident
evident hardness
hardness decrease
decrease within
within the the
low-temperature tempering range, but the material hardness remains
low-temperature tempering range, but the material hardness remains almost constant within the almost constant within the
high-temperature
high-temperature tempering tempering range,
range, at at aa level
level of of 750
750 HVHV 10. The application
10. The application of of SZT
SZT increases
increases the the
material ◦ C; the
material hardness within the low-temperature tempering range, up to a temperature of 450
hardness within the low-temperature tempering range, up to a temperature of 450 °C; the
highest bulk hardness has been recorded for the steel processed in cold nitrogen gas, at − 140 ◦ C.
highest bulk hardness has been recorded for the steel processed in cold nitrogen gas, at −140 °C.
Tempering
Tempering withinwithin the the secondary
secondary hardening
hardening temperature
temperature range, range, however,
however, decreases
decreases the the hardness
hardness
significantly, and except for SZT at − 140 ◦ C, its values are somewhat lower than what is obtained
significantly, and except for SZT at −140 °C, its values are somewhat lower than what is obtained by
by
CHT.CHT.
For
For the thedemonstration
demonstration of theofflexural strength variations,
the flexural low- (at 170 ◦low-
strength variations, C) and(athigh-temperature
170 °C) and
tempered steel specimens (at 530 ◦ C) were selected. The reason for this is that cold-work tool
high-temperature tempered steel specimens (at 530 °C) were selected. The reason for this is steel
that
grades are commonly used in the state after tempering to either
cold-work tool steel grades are commonly used in the state after tempering to either “primary“primary hardness” (tempering at
low temperatures) or “secondary hardness” (tempering at high temperatures, usually around 500 ◦ C).
hardness” (tempering at low temperatures) or “secondary hardness” (tempering at high
The graphical usually
temperatures, interpretations
around 500 of the
°C).properties
The graphical are in Figure 7 (flexural
interpretations of thestrength)
propertiesandarein in Figure
Figure 87
(work of fracture). The results obtained after tempering at 170 ◦ C infer that the flexural strength
(flexural strength) and in Figure 8 (work of fracture). The results obtained after tempering at 170 °C
is improved
infer that thebecause
flexural of the application
strength is improved of SZTs
becauseat −of140
theand −269 ◦ C,
application of but
SZTs it at
is −140
deteriorated
and −269after °C,
SZT − ◦ W
but it is deteriorated after SZT at −196 °C. The corresponding work of fracture values, Wof, values
at 196 C. The corresponding work of fracture values, of , follow the flexural strength follow
well; the higher
the flexural the flexural
strength values strength,
well; the the morethe
higher energy is required
flexural strength,for specimen
the more energy fracture. The results
is required for
obtained after tempering at 530 ◦ C indicate that the flexural strength is almost unaffected by SZTs
specimen fracture. The results obtained after tempering at 530 °C indicate that the flexural strength
at −140 and
is almost −196 ◦ C.by
unaffected It SZTs
is, however,
at −140improved
and −196 slightly
°C. It is,for SZT at the
however, boilingslightly
improved temperature
for SZT of liquid
at the
helium.
boiling temperature of liquid helium. Compared to the state after CHT, the obtained results ofonly
Compared to the state after CHT, the obtained results of the work of fracture manifest the
weak variations after SZT at − 140 and − 196 ◦ C. Alternatively, the work of fracture is significantly
work of fracture manifest only weak variations after SZT at −140 and −196 °C. Alternatively, the
higher forfracture
the material that was subjected to the at −269 ◦ C, which corresponds well to the flexural
work of is significantly higher for theSZT
material that was subjected to the SZT at −269 °C,
strength changes.
which corresponds well to the flexural strength changes.
Figure 9a–d shows an overview of the typical fracture surfaces of the specimens that were
tempered at 170 ◦ C after CHT, and after SZT at different temperatures. The fracture surface of the
CHT specimen is shown in the SEM micrograph in Figure 9a. It appears smooth and shiny, with
very little surface roughness. This is typical for hard and brittle metallic materials, for instance, for
high-carbon high chromium ledeburitic die steels with a very high hardness, ball bearing steels, and
so on. The fracture surfaces of the specimens that were subjected to different SZTs manifest a clearly
evident topography as they appear rougher (Figure 9b–d). By comparing the SEM fractographs with
the obtained values of both the flexural strength and the work of fracture, it is evident that there is
close relationship between these characteristics, except for the specimen that was subjected to SZT at
−196 ◦ C.
Metals 2018,
Metals 8, x1047
2018, 8, FOR PEER REVIEW 8 8of
of 16
15
Metals 2018, 8, x FOR PEER REVIEW 8 of 16

Figure 7. Flexural strength for CHT and differently SZT treated specimens made of Vanadis 6 steel.
Figure 7.
Figure Flexuralstrength
7. Flexural strength for
for CHT
CHT and
and differently
differently SZT
SZT treated
treated specimens
specimens made of Vanadis 6 steel.

