Sei sulla pagina 1di 9

Materials Science & Engineering A 624 (2015) 3240

Contents lists available at ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

The effect of deep cryogenic treatments on the mechanical properties


of an AISI H13 steel
Marcos Prez n, Francisco Javier Belzunce
Department of Materials Science and Metallurgic Engineering, Polytechnic School of Engineering, University of Oviedo, Campus Universitario, Gijn, Spain

art ic l e i nf o

a b s t r a c t

Article history:
Received 6 May 2014
Received in revised form
2 October 2014
Accepted 17 November 2014
Available online 25 November 2014

Cryogenic treatments are considered a good way to reduce the retained austenite content and improve
the performance of tool steels. Four different heat treatments, two of which included a deep cryogenic
stage, were applied in this study to an H13 tool steel, subsequently determining its mechanical
properties by means of tensile, hardness and fracture toughness tests. Furthermore, scanning electron
microscopy and X-ray diffraction analysis were performed to gain an insight into the microstructural
evolution of these heat treatments during all the stages.
It was concluded that the application of a deep cryogenic treatment to H13 steel induces higher
thermal stresses and structural defects, producing a dispersed network of ne carbides after the
subsequent tempering stages, which were responsible for a signicant improvement in the fracture
toughness of this steel without modifying other mechanical properties. Although the application of a
deep cryogenic treatment reduces the retained austenite content, there is a minimum innate content
which cannot be transformed by heat treatment. Nevertheless, this austenite is hence believed to be
stable enough and should not transform during the normal service life of forging dies.
& 2014 Elsevier B.V. All rights reserved.

Keywords:
AISI H13 tool steel
Deep cryogenic treatments
Mechanical properties
Microstructural evolution
X-ray diffraction

1. Introduction
Traditional hot forging manufacturers are constantly improving
productivity in order to be competitive and maintain or increase their
market share and prots. In this respect, there are three main reasons
why companies are investing to optimize their forging processes: the
gradual increase in energy costs (electricity and gas) in recent years
[1], the competitiveness of rms from emerging countries, and the
development of more economic production technologies with good
mechanical properties, such as modern casting [2].
As a result, forging companies are forced to minimize their
production costs and achieve more efcient processes by means of
the following methods [3]:

 Reduction in labour costs via the increased use factory automation systems.

 Increasing the lifetime of forging dies.


 Optimizing the use of available equipment.
 Material optimization by reducing ashes and nishing
operations.

 Shortening the design cycle.


n

Corresponding author. Tel.: 34 985182024; fax: 34 985182022.


E-mail address: marcosperezrd@gmail.com (M. Prez).

http://dx.doi.org/10.1016/j.msea.2014.11.051
0921-5093/& 2014 Elsevier B.V. All rights reserved.

As regards the service lifetime of dies, hot forging involves


extreme working conditions which result in a series of life-limiting
failure mechanisms. Although dies are inevitably replaced after a
certain period of time, the aim is to maximize their lifetime. As die
costs represent between 15 and 30% of the total costs for a closed
die forging process [3,4], producing more parts with the same die
leads to a notable cost reduction. To do so, computer design and
simulation software techniques are useful tools to improve the
design of forging dies. Furthermore, material selection and heat
treatments (HT) are also factors of major importance [5].
Quenched and tempered H13 hot work tool steel is thus mainly
used to make hot forging dies in virtue of its excellent properties:
high hardness and mechanical strength, good toughness and
signicant resistance to shock, thermal fatigue and wear [58].
This paper discusses the performance of four different heat
treatments applied to an H13 steel. Two different quenching media
(gas and oil) and the effects of a cryogenic stage were studied.
In general terms, the application of deep cryogenic treatments
to tool steels is acknowledged to provide the following benets: a
decrease in retained austenite content, a uniformly dispersed
network of ne carbides and a homogenized microstructure.
Reducing the retained austenite content is very convenient, as it
enhances the dimensional stability of the dies, as reported by
Surberg et al. [9], leading to increased service lifetimes. The main

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

advantage of cryogenic treatments, on the other hand, seems to be


that of modifying the carbide distribution pattern, as this modication has a positive effect on hardness, mechanical strength,
fracture toughness and structure homogeneity. The typical carbide
morphologies present in tool steels are: quite large primary
carbides, which do not dissolve during the austenitizing treatment, secondary microcarbides, which precipitate during quenching, and secondary nanocarbides, which are formed during the
tempering stage [10,11]. It is worth noting that cryogenic treatments only modify the nanocarbides precipitated during the
tempering stages, but primary and secondary carbides do not
experience any modication in their morphologies [11]. The
maintenance at  196 1C gives rise to a ne modulated structure
with carbon enriched regions and structural defects. This structure
leads to a ner and more homogeneous carbide precipitation
during the tempering phase and to an increase in the volume
fraction of nanocarbides [10,12,13].
As a consequence of the transformation of the retained austenite and the optimal carbide distribution, Baldissera et al. [14]
reported that cryogenic treatments produce a more homogeneous
structure which delays the nucleation of cracks and increases the
fatigue life of the tool.

2. Experimental methods
2.1. Steel and heat treatments
An H13 steel regularly used to make hot forging dies was
studied in this research. Its chemical composition is given in
Table 1.
Tensile and fracture specimens were subjected to four different
heat treatments involving quenching and tempering. Two quenching media (nitrogen gas and oil) and two types of sequences
(quenching (Q) tempering (T) and quenching (Q) cryogenic
treatment (C) tempering (T)) were carried out. The sequences
and characteristics of the applied treatments can be seen in
Table 2.
Apart from the four sequences described above, an extended
study was conducted to gain an insight into the effects of deep
cryogenic treatments and assess the inuence of the other factors
of the studied heat treatments, such as quenching media and
tempering stages. Two 20  10  8 mm specimens (samples 1
and 2) were cut from the non-deformed central section of the
fracture specimens and were likewise subjected to the heat
treatments described in Table 2. In this case, the effect of quench
severity was studied, comparing oil and air as quenching media.
2.2. Mechanical characterization
The mechanical properties of the H13 steel were measured at
room temperature by means of tensile, hardness and fracture tests.
Tensile tests were performed on cylindrical tensile specimens
of 10 mm diameter with a calibrated length of 70 mm using a MTS
tensile machine equipped with a 250 kN load cell at a displacement rate of 2 mm/min. Three specimens were tested for each
heat treatment. In order to calculate elongation (e), an initial
gage length of 50 mm was considered on each specimen. The
experimental procedure and the determination of the main
Table 1
H13 steel chemical composition (weight%).
%C

%Si

%Mn

%Cr

%Mo

%V

0.39

1.0

0.4

5.2

1.4

0.9

33

parameters were carried out as per the UNE-EN ISO 68921:2009 standard.
The hardness of the steel was determined by means of the
Rockwell C hardness test using a HOYTOM hardness tester under a
load of 150 kg. For treatments HT1 to HT4, samples were extracted
from the non-deformed central section of the SENB specimens
once the fracture toughness test was completed. Five indentations
were made in the thickness direction to appreciate any hardening
due to the cooling rate gradient existing between the surface and
the centre of the samples. Moreover, for samples 1 and 2, Rockwell
C hardness measurements were performed after each stage
(quenching (Q), deep cryogenic treatment (C) and tempering (T))
of the overall heat treatment to appreciate the evolution of
hardness and to compare the results with the Full Width at Half
Maximum (FWHM) parameter, which was obtained in the course
of the X-ray diffraction analysis.
In line with bibliographic references [8,15], KIC fracture toughness tests following the ASTM E1820 standard [16] were carried
out due to the fact that H13 steel is quite brittle at room
temperature. Three single edge notched bend (SENB) specimens,
each with a width (W) of 40 mm, a thickness (B) of 20 mm, a span
(S) of 160 mm (4 W) and a crack size-to-width ratio (a/W) of 0.45,
were tested for each heat treatment. Tests were performed on an
MTS tensile machine, equipped with a 100 kN load cell at a
displacement rate of 0.5 mm/min. Specimens had been fatigue
pre-cracked to the required nominal a/W using a load ratio of 0.1, a
frequency of 3 Hz and a maximum stress intensity factor, K, of
25 MPa  m1/2 in order to prevent sudden fracture of the specimens [15]. The conditional result, KQ, which was determined from
the load versus COD record according to the ASTM E1820 standard,
was consistent with size and linear elastic requirements, so the
results obtained could be considered as material fracture toughness (KIC) [16].
2.3. Fractographic and microstructural characterization
Fractographic and microstructural analyses were carried out to
understand and justify the mechanical testing results. The steel
microstructures were studied by optical and scanning electron
microscopy. The latter technique was also used to characterise the
fracture surfaces.
As far as microstructural examination is concerned, metallographic samples were prepared from the non-deformed central
section of the SENB specimens. These were ground, polished with
diamond paste and nally etched with picral. All samples were
rst observed under a Nikon Epiphot 200 optical microscope and
then, under a FEG-SEM Carl Zeiss GEMINI microscope.
Fracture surfaces from the SENB specimens were likewise
analysed under a JEOL JSM-5600 scanning electron microscope
at different magnications to reveal the predominant failure
micromechanisms. This examination focused mainly on observing
the onset zone of fracture growth from the fatigue pre-crack.
2.4. X-ray diffraction analysis
X-ray diffraction was applied to determine the residual stress
state induced in the heat treatments, the structural distortion, via the
Full Width at Half Maximum (FWHM) parameter, and the retained
austenite content. Measurements were performed for the four
applied heat treatments (HT1HT4), on the non-deformed central
sections of the SENB specimens, as well as on samples 1 and 2.
The FWHM parameter is considered an indicator of the structural distortion of the material. In particular, it is related to the
dislocation density and to the existence of micro-residual stresses
[17]. As thermal and transformation stresses can give rise to local
plastic deformation and to an increase in dislocation density,