MetalsFigure
2018, 8,8.
x FOR PEER
Work
Work REVIEW
of fracture Wof for
forCHT
CHTand
anddifferently
differentlySZT
SZTtreated
treatedspecimens
specimensmade
madeof
ofVanadis 9 of 16
Vanadis 66 steel.
steel.
Figure 8. Work of fracture Wofoffor CHT and differently SZT treated specimens made of Vanadis 6 steel.

Figure 9a–d shows an overview of the typical fracture surfaces of the specimens that were
Figure 9a–d shows an overview of the typical fracture surfaces of the specimens that were
tempered at 170 °C after CHT, and after SZT at different temperatures. The fracture surface of the
tempered at 170 °C after CHT, and after SZT at different temperatures. The fracture surface of the
CHT specimen is shown in the SEM micrograph in Figure 9a. It appears smooth and shiny, with
CHT specimen is shown in the SEM micrograph in Figure 9a. It appears smooth and shiny, with
very little surface roughness. This is typical for hard and brittle metallic materials, for instance, for
very little surface roughness. This is typical for hard and brittle metallic materials, for instance, for
high-carbon high chromium ledeburitic die steels with a very high hardness, ball bearing steels, and
high-carbon high chromium ledeburitic die steels with a very high hardness, ball bearing steels, and
so on. The fracture surfaces of the specimens that were subjected to different SZTs manifest a
so on. The fracture surfaces of the specimens that were subjected to different SZTs manifest a
clearly evident topography as they appear rougher (Figure 9b–d). By comparing the SEM
clearly evident topography as they appear rougher (Figure 9b–d). By comparing the SEM
fractographs with the obtained values of both the flexural strength and the work of fracture, it is
fractographs with the obtained values of both the flexural strength and the work of fracture, it is
evident that there is close relationship between these characteristics, except for the specimen that
evident that there is close relationship between these characteristics, except for the specimen that
was subjected to SZT at −196 °C.
was subjected to SZT at −196 °C.

Figure 9.
9. SEM
SEMmicrographs
micrographsshowing
showingthethe fracture
fracture surfaces
surfaces of the
of the specimens
specimens for flexural
for flexural strength
strength after
after
tempering at 170at◦ C:
tempering 170(a)
°C: (a) CHT,
CHT, (b)at
(b) SZT SZT−140 ◦ C, (c)
at −140 °C,SZT
(c) SZT ◦ C, °C,
at −196
at −196 andand (d) SZT
(d) SZT at −at269 ◦ C.°C.
−269

However, detailed SEM micrographs (Figures 10a–d) clearly delineate the differences in the
fracture surface morphology of the differently heat-treated specimens. By comparing these
micrographs, it can be seen that the fracture surface roughness decreases in turn for the SZT
specimens at −140 °C, SZT specimens at −269 °C, SZT specimens at −196 °C, and CHT steel. The
MetalsFigure
2018, 8, 9. SEM micrographs showing the fracture surfaces of the specimens for flexural strength9 of 15
1047
after tempering at 170 °C: (a) CHT, (b) SZT at −140 °C, (c) SZT at −196 °C, and (d) SZT at −269 °C.

However,
However,detailed
detailedSEMSEM micrographs
micrographs (Figure 10a–d)
(Figures clearly
10a–d) delineate
clearly the differences
delineate in the fracture
the differences in the
surface morphology of the differently heat-treated specimens. By comparing
fracture surface morphology of the differently heat-treated specimens. By comparing these these micrographs, it
can be seen that at − ◦
micrographs, it the
canfracture
be seensurface
that theroughness
fracture decreases in turn fordecreases
surface roughness the SZT specimens
in turn for 140SZT
the C,
SZT specimens at − 269 ◦ C, SZT specimens at −196 ◦ C, and CHT steel. The detailed micrographs
specimens at −140 °C, SZT specimens at −269 °C, SZT specimens at −196 °C, and CHT steel. The
reveal
detailedthat the area fraction
micrographs revealofthat
the the
cleavage fracture of
area fraction propagation
the cleavage is much more
fracture pronounced
propagation is in the
much
◦ ◦
specimens of SZT at
more pronounced in−the
196specimens
C than what is seen
of SZT for SZT
at −196 at −what
°C than 140 and 269 for
is seen C. The
SZTfracture propagated
at −140 and 269 °C.
mostly in the so-called “low-energy” ductile manner in samples after SZT at − 140 and − 269 ◦ C, while
The fracture propagated mostly in the so-called “low-energy” ductile manner in samples after SZT
the fracture at −196 ◦ manifest a clearly visible
at −140 and surfaces
−269 °C, of the CHT
while materialsurfaces
the fracture and theofsteel
theafter
CHTSZTmaterial andCthe steel after SZT at −196
cleavage
°C manifestpropagation micromechanism.
a clearly visible cleavage propagation micromechanism.

10.SEM
Figure 10. SEMmicrographs
micrographs showing
showing details
details of fracture
of the the fracture
surfacessurfaces
of the of the specimens
specimens for the
for the flexural
flexural strength
strength after tempering at 170 ◦ C:
after tempering at (a)
170CHT,
°C: (a)
(b)CHT,
SZT at 140 ◦atC,−140
(b)−SZT °C, at
(c) SZT (c)−SZT ◦ C,
196 at −196
and°C,
(d)and
SZT(d)
at

−269 C. DC—decohesion at the matrix/carbide interface; FCs—fractured carbides; PLD—plastic
SZT at −269 °C. DC—decohesion at the matrix/carbide interface; FCs—fractured carbides;
PLD—plastic CL—cleavage
deformation; deformation; CL—cleavage
fracture mode.fracture mode.

Detailed SEM micrographs


micrographsofofthe the specimens tempered at 530 ◦ C (Figure 11a–d) assist to
Detailed SEM specimens tempered at 530 °C (Figure 11a–d) assist to draw
draw the conclusion
the conclusion that fracture
that the the fracture surfaces
surfaces of CHT
of CHT and,
and, differently,SZT
differently, SZTsteel,
steel,do
do not
not manifest
manifest
significant differences in terms of their topography. All of the surfaces contain a great number
of micro-dimples and holes, which correspond to the extraction of small globular carbides (SGCs)
from the fracture surface during the crack propagation. This phenomenon becomes more clearly
evident in the SZT specimens, as they contain a much higher amount and population density of these
particles [10,11,14,18]. Besides that, micro-voids are also formed at the other carbide/matrix interfaces,
however, coarser particles more easily undergo cleavage failure rather than a decohesive fracture
propagation manner. This is due to the fact that coarser carbides are mostly M7 C3 (average size of
2.5 µm), while the finer ones are MC (average size of 1.6 µm) [14,22]. The size effect of carbides on the
fracture propagation manner has been investigated extensively. It has been established that the larger
size of M7 C3 carbides makes them very prone to cracking [17,31–35]. Also, the crystallography of the
carbides should be taken into consideration when assessing their role in fracture behavior. The MC
carbides are cubic, while the M7 C3 carbides are hexagonal. Casellas et al. [36] and Lin et al. [37],
for instance, studied the fracture toughness of the typical carbides that occur in cold work tool
steels. They determined the fracture toughness of the MC phase to be 3.7 ± 0.6 MPa·m1/2 . For the
M7 C3 -carbides, the fracture toughness varied over a wide range, from 0.5 to 4 MPa·m1/2 , whereas the
lowest values correspond to the orientation parallel of the carbide’s main axis; meanwhile, the higher
data was acquired in the orientation perpendicular to the carbide’s axis. One can thus summarize that
Also, the crystallography of the carbides should be taken into consideration when assessing their
role in fracture behavior. The MC carbides are cubic, while the M7C3 carbides are hexagonal.
Casellas et al. [36] and Lin et al. [37], for instance, studied the fracture toughness of the typical
carbides that occur in cold work tool steels. They determined the fracture toughness of the MC
phase to be
Metals 2018, 3.7 ± 0.6 MPa·m1/2. For the M7C3-carbides, the fracture toughness varied over a10wide
8, 1047 of 15
range, from 0.5 to 4 MPa·m1/2, whereas the lowest values correspond to the orientation parallel of
the carbide’s main axis; meanwhile, the higher data was acquired in the orientation perpendicular
M7the
to C3 -carbides
carbide’sare more
axis. Onebrittle thansummarize
can thus MC, despitethat
their
M7lower hardness
C3-carbides are(1800
moreHV vs. 2500–3000
brittle HV for
than MC, despite
the MC).
their lower hardness (1800 HV vs. 2500–3000 HV for the MC).

Figure 11.
11. SEM
SEMmicrographs
micrographs showing
showing details
details of fracture
of fracture surfaces
surfaces of theof the broken
broken specimens
specimens for
for flexural
flexural
strengthstrength
after tempering at 530 ◦ C:
after tempering at (a)
530CHT,
°C: (a)
(b)CHT,
SZT at 140 ◦atC,−140
(b)−SZT °C, (c)
(c) SZT at −SZT ◦ C,
196 at −196
and°C,
(d)and
SZT(d)
at
−269at◦ C.
SZT −269 °C.