34

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

two factors related to steel hardness, the FWHM parameter is


generally acknowledged as a complementary index to characterise
the hardening of the steel [18,19].
In the present study, X-ray diffraction measurements were
performed on a Stresstech XSTRESS 3000 G3R diffractometer. In
order to determine the residual stress state and the FWHM
parameter, the X-ray diffraction technique employed was the
sin2 method, following the recommendations by National Physical Laboratory (NPL) [20], while the determination of FWHM
consisted in the analysis of the diffraction peak broadening. A CrK X-ray source was used, employing a wavelength of 0.2291 nm.
Measurements were taken on the (2 1 1) diffraction peak of the
martensite, which was recorded at a 2 angle of approximately
1561, and the diffraction elastic constant of the selected diffraction
plane, E/(1 ), was 168,900 MPa [21]. Diffraction data were
determined in three different directions on the sample plane:
45, 0 and 451, subsequently calculating the average result. The
longitudinal direction of the samples was designated as direction
01. Nine tilt angles () between -45 and 451, a 20-second
exposure time for each single measurement and a 2 mm diameter
collimator were also used. The diffraction peaks were processed
via a Pseudo-Voigt adjustment and background noise was minimized by means of a parabolic expression.
The retained austenite content was calculated following the
experimental procedure explained in the ASTM E975-03 standard
and in [22]. Samples cut from the non-deformed section of the
SENB specimens were ground and polished following conventional
procedures before measuring the residual austenite content using
the four peak method (two peaks of the ferrite and two peaks of
the austenite). The parameters employed in these measurements
are given in Table 3. It should be noted that the presence of
carbides was not considered in a rst approximation. In a second
approximation, however, the volume fraction of carbides, Vc,
present in every heat treatment was determined by applying
image analysis to the FEG-SEM micrographs, the retained austenite content being corrected accordingly. In contrast, the evolution
of the retained austenite content throughout all the processing
sequences (samples 1 and 2) was calculated without considering

the presence of carbides. In this case, therefore, the obtained


results are only comparative.

3. Results
3.1. Mechanical properties
The tensile test (elastic modulus (E), yield strength (ys),
ultimate tensile strength (UTS), elongation (e) and area reduction
(Z)) and hardness results are shown in Table 4. The differences
between the four heat treatments were not signicant, HT1 being
affording the highest yield strength and ultimate tensile strength.
Cryogenic treatments did not seem to affect the tensile properties
of the steel. No signicant differences in hardness were observed
between the different heat treatments either.
In addition, the evolution of the average hardness throughout
the overall heat treatment is shown in Fig. 1. It can be seen that the
highest hardness value was obtained after the quenching stage
due to the presence of a non-tempered hard martensite. As it will
be seen later, at the end of the cryogenic stage a certain amount of
retained austenite transforms into martensite, while a slight softening of 0.7 and 0.6 HRC was observed for air and oil quenching,
respectively. It seems that the early carbide precipitation during
the heating stage from the cryogenic temperature (  196 1C) to
room temperature leads to a slight decrease in the carbon supersaturation of the martensite and resulting in a slight softening of
the steel [23]. The increase in the dislocation density of the steel
during the cryogenic treatment (plastic deformation due to thermal stresses), especially during the cooling and the holding stages,
does not result in hardening. This is because it is compensated by
structural relaxation and a decrease in structural distortion due to
the onset of the carbide precipitation which takes place during the
heating stage to room temperature.
On the other hand, hardness decreased during the course of
tempering in virtue of carbide precipitation and the subsequent
decrease in the carbon content of the martensite. Nevertheless, the
difference between the quenching (Q) and the rst tempering

Table 2
Studied processing sequences: heat treatments (HT) 14 and samples 1-2.
Heat treatments and
samples

Processing sequences

HT1

Austenitizing
quench
Austenitizing
quench
Austenitizing
quench
Austenitizing
quench

HT2
HT3
HT4
Sample 1
Sample 2

at 1020 1C for 30 min. followed by gas

Triple tempering at 590 1C for 2 h

at 1020 1C for 30 min. followed by gas

Cryogenic treatment at  196 1C for 12 h

Triple tempering at 590 1C for 2 h

at 1020 1C for 30 min. followed by oil

Triple tempering at 590 1C for 2 h

at 1020 1C for 30 min. followed by oil

Cryogenic treatment at  196 1C for 12 h

Triple tempering at 590 1C for 2 h

Austenitizing at 1020 1C for 30 min. followed by air


quench (Q)
Austenitizing at 1020 1C for 30 min. followed by oil
quench (Q)

Cryogenic treatment at  196 1C for 12 h


(C)
Cryogenic treatment at  196 1C for 12 h
(C)

Triple tempering at 590 1C for 2 h


(3 T)
Triple tempering at 590 1C for 2 h
(3 T)

Table 3
X-ray diffraction parameters for retained austenite measurement.
Radiation

Cr K

Filter
Austenite measured planes
Ferrite measured planes

Vanadium
(220)
(211)

(200)
(200)

Carbides (%)

0% or Vc

collimator (mm)
Diffraction angles of austenite (2)(1)
Diffraction angles of ferrite (2)(1)

5
130
156.4

80
106.1

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

35

Table 4
Tensile properties, Rockwell hardness and fracture toughness of the H13 steel after different heat treatments (HT). Mean values (standard deviation).
Heat treatment