above-mentioned findings
The above-mentioned findingsare areininclose
closecorrelation
correlationtotothe
the microstructural
microstructural characteristics.
characteristics. It
It is shown in Figures 3–5 that the application of SZT produces a considerably
is shown in Figures 3–5 that the application of SZT produces a considerably enhanced number and enhanced number
and population
population density
density of small
of small globularglobular
carbidescarbides
(SGCs),(SGCs),
and that and
thisthat this material
material state isretained
state is largely largely
retained
after after tempering.
tempering. Thealso
The steel steelcontains
also contains eutectic
eutectic and and secondary
secondary carbides
carbides afterthe
after theapplied
applied heat
treatment schedule. However,
However, the population
population density
density of these particles is unaffected by quenching
and/or quenching
and/or quenching followed
followed by by SZT
SZT [11,14,18]. The The higher
higher number
number of of carbides
carbides produces
produces more
matrix/carbide interfaces. During crack propagation, the rigid carbides cannot co-deform withwith
matrix/carbide interfaces. During crack propagation, the rigid carbides cannot co-deform the
the matrix.
matrix. Consequently,
Consequently, they
they cancan either
either crack
crack byby cleavage,ororassist
cleavage, assisttotodecohesion
decohesionalways
alwaysfollowed
followed by
a ductile microvoid coalescence fracture micromechanism. The higher the matrix/carbide interfaces
number, the higher is the probability to form microvoids at these interfaces.
Another issue is the retained austenite amount. In the current investigation, the maximum
retained austenite amount was found in the CHT material, and it was significantly reduced due to
the application of SZT (Figure 12). Hence, one can expect that the CHT material would manifest a
better flexural strength than what is obtained by the SZTs. However, SZT produces much more carbide
particles, and thereby much more matrix/carbide interfaces, which, on the contrary, may act in favor
of the better flexural strength of the SZT material.
The refinement of the martensitic domains was first recorded for 12% Cr–4% V ledeburitic steel
by Tyshchenko et al. [12]. A plausible explanation for this phenomenon was delineated later by
Jurči et al. [11]. He suggested dividing the γ to α’ transformation into two components, namely the
diffusion-less (athermal) component and a time-dependent isothermal component. This is active
during the steel holding at a SZT temperature, and is always connected with the extensive plastic
deformation of newly formed “virgin” martensite, with the capture of carbon atoms by gliding
dislocations, and thereby with mass transfer. The mass transfer, albeit in a limited extent, is associated
with controlling the martensitic plates growth. In addition, the martensitic domains grow freely
Metals 2018, 8, x FOR PEER REVIEW 11 of 16

a ductile microvoid coalescence fracture micromechanism. The higher the matrix/carbide interfaces
number, the higher is the probability to form microvoids at these interfaces.
Metals 2018, 8, 1047 11 of 15
Another issue is the retained austenite amount. In the current investigation, the maximum
retained austenite amount was found in the CHT material, and it was significantly reduced due to
the application
within the primary of SZT (Figuregrains
austenite 12). Hence, one can expect
at the beginning of thethat the CHT material
transformation. would manifest
Alternatively, the space a
better flexural
for their growthstrength than what
is considerably is obtained
limited during by
the the SZTs. However,
isothermal SZT
hold at the produces much as
cryo-temperature, more
the
carbide
much spaceparticles, and thereby
is already filled bymuch more matrix/carbide
the martensite interfaces,
formed athermally. Thewhich,
overallon the contrary,
refinement of themay act
matrix
in favor of the better
microstructure is thenflexural strengthof
a consequence ofthe
theabove
SZT material.
mentioned two phenomena.

◦ C:
12. Retained
Figure 12.
Figure Retained austenite
austeniteamount
amountininCHT
CHTandanddifferently SZT
differently SZTspecimens after
specimens tempering
after at 170
tempering at 170
(a) at −at ◦ at − ◦ −269 ◦
°C:CHT, (b) SZT
(a) CHT, (b) SZT 140
−140C,°C,
(c) (c)
SZTSZT at196
−196C,°C,
and (d)(d)
and SZT at at
SZT −269 C.
°C.

The refinement
The grain refinement
of the has beneficial
martensitic a effectwas
domains on both the strength
first recorded and toughness
for 12% of the metallic
Cr–4% V ledeburitic steel
materials. This mechanism is known as “grain boundary strengthening”, and can be
by Tyshchenko et al. [12]. A plausible explanation for this phenomenon was delineated later expressed by the
by
Hall-Petch equation [38], as follows:
Jurči et al. [11]. He suggested dividing the γ to α’ transformation into two components, namely the
σo + k ∗ d1/2
diffusion-less (athermal) component andσa=time-dependent isothermal component. This is active(1)
duringσthe
where steel holding at a SZT temperature, and is always connected with the extensive plastic
o is a friction stress (it involves contributions from both the solutes and foreign particles); k is
deformation
the Hall–Petchofconstant
newly (specific
formed “virgin” martensite,
for each material); andwith thegrain
d is the capture
size,of carbon
which atoms represented
is usually by gliding
dislocations, and thereby with mass transfer. The mass transfer, albeit in
by the average grain diameter. In the particular case of martensitic structures (non-polyhedral), a limited extent,the
is
associated of
parameter with
graincontrolling
size d cannotthebemartensitic plates
taken into the growth. In Instead,
consideration. addition,thethe martensiticdimension
characteristic domains
grow freely within the primary austenite
d represents the width of the martensitic domains. grains at the beginning of the transformation.
Alternatively,
One can thus the summarize
space for theirthatgrowth is considerably
the better flexural strengthlimited during
of SZT the is
material isothermal
due to thehold at the
combined
cryo-temperature,
effect of the overallas the much space
microstructure is already
refinement andfilled by the martensite
the enhanced populationformed
densityathermally. The
of the carbides,
overall refinement of the matrix microstructure is then a consequence of the above
despite the fact that the hardness is higher and the retained austenite is significantly reduced. mentioned two
phenomena.
In tempering within the common secondary hardening temperature range, the role of the
The grain refinement
retained austenite has beneficialas
can be disregarded, a effect on both
the steel the strength
is almost free of and
this toughness of the metallic
phase [14,18,39]. Hence,
materials. This mechanism is known as “grain boundary strengthening”,
the slightly improved or almost equal flexural strength of SZT steel (as compared with the and can be expressed by
state after
the Hall-Petch equation [38], as follows:
CHT) can be attributed only to the higher population density of the SGCs, and to the martensite
refinement. The presence of a higher population density
σ = σo + k ∗ d 1/ 2 of SGCs essentially contributes to (1)
decohesive fracture propagation manner (see also the discussion to the state after low-temperature
the

tempering).
where Alternatively,
σo is a friction stress (ithigh-temperature tempering
involves contributions induces
from both the extensive
the solutes precipitation
and foreign particles);ofk
nano-sized
is carbides
the Hall–Petch [14,18].(specific
constant These carbides
for eacharematerial);
coherent and
withdthe matrix,
is the grainand make
size, it more
which brittle.
is usually
Therefore,
representedone bycantheconclude
average that thediameter.
grain final flexural strength
In the of thecase
particular steelofis always a result
martensitic of the
structures
competition between
(non-polyhedral), the the precipitation
parameter of grainstrengthening
size d cannotmechanism of carbides,
be taken into and the mechanisms
the consideration. Instead, the
affecting microplastic
characteristic dimension performance
d represents ofthe
thewidth
matrixof(because of the enhanced
the martensitic domains. number of SGCs and the
overall
Onestructural
can thus refinement).
summarize that the better flexural strength of SZT material is due to the
combined
Figureeffect of thethe
13 shows overall microstructure
variations refinement
of the fracture and the
toughness of enhanced
the low- and population density of
high-temperature
tempered specimens that were CHT and SZT at different temperatures.
Therefore, one can conclude that the final flexural strength of the steel is always a result of the
competition between the precipitation strengthening mechanism of carbides, and the mechanisms
affecting microplastic performance of the matrix (because of the enhanced number of SGCs and the
overall structural refinement).
Figure 13 shows the variations of the fracture toughness of the low- and high-temperature
Metals 2018, 8, 1047 12 of 15
tempered specimens that were CHT and SZT at different temperatures.

Figure 13.
Figure Fracture toughness
13. Fracture toughness for
for differently
differently heat-treated
heat-treated specimens
specimens made
made of
of Vanadis
Vanadis 66 steel.
steel.

In CHT KICKIC ◦ C than what


In CHT steel,
steel,the
thefracture
fracturetoughness
toughness is comparably
is comparablyhigher afterafter
higher tempering at 170
tempering at 170 °C than
was achieved after tempering at 530 ◦ C. The SZTs at −140 and −269 ◦ C influence the fracture toughness
what was achieved after tempering at 530 °C. The SZTs at −140 and −269 °C influence the fracture
only marginally
toughness only after low-temperature
marginally tempering, buttempering,
after low-temperature improve it but
substantially
improve after tempering inafter
it substantially the
tempering in the common normal secondary hardening temperature range. In addition, there isof
common normal secondary hardening temperature range. In addition, there is a detrimental effect a
SZT at − 196 ◦ C after low-temperature tempering, while a slightly positive effect for a high-temperature
detrimental effect of SZT at −196 °C after low-temperature tempering, while a slightly positive
tempered
effect for astate is observed. tempered state is observed.
high-temperature
The observed
The observedvariations
variations of of
the the
fracture toughness
fracture with an
toughness aggregate
with effect of effect
an aggregate SZT andof tempering
SZT and
can be referred to by the following factors:
tempering can be referred to by the following factors:
-- The higher
The higher retained
retained austenite
austenite amount
amount acts
acts in
in favor
favor of
of better KICIC, ,as
better K asreported
reportedrecently
recently[15,22,38].
[15,22,38].
ItIt is
isshown
shown(Figure 12) that
(Figure 12) the
thatretained austeniteaustenite
the retained amount decreases
amount substantially with substantially.
decreases substantially with
- The application
substantially. of SZT produces a considerably higher amount and population density of
- cementite
The particles
application (of size
of SZT 100–500a nm,
produces small globular
considerably highercarbides
amount(SGCs)), whereas the
and population SZT of
density at
−140 ◦ C acts
cementite more effectively
particles in this way
(of size 100–500 nm, [11,14,15,18,40]. Tempering
small globular carbides alwayswhereas
(SGCs)), reduces the
the amount
SZT at
of these
−140 °C carbides.
acts moreDespite that,in
effectively however,
this way it remains much higher
[11,14,15,18,40]. than what
Tempering can reduces
always be obtainedthe
by CHTof
amount (Figure
these 5). The crack
carbides. propagates
Despite by a decohesive
that, however, it remainsmechanism
much higher at the
thancarbide/matrix
what can be
interfacesby
obtained [15,22],
CHTwhich is connected
(Figure with micro-plastic
5). The crack propagates by deformation.
a decohesive The mechanism
more carbides,at the
higher is the probability to form the microvoids at the interfaces, and the higher the deformation
energy is that is needed for the fracture propagation.
- The precipitation of carbides makes the matrix stiffer and less amenable to deform plastically.
In particular, the M7 C3 carbides that precipitate at high tempering temperatures [14,18] enhance
the matrix hardness substantially.