HT1

E (GPa)
ys (MPa)
UTS (MPa)
e (%)
Z (%)
Hardness (HRC)
KIC fracture toughness (MPam1/2)a

208
1303
1497
16.8
38.8
43.6
54.8

HT2
( 75)
( 75)
( 72)
( 70.9)
( 70.4)
( 71.1)
( 71.7)

210
1265
1469
14.2
41.8
43.6
67.1

HT3
( 7 5)
( 7 9)
( 7 7)
( 7 1.0)
( 7 1.8)
( 7 0.6)
( 7 5.7)

211
1255
1464
12.1
43.4
43.4
62.3

HT4
( 7 6)
( 7 4)
( 7 3)
( 7 1.8)
( 7 1.8)
( 7 0.8)
( 7 1.5)

211
1253
1469
18.0
41.3
43.1
77.4

( 7 3)
( 7 10)
( 7 10)
( 7 1.3)
( 7 2.2)
( 7 1.1)
( 7 6.6)

All the fracture tests comply with the ASTM E1820-09 standard.

Fig. 1. Evolution of the Rockwell C hardness (HRC) and the Full Width at Half
Maximum (FWHM) parameter throughout the overall heat treatment (samples
1 and 2): quenching (Q), cryogenic (C) and tempering (T) stages.

stage (Q C 1T) was not notable, 1.3 Rockwell C units for the air
quenching and 3.6 units for oil quenching. This slight variation
may be explained by the transformation of the retained austenite,
a soft constituent, into martensite, compensating the loss of
hardness due to the carbon-depleted martensite. Moreover, the
secondary carbide precipitation, which takes place in tool steels
during tempering at 500600 1C, also generates an additional
increase in hardness. The effect of the second and third tempering
was even lower.
Comparing the two heat treatments (air and oil quenching), it
can be seen that the hardness values were not very different.
The evolution of the FWHM parameter throughout the quenching, cryogenic treatment and three tempering stages is also shown
in Fig. 1 for air and oil quenching. This parameter was found to
behave identically for both heat treatments. The highest values
were obtained after the quenching and the cryogenic stages, the
FWHM value subsequently decreasing signicantly during the rst
tempering. In fact, the FWHM parameter shows an 80% decrease
after the rst tempering when quenching in air, and a 90%
decrease when quenching in oil. The second and third tempering
sequences had practically no inuence on this parameter, as can be
seen from Fig. 1.
Fig. 1 also shows that the evolution of the Rockwell hardness
and the FWHM parameter followed the same pattern, although
the decrease in the former during the tempering sequences was
more progressive, while the latter mainly decreased during the
rst tempering. As a result of this observation, it may be stated
that the FWHM is a hardening parameter linked to microstructural
evolution, as the most representative microstructural changes
are produced during the rst tempering sequence, i.e. the tempering of the quenched martensite (carbide precipitation and
decrease in the carbon content of the martensite) and the
transformation of the retained austenite into martensite. In

contrast, the second and third tempering stages give rise to less
signicant microstructural changes, only the tempering of the
fresh martensite which had been transformed during the cooling
of the rst tempering stage.
Signicant differences were observed, however, in the fracture
toughness results, which are shown in Table 4. HT1 provided the
lowest fracture toughness, whereas HT4 offered the highest value.
The positive effect of the cryogenic treatments is especially worthy
to note. Both HT2 and HT4 respectively gave rise to a 22.5 and a
24% increase compared to their respective treatments without
the cryogenic phase, HT1 and HT3. Moreover, a 41% increase in
fracture toughness was obtained via HT4 compared with HT1. The
quenching medium was also seen to affect the toughness of the
steel due to the effect of the cooling rate, oil being a more severe
quench medium than gas: oil quenching (HT3 and HT4) led to
higher toughness values than gas quenching (HT1 and HT2).
The fracture surface of the SENB specimens constituted an
important source of information for justifying the obtained results.
SEM analysis, which focused on the onset zone of crack growth
initiation from the fatigue pre-crack, is reported in Fig. 2 (the
bottom of the gures corresponds to the fatigue pre-crack and the
top, to the crack growth that led to sudden failure of the fracture
specimens). Fracture pattern observation revealed the existence
of a small region where ductile micromechanisms were active
(nucleation, growth and coalescence of microcavities) right at the
onset of the crack growth. The initial crack blunted and a plastic
zone developed in the crack front before complete failure of the
specimen. Although all the specimens behaved in a brittle way
(cleavage fracture), the higher fracture toughness resulting from
the heat treatments provided with a cryogenic stage was directly
linked to the higher plasticity produced in the crack front, this
nding being especially relevant after the HT4 treatment, as can be
seen in Fig. 2d.
3.2. Microstructural analysis
As expected, microstructural analysis of the samples using
optical microscopy revealed the presence of a tempered martensite microstructure, with ne, spherical, homogeneously dispersed
carbides. In order to justify the fracture toughness results, samples
were analysed under an FEG-SEM microscope at 100,000  . Fig. 3
shows the typical carbide pattern of tool steels [10,11]: primary
carbides, secondary micrometric carbides and secondary nanometric carbides.
It seems clear that carbides precipitated in the course of the
HT1 treatment (Fig. 3a) have a relatively coarse and sometimes
elongated appearance, whereas the HT4 microstructure (Fig. 3d)
shows a more uniform distribution of ner carbides. These results
are in agreement with those reported by other authors [1013] in
regard to the effect of deep cryogenic treatments on secondary
carbides: cryogenic treatments produce a ner, homogeneously
distributed carbide network and also an increase in the carbide
volume fraction. Meanwhile, both HT2 and HT3 (Fig. 3b and c) led