In the case of low-temperature tempering, the KIC values were almost the same for the specimens
after CHT and after SZT at −140 ◦ C; the differences in the γR amount are compensated by variations
in the number and population density of SGCs. The SZT at −196 ◦ C further reduces the γR amount,
and produces lower amounts of SGCs—this is why a low KIC level has been recorded.
As above mentioned, the role of the γR can be disregarded in the case of tempering into the
common secondary hardening temperature range. The resulting KIC level is then a function of the
number and population density of SGCs; this is the highest one after SZT at −140 ◦ C. Also, the
variations in the extent of the carbides precipitation should be considered; SZT at −196 ◦ C suppresses
the precipitation of M7 C3 , which results in a lower matrix hardness, and thereby in a slightly higher
KIC than what is obtained after CHT.
Metals 2018, 8, 1047 13 of 15

Finally, SZT at −140 ◦ C lowers neither the flexural strength of the steel (Figures 7 and 8) nor its
fracture toughness (Figure 13) in the as-low-temperature tempered state, at a significantly enhanced
hardness. In addition, this kind of treatment improves all of these mechanical properties after
tempering within the common secondary hardening temperature range. It has been demonstrated
recently [39] that SZTs at −196 ◦ C do not act in favor of simultaneous improvement of hardness
and toughness after low-temperature tempering, while these properties are slightly improved after
high temperature tempering. Therefore, the treatment at −140 ◦ C enhances the hardness along with
the toughness to a much more remarkable extent, as compared to the state after conventional heat
treatment. The treatment in liquid helium seems to also be a promising way how to improve the
hardness and toughness simultaneously. The temperature of the treatment is very low, however,
and practically there is limited transport of atoms possible at 4 K. Hence, all of the processes being
responsible for the ameliorations of mechanical properties inevitably take a very long time, which is
unacceptable from the point of view of the overall economy of the treatment, due to the high price of
helium, among others.

4. Conclusions
The results obtained by the systematic investigations of the effect of the sub-zero treatments
at different temperatures (−140, −196, and −269 ◦ C), and the tempering on the hardness, fracture
toughness, and flexural strength, allow for making the following main conclusions:
(i) The hardness is generally improved by the SZTs applied; this improvement is the most
pronounced for SZT at −140 ◦ C, while other SZT conditions gave a less remarkable hardness
increase. The observed improvement has been assigned to the increased volume fraction of
small globular cementite carbides, because of the microstructural changes that occurred at
subzero temperatures.
(ii) The fracture toughness level is generally unaffected or somewhat worsened for the material after
low-temperature tempering. In high-temperature tempered samples (in secondary hardening
range), an improvement of the fracture toughness has been recorded. A higher number of small
globular cementite particles increases the number of nucleation sites, by decohesion particles to
the matrix interface boundary. On the other hand, the microstructure refinement contributes to
the higher fracture resistance of the matrix. As a consequence, the potential for fracture toughness
enhancement is not too high, but, under improvement of the hardness, there is no risk for a
toughness decrease.
(iii) The flexural strength is slightly improved by SZTs, except for the case of SZT in liquid nitrogen,
where no effect or a slight worsening has been recorded. Although changes in the flexural
strength with applied SZTs conditions are analogous to the hardness changes, there is another
aspect going against the better improvement. These are associated with increased number of
decohesion sites at the carbide-matrix interface. contributing to a low energy ductile fracture.
(iv) The SZTs at −140 and −269 ◦ C make it possible to obtain the simultaneous improvement of both
the hardness and toughness in a much greater extent than the SZTs carried out at other treatment
temperatures. However, the use of the boiling temperature of liquid helium may be associated
with excessive production costs.

Author Contributions: Writing of manuscript, P.J. and I.D.; methods of research, P.J. and I.D.; realization of heat
treatments, P.P. and Z.M.; microstructural investigations, P.J. and I.D.; investigations of mechanical properties,
I.D.; project administration, P.J.; funding acquisition, P.J.
Funding: This research was funded by the Scientific Grant Agency of the Ministry of Education, Science, Research,
and Sport of the Slovak Republic and the Slovak Academy of Sciences, under project no. 1/0264/17; the Scientific
Grant Agency of the Ministry of Education, Science, Research, and Sport of the Slovak Republic and the Slovak
Academy of Sciences.
Acknowledgments: The authors acknowledge that the paper is a result of the experiments realized within the
project VEGA 1/0264/17. In addition, this publication is the result of the project implementation “Centre for
Metals 2018, 8, 1047 14 of 15