36

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

Fig. 2. Fracture toughness surfaces. Crack growth from the pre-crack: (a) HT1, (b) HT2, (c) HT3 and (d) HT4 (500  ).

to intermediate carbide distributions compared to HT1 and HT4


(Fig. 3a and d).
The smaller size of the precipitated carbides and their more
homogeneous distribution pattern seem to be the reason for the
enhanced toughness of the H13 steel as a consequence of the
cryogenic treatments, treatment HT4 being the most suitable heat
treatment to obtain the highest fracture toughness.
The volume fraction of carbides is shown in Table 5 for heat
treatments 14. As expected, quenched and tempered H13 steel
has a notable carbide content, up to 10% in every case. The results
show that cryogenic treatments increased the carbide percentage
of H13 steel 22.3 and a 8.2% when comparing the treatments
without a cryogenic stage (HT1 and HT3) with those that included
this stage (HT2 and HT4), respectively. Besides, the effect of more
severe quenching, such as oil quenching, is similar to the effect of a
cryogenic stage, resulting in an increase in carbide content of 30%
between HT1 and HT4. The obtained data are in accordance with
the KIC fracture toughness; hence, the higher the carbide content,
the higher the toughness, as the carbon content of the martensitic
matrix phase will be lower.
The results of the measured retained austenite content (V)
are given in Table 6. For heat treatments 14, these were
calculated considering the presence of carbides (Table 5) and
without considering their presence. The retained austenite content
suffers a slight drop when the volume fraction of carbides is taken
into account and more precise results are obtained.
In any case, as a result of the triple tempering applied to the
H13 steel, the nal residual austenite content is around or below
3%. Those heat treatments with a cryogenic stage (HT2 and HT4)

show a lower retained austenite content than their respective heat


treatments without it (HT1 and HT4). Nevertheless, there is a low
austenite content that could not be transformed even through the
application of the mentioned cryogenic treatments. It is hence
considered not to have any negative effect during the forging
process, as it will not be susceptible to transformation during the
service life of the die. Along these same lines, according to Villa
et al. [24], the application of cryogenic treatments contribute to
the mechanical stabilization of the retained austenite, avoiding the
formation of martensite during the normal service life of dies.
Fig. 4 gives the results of the retained austenite measurements
after each stage of the overall heat treatment (air and oil quenching). Firstly, the effect of the quenching medium on the retained
austenite content is worth noting. At the end of the quench stage,
oil quenching led to lower austenite content than air quenching,
due to the formers higher cooling rate. In fact, the higher retained
austenite content after air quenching is a consequence of the
stabilization of the austenite owing to the lower cooling rate.
The cryogenic stage modies the residual austenite content,
especially for air quenching; the reduction being less signicant
for oil quenching. These results are in accordance with the revised
literature, since a reduction in the retained austenite content is
usually reported when a cryogenic stage is applied [2429].
Tempering was found to be a good way to transform the retained
austenite into martensite in this steel. In fact, as can be observed in
Fig. 4, the retained austenite fraction decreased to its minimum
value in both the air and oil quenching heat treatments during the
rst tempering cycle. The application of a second and a third
tempering did not provide any further reduction in the austenite

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

37

Fig. 3. H13 steel microstructure at 100,000  magnication (FEG-SEM): (a) HT1, (b) HT2, (c) HT3 and (d) HT4.

Table 5
Volume fraction of carbides (%) measured in the different heat treatments (HT).
Heat treatment

HT1

HT2

HT3

HT4

Carbide content, Vc (%)

11.2

13.7

13.4

14.5

fraction. Hence, H13 steel will have a minimum innate retained


austenite content which cannot be transformed by heat treatment.
This is in agreement with Li et al. [27] who also reported an innate
retained austenite content after the tempering stage.

3.3. Surface residual stresses induced by the heat treatments


Fig. 5 shows the evolution of the average surface residual stress
during the overall heat treatment (air and oil quenching). As
expected, the results revealed the presence of a tensile residual
stress state in the surface of all these specimens. The combination
of thermal contraction and the transformation of austenite into
martensite give rise to a surface tensile residual stress state which
is linked rst and foremost to the severity of the quenching media:
the higher the severity of the media, the higher the residual stress
after the quenching stage.
During the cryogenic stage, a slight relaxation in stress took
place for both heat treatments due to the incipient ne carbide
precipitation which occurs during the heating phase from  196 1C
to room temperature [25]. In fact, according to Das et al. [23], the
carbon atoms present in the supersaturated martensite segregate
to nearby structural defects during the cryogenic stage, forming

clusters. These clusters lead to the precipitation of small carbides


in the following warming up to room temperature.
A noticeable relaxation in stress then took place during the rst
tempering stage. This was the result of carbide precipitation and
the decrease in the carbon content of the martensite [30]. On the
contrary, the second and third tempering stages had apparently no
effect on the stress state of the samples. It is well known that most
of the carbide precipitation and the loss in carbon of the martensite basically take place during the rst tempering. The aim of the
second tempering was to transform the martensite formed during
the cooling of the previous tempering. The third tempering
seemed to have no signicant effect, however, and could be
omitted, given that no new fresh martensite was produced in
the course of the second tempering, as can also be seen in Fig. 4.