Development and Application of Advanced Diagnostic Methods in Processing of Metallic and Non-Metallic
Materials—APRODIMET”, ITMS: 26220120014, supported by the Research and Development Operational
Programme funded by the ERDF.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Stratton, P.F. Optimising nano-carbide precipitation in tool steels. Mater. Sci. Eng. 2007, A449–451, 809–812.
[CrossRef]
2. Reitz, W.; Pendray, J. Cryoprocessing of Materials: A Review of Current Status. Mater. Manuf. Proc. 2001, 16,
829–840. [CrossRef]
3. Sweeney, T.P. Deep cryogenics: the great cold debate. J. Heat. Treat. 1986, 2, 28–33.
4. Yugandhar, T.; Krishnan, P.K.; Bhaskar Rao, C.V.; Kalidas, R. Cryogenic Treatment and It’s Effect on Tool
Steel. In Proceedings of the 6th International Tooling Conference, Karlstad, Sweden, 10–13 September 2002;
Bergstrom, J., Fredriksson, G., Johansson, M., Kotik, O., Thuvander, F., Eds.; Karlstad University: Karlstad,
Sweden, 2012; pp. 671–684.
5. Das, D.; Dutta, A.K.; Ray, K.K. On the enhancement of wear resistance of tool steels by cryogenic treatment.
Philos. Mag. Lett. 2008, 88, 801–811. [CrossRef]
6. Das, D.; Ray, K.K. Structure-property correlation of sub-zero treated AISI D2 steel. Mater. Sci. Eng. A 2012,
541, 45–60. [CrossRef]
7. Akhbarizadeh, A.; Golozar, M.A.; Shafeie, A.; Kholghy, M. Effects of austenitizing time on wear behaviour
of D6 tool steel after deep cryogenic treatment. J. Iron Steel Res. 2009, 16, 29–32. [CrossRef]
8. Collins, D.N. Deep cryogenic treatment of tool steels—A review. Heat Treat. Met. 1996, 23, 40–42.
9. Das, D.; Dutta, A.K.; Ray, K.K. Sub-zero treatments of AISI D2 steel: Part I. Microstructure and hardness.
Mater. Sci. Eng. A 2010, 527, 2182–2193. [CrossRef]
10. Jurči, P.; Kusý, M.; Ptačinová, J.; Kuracina, V.; Priknerová, P. Long-term Sub-zero Treatment of P/M Vanadis
6 Ledeburitic Tool Steel—A Preliminary Study. Manuf. Technol. 2015, 15, 41–47.
11. Jurči, P.; Dománková, M.; Čaplovič, L.; Ptačinová, J.; Sobotová, J.; Salabová, P.; Prikner, O.; Šuštaršič, B.;
Jenko, D. Microstructure and hardness of sub-zero treated and no tempered P/M Vanadis 6 ledeburitic tool
steel. Vacuum 2015, 111, 92–101. [CrossRef]
12. Tyshchenko, A.I.; Theisen, W.; Oppenkowski, A.; Siebert, S.; Razumov, O.N.; Skoblik, A.P.; Sirosh, V.A.;
Petrov, J.N.; Gavriljuk, V.G. Low-temperature martensitic transformation and deep cryogenic treatment of a
tool steel. Mater. Sci. Eng. A 2010, 527, 7027–7039. [CrossRef]
13. Meng, F.; Tagashira, K.; Azuma, R.; Sohma, H. Role of Eta-carbide Precipitation’s in the Wear Resistance
Improvements of Fe-12Cr-Mo-V-1.4C Tool Steel by Cryogenic Treatment. ISIJ Int. 1994, 34, 205–210.
[CrossRef]
14. Jurči, P.; Dománková, M.; Hudáková, M.; Ptačinová, J.; Pašák, M.; Palček, P. Characterization of
microstructure and tempering response of conventionally quenched, short- and long-time sub-zero treated
PM Vanadis 6 ledeburitic tool steel. Mater. Charact. 2017, 134, 398–415. [CrossRef]
15. Sobotová, J.; Jurči, P.; Dlouhý, I. The effect of sub-zero treatment on microstructure, fracture toughness, and
wear resistance of Vanadis 6 tool steel. Mater. Sci. Eng. A 2016, 652, 192–204. [CrossRef]
16. Collins, D.N.; Dormer, J. Deep Cryogenic Treatment of a D2 Cold-Work Tool Steel. Heat Treat. Met. 1997, 24,
71–74.
17. Das, D.; Sarkar, R.; Dutta, A.K.; Ray, K.K. Influence of sub-zero treatments on fracture toughness of AISI D2
steel. Mater. Sci. Eng. A 2010, 528, 589–603. [CrossRef]
18. Jurči, P.; Dománková, M.; Ptačinová, J.; Pašák, M.; Kusý, M.; Priknerová, P. Investigation of the
Microstructural Changes and Hardness Variations of Sub-Zero Treated Cr-V Ledeburitic Tool Steel Due to
the Tempering Treatment. J. Mater. Eng. Perform. 2018, 27, 1514–1529. [CrossRef]
19. Surberg, C.H.; Stratton, P.; Lingenhole, K. The effect of some heat treatment parameters on the dimensional
stability of AISI D2. Cryogenics 2008, 48, 42–47. [CrossRef]
20. Molinari, A.; Pellizzari, M.; Gialanella, S.; Straffelini, G.; Stiasny, K.H. Effect of deep cryogenic treatment on
mechanical properties of tool steels. J. Mater. Proc. Technol. 2001, 118, 350–355. [CrossRef]
Metals 2018, 8, 1047 15 of 15