4. Discussion
Primary vanadium-rich carbides (MC) are expected to be
formed in H13 steel austenitized at a temperature of 1020 1C [8].
Moreover, molybdenum/iron-rich (M6C) and chromium/iron-rich
(M7C3) carbides might also be present due to the application of
low austenitizing temperatures [8]. On the other hand, secondary
micrometric carbides (M7C3) are supposed to form in grain
boundaries during quenching [8,31]. Anyway, deep cryogenic
treatment and tempering do not affect the morphology of coarse
primary and micrometric secondary carbides [10,11], but only
secondary nanocarbides are modied by these heat treatments.
According to many researchers [30,32] and also bearing in
mind the results obtained in this study, a high internal stress state

38

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

Table 6
Retained austenite content as a function of the heat treatment (HT) and the carbide correction. Volume fraction, % (error, %).
Heat treatment

HT1

HT2

HT3

HT4

Retained austenite content without carbide correctiona


Retained austenite content after carbide correctionb

3.7 ( 71.3)
3.3 ( 71.1)

2.3 ( 7 1.2)
2.0 ( 7 1.0)

2.9 (7 1.1)
2.5 (7 1.0)

2.5 ( 7 1.0)
2.2 ( 7 1.0)

a
b

V V 1.
V V Vc 1.

Fig. 4. Evolution of the retained austenite content (%) throughout the overall heat
treatment (without carbide correction): quenching (Q), cryogenic (C) and tempering (T) stages.

Fig. 5. Evolution of the surface residual stress versus the quenching medium for
each stage of the overall heat treatment (quenching (Q), cryogenic (C) and
tempering (T) stages).

is generated during quenching and cryogenic stages due to


thermal stresses and the transformation of austenite into martensite. In the case under study here only a fraction of the retained
austenite was transformed into martensite during the cryogenic
stage, particularly when quenching was performed in air. Nonetheless, thermal stresses give rise to a great number of structural
defects in the course of this latter stage and the carbonsupersaturated martensite becomes unstable. Carbon atoms move
towards the recently created defects forming clusters in virtue of
two different mechanisms: a short-range diffusion mechanism
[30,32], or as a consequence of the capture and transport of carbon
atoms by moving dislocations [28]. In any case, martensite is
decomposed and carbide precipitation initiates during the warming up phase to room temperature [29,33], producing a reduction
in the residual stress state. However, the main microstructural
changes are produced in the course of the rst tempering stage, as
a signicant precipitation of carbides takes place in the numerous
defects previously created during the quenching and the cryogenic
stages, resulting in a homogeneously dispersed network of very
ne carbides. Furthermore, retained austenite transforms into
martensite during the subsequent cooling phase [28,34], leading
to a nal amount of approximately 2%.
The carbides evolution during tempering is as follows [35]. First
of all, the precipitation of ne -carbide (transition carbide) takes

place in the structural defects created during the quenching and


cryogenic stages [33]. Secondly, there is an in-situ nucleation of
M3C carbides (cementite) on -carbides [35]. The last step is the
formation of alloy carbides responsible for secondary hardening. In
the case of chromium- containing tool steels, M3C carbides transform to M7C3, whereas in tool steels containing molybdenum, M3C
carbides transform to M2C. As H13 steel has both chromium and
molybdenum, M3C carbides are supposed to transform into M7C3
and M2C. This hypothesis is conrmed with the results obtained by
Mesquita et al. [36] for H11 steel, whose chromium and molybdenum contents are similar to H13 ones. Furthermore, according to
other researchers [31,36,37], M7C3 and M2C carbides may transform to their respective more stable carbides, M23C6 and M6C,
when tempering is prolonged or repeated. Thus, after triple
tempering, H13 steel is expected to have M23C6, M6C and MC
nanocarbides [37], the latter having apparently precipitated exsitu due to their smaller size.
On the other hand, oil quenching gives rise to a lower retained
austenite content at the end of the quenching stage compared to
gas or air quenching, due to the higher severity of the former
process. It also produces a similar effect to cryogenic treatment:
the higher cooling rate characteristic of oil quenching generates a
martensite subjected to higher residual stresses.
In contrast, the nucleation and coarsening of carbides during a
heat treatment without a cryogenic stage is quite different to the
process described above. In this case, the internal residual stresses
at the end of the quenching stage are lower and the carbonsupersaturated martensite is now more stable. Consequently,
carbide precipitation starts at a higher temperature during the
course of the rst tempering stage, resulting in a lower fraction of
carbides, which are also able to become larger in size. As a result,
although the nal austenite content is not dependent on the
application of the cryogenic stage, the matrix martensite phase
will have a lower carbon content when a cryogenic stage is
employed. It will thus be able to develop some ductile fracture
micromechanisms before cleavage fracture and will accordingly be
tougher, while the rest of the mechanical properties of the steel
will be virtually the same.
Subsequently, the main effects of the second tempering are to
temper the fresh martensite produced during the cooling of the
rst tempering and to continue the general tempering of the rest
of the microstructure, mainly the carbide coarsening phase. Moreover, the third tempering of this steel seemed to have no
signicant effect, and could be omitted, given that no new fresh
martensite was produced in the course of the second tempering.
All the points discussed here have also been represented in Fig. 6.

5. Conclusions
The tensile mechanical properties and hardness of an H13 steel
were barely changed in the four applied heat treatments, both the
effect of the quench severity and of the cryogenic treatment being
non-signicant.
Rockwell hardness progressively decreased during the three
tempering stages, whereas the FWHM parameter suffered a

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

39

Fig. 6. Phase transformations, carbide precipitation and coarsening at each stage of the heat treatment, highlighting the effect of the cryogenic treatment.

marked decrease only after the rst tempering. As a result of this


nding, it can be stated that the FWHM is a hardening parameter
that is strongly linked to microstructural evolution given that the
most representative microstructural changes were produced during the rst tempering stage. Furthermore, the second and third
tempering stages barely altered the microstructure of the H13
steel as reected by the stabilization of the residual stress state,
the FWHM parameter and the retained austenite content. Furthermore, the third tempering could even be omitted.
Cryogenic treatments notably improved the fracture toughness
of the H13 steel, however. Those treatments with a cryogenic
stage, HT2 (gas) and HT4 (oil), generated a 22.5% and 24% increase
compared to their respective treatments without the cryogenic
stage, HT1 and HT3. The quenching medium also affected the
fracture toughness of the steel due to the effect of the cooling rate.
It has been reported that the ne, homogeneously dispersed
carbide precipitation resulting from the cryogenic stage is responsible for the observed enhancement in toughness. In fact, the
volume fraction of carbides also increased due to the cryogenic
treatments and the use of a more severe quenching. Moreover, the
martensite phase will consequently have a lower carbon content,
giving rise to a tougher matrix phase.
Deep cryogenic treatments reduce the retained austenite content in H13 steel, even though an innate mechanically stable
amount is still present even after triple tempering.
Oil quenching also produced a higher residual stress state and a
lower retained austenite content at the end of the quenching
phase compared to air or gas quenching because of the higher
severity of the former medium.
Bearing in mind the working conditions of hot forging dies, oil
quenching followed by a cryogenic stage (HT4) is considered to be
the best overall heat treatment to increase the fracture toughness
of H13 steel.

Acknowledgements
The authors are grateful for the nancial support for this study
provided by CIE Galfor S.A. and the Spanish Ministry of Economy
and Competitiveness, through project CN-12-025-CDTI.