21. Rhyim, Y.M.; Han, S.H.; Na, Y.S.; Lee, J.H. Effect of deep cryogenic treatment on carbide precipitation and
mechanical properties of tool steel. Solid State Phenom. 2006, 118, 9–14. [CrossRef]
22. Ptačinová, J.; Sedlická, V.; Hudáková, M.; Dlouhý, I.; Jurči, P. Microstructure—Toughness relationships in
sub-zero treated and tempered Vanadis 6 steel compared to conventional treatment. Mater. Sci. Eng. A 2017,
702, 241–258. [CrossRef]
23. Ptačinová, J.; Jurči, P.; Dlouhý, I. Fracture toughness of ledeburitic Vanadis 6 steel after sub-zero treatment
for 17 H and double tempering. Mater. Technol. 2017, 51, 729–733. [CrossRef]
24. Gavriljuk, V.G.; Theisen, W.; Sirosh, V.V.; Polshin, E.V.; Kortmann, A.; Mogilny, G.S.; Petrov, Yu.N.;
Tarusin, Y.V. Low-temperature martensitic transformation in tool steels in relation to their deep cryogenic
treatment. Acta Mater. 2013, 61, 1705–1715. [CrossRef]
25. Nemec, M.; Jurči, P.; Kosnáčová, P.; Kučerová, M. Evaluation of structural isotropy of Cr-V ledeburitic steel
made by powder metallurgy of rapidly solidified particles. Kovove Mater. 2016, 54, 453–462. [CrossRef]
26. Jurči, P.; Dlouhý, I. Fracture Behaviour of P/M Cr-V Ledeburitic Steel with Different Surface Roughness.
Mater. Eng.-Materiálové inžinierstvo 2011, 18, 36–43.
27. ISO 12135. Metallic Materials-Unified Method of Test for the Determination of Quasistatic Fracture
Toughness. Available online: https://www.iso.org/standard/60891.html (accessed on 15 November 2016).
28. Jurči, P. Cr-V Ledeburitic cold-work tool steels. Mater. Technol. 2011, 45, 383–394.
29. ASTM E975-13. Standard Practice for X-ray Determination of Retained Austenite in Steel with Near Random
Crystallographic Orientation; ASTM Book of Standards: West Conshohocken, PA, USA, 2004; Volume 3.01.
30. Bílek, P.; Sobotová, J.; Jurči, P. Evaluation of the microstructural changes in Cr-V ledeburitic tool steels
depending on the austenitization temperature. Mater. Technol. 2011, 44, 489–493.
31. Pirtovšek, T.V.; Kugler, G.; Terčelj, M. The behaviour of the carbides of ledeburitic AISI D2 tool steel during
multiple hot deformation cycles. Mater. Charact. 2013, 83, 97–108. [CrossRef]
32. Mostafari, S.; Fotouhi, M.; Motasemi, A.; Ahmadi, M.; Sindi, C.T. Acoustic emission methodology to evaluate
the fracture toughness in heat treated AISI D2 tool steel. J. Mater. Eng. Perform. 2012, 21, 2106–2116.
[CrossRef]
33. Wang, J.; Guo, W.; Sun, H.; Li, H.; Gou, H.; Zhang, J. Plastic deformation behaviors and hardening mechanism
of M7C3 carbide. Mater. Sci. Eng. A 2016, 662, 88–94. [CrossRef]
34. Fukaura, K.; Yokoyama, Y.; Yokoi, D.; Tsuji, N.; Ono, K. Fatigue of cold-work tool steels: effect of heat
treatment and carbide morphology on fatigue crack formation, life, and fracture surface observations.
Metall. Mater. Trans. A 2004, 35, 1289–1300. [CrossRef]
35. Sohar, C.R.; Betzwar-Kotas, A.; Gierl, C.; Weiss, B.; Danninger, H. Fractographic evaluation of gigacycle
fatigue crack nucleation and propagation of a high Cr alloyed cold work tool steel. Int. J. Fatigue 2008, 30,
2191–2199. [CrossRef]
36. Casellas, D.; Caro, J.; Molas, S.; Prado, J.M.; Valls, I. Fracture toughness of carbides in tool steels evaluated by
nanoindentation. Acta Mater. 2007, 55, 4277–4286. [CrossRef]
37. Lin, C.M.; Chang, C.M.; Chen, J.-H.; Wu, W. Hardness, toughness and cracking systems of primary
(Cr, Fe)23C6 and (Cr, Fe)7C3 carbides in high-carbon Cr-based alloys by indentation. Mater. Sci. Eng.
A 2010, 527, 5038–5043. [CrossRef]
38. Hansen, N. Hall-Petch relation and boundary strengthening. Scr. Mater. 2004, 51, 801–806. [CrossRef]
39. Jurči, P.; Ptačinová, J.; Sahul, M.; Dománková, M.; Dlouhý, I. Metallurgical principles of microstructure
formation in sub-zero treated cold-work tool steels—A review. Matériaux Tech. 2018, 106, 104–112. [CrossRef]
40. Ďurica, J.; Ptačinová, J.; Hudáková, M.; Kusý, M.; Jurči, P. Microstructure and Hardness of Cold Work
Vanadis 6 Steel after Subzero Treatment at −140 degrees C. Adv. Mater. Sci. Eng. 2018, 2018, 6537509.
[CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like