References
[1] EUROFORGE.org [Internet]. Hagen: EUROFORGE; c2014 [cited 2014 Mar 14].
Energy benchmarking. Available from: http://www.euroforge.org/leadmin/
user_upload/Downloads/EUROFORGE_Energy_BM_2012_13.pdf.
[2] W. Menk, L. Kniewallner, S. Prukner, MTZ worldw 68 (5) (2007) 2324.
[3] M. Shirgaokar, Technology to Improve Competitiveness in Warm and Hot
Forging -Increasing Die Life and Material Utilization- (Dissertation), The Ohio
State University, Columbus (OH), 2008.
[4] M. Bayramoglu, H. Polat, N. Geren, J. Mater. Process. Technol. 205 (2008)
394403. http://dx.doi.org/10.1016/j.jmatprotec.2007.11.256.
[5] S. Babu, D. Ribeiro, R. Shivpuri, Material and Surface Engineering For Precision
Forging Dies, Precision Forging Consortium (US), Ohio Aerospace Institute and
National Center for Manufacturing Sciences, 1999 (Final report).
[6] J.R. Davis, ASM Specialty Handbook-Tool Materials, ASM International, Materials Park (OH) (1995) 119153.
[7] R. Shivpuri, in: S.L. Semiatin (Ed.), ASM Metals Handbook, Volume 14, ASM
International, Materials Park (OH), 2005, pp. 4761.
[8] G. Roberts, G. Krauss, R. Kennedy (Eds.), Tool Steels, fth ed.,ASM International, Materials Park (OH), 1998.
[9] C.H. Surberg, P. Stratton, K. Lingenhle, Cryogenics 48 (2008) 4247. http://dx.
doi.org/10.1016/j.cryogenics.2007.10.002.
[10] P. Farina, Efeito das adies de tratamentos criognicos e de alvio de tenses
no ciclo trmico do ao ferramenta (PhD Thesis), University of So Paulo, 2011.
[11] P. Farina, C. Barbosa, H. Goldenstein, Proceedings of the 18th IFHTSE Congress
(2010) 54175426.
[12] P. Farina, A. Farina, C. Barbosa, H. Goldenstein, Tecnol. Metal. Mater. Miner 9
(2) (2012) 140147.
[13] M. Pellizzari, A. Molinari, S. Gialanella, G. Straffelini, Metall. Italiana 93 (1)
(2001) 2127.
[14] P. Baldissera, C. Delprete, Open Mech. Eng. J. 2 (2008) 111.
[15] J.C. Bergeron, E. Burns, J. Bushie, H. Sandberg, A.V. Heuvel, A. Carrico, Failure
Analysis of H13 Gear Blank Forging Dies (Final report), Michigan Technological
University, Houghton (MI), 2004.
[16] ASTM E 1820-09. Standard Test Method for Measurement of Fracture Toughness, American Society for Testing and Materials.
[17] I.C. Noyan, J.B. Cohen, Residual stress: measurement by diffraction and
interpretation, Springer, New York, 1987.
[18] A.T. Vielma, Parmetros de shot peening y sus efectos en el comportamiento a
fatiga del acero F1272 (PhD Thesis), University of Oviedo, 2013.
[19] I. Fernndez-Pariente, S. Bagherifard, M. Guagliano, R. Ghelichi, Eng. Fract.
Mech. 103 (2013) 29. http://dx.doi.org/10.1016/j.engfracmech.2012.09.014.
[20] M.E. Fitzpatrick, A.T. Fry, P. Holdway, F.A. Kandil, J. Shackleton, L. Suominen,
Measurement Good Practice Guide No. 52.Determination of residual stresses
by X-ray diffraction Issue 2 (Report), National Physical Laboratory, Teddington (UK), 2005.
[21] P.S. Prevey, in: ninth ed.,in: R.E. Whan (Ed.), ASM Metals Handbook, Volume
10, ASM International, Materials Park (OH), 1986, pp. 380392.
[22] Software user's guide, preliminary. Finland: Stresstech Oy, 2010.
[23] D. Das, A.K. Dutta, K.K. Ray, Mater. Sci. Eng. A 527 (2010) 21822193. http://dx.
doi.org/10.1016/j.msea.2009.10.070.

40

M. Prez, F.J. Belzunce / Materials Science & Engineering A 624 (2015) 3240

[24] M. Villa, K. Pantleon, M.A.J. Somers, Acta Mater. 65 (2014) 383392. http://dx.
doi.org/10.1016/j.actamat.2013.11.007.
[25] A. Bensely, S. Venkatesh, D. Mohan Lal, G. Nagarajan, A. Rajadurai, K. Junik, Mater.
Sci. Eng. A 479 (2008) 229235. http://dx.doi.org/10.1016/j.msea.2007.07.035.
[26] M. Preciado, M. Pellizzari, J. Mater. Sci. 49 (2014) 81838191. http://dx.doi.org/
10.1007/s10853-014-8527-2.
[27] S. Li, Y. Xie, X. Wu, Cryogenics 50 (2010) 8992. http://dx.doi.org/10.1016/j.
cryogenics.2009.12.005.
[28] A.I. Tyshchenko, W. Theisen, A. Oppenkowski, S. Siebert, O.N. Razumov, A.
P. Skoblik, et al., Mater. Sci. Eng. A 527 (2010) 70277039. http://dx.doi.org/
10.1016/j.msea.2010.07.056.
[29] D. Yun, L. Xiaoping, X. Hongshen, Heat Treat. Met. 25 (3) (1998) 5559.
[30] D. Senthilkumar, I. Rajendran, M. Pellizzari, J. Siiriainen, J. Mater. Process.
Technol. 211 (2011) 396401. http://dx.doi.org/10.1016/j.jmatprotec.2010.10.018.

[31] G. Krauss, Steels: Processing, Structure, and Performance, ASM International,


Materials Park (OH) (2005) 535560.
[32] K. Amini, A. Akhbarizadeh, S. Javadpour, Vac. 86 (2012) 15341540. http://dx.
doi.org/10.1016/j.vacuum.2012.02.013.
[33] F. Meng, K. Tagashira., R. Azuma, H. Sohma, ISIJ Int. 34 (2) (1994) 205210.
http://dx.doi.org/10.2355/isijinternational.34.205.
[34] M. Koneshlou, K. Meshinchi Asl, F. Khomamizadeh, Cryogenics 51 (2011)
5561. http://dx.doi.org/10.1016/j.cryogenics.2010.11.001.
[35] G. Krauss, Steels: Processing, Structure, and Performance, ASM International,
Materials Park (OH) (2005) 327352.
[36] R.A. Mesquita, C.A. Barbosa, E.V. Morales, H.-J. Kesetenbach, Metall. Mater.
Trans. A 42A (2011) 461472. http://dx.doi.org/10.1007/s11661-010-0430-0.
[37] X. Hu, L. Li, X. Wu, M. Zhang, Int. J. Fatigue. 28 (2006) 175182. http://dx.doi.
org/10.1016/j.ijfatigue.2005.06.042.

Potrebbero piacerti anche