You are on page 1of 9

Metallogr. Microstruct. Anal.

DOI 10.1007/s13632-015-0206-4

TECHNICAL ARTICLE

Effect of Deep Cryogenic Treatment on the Microstructure


and Mechanical Properties of HY-TUF Steel
Ahmad Zare1 • H. Mansouri1 • S. R. Hosseini1

Received: 26 November 2014 / Revised: 8 January 2015 / Accepted: 17 April 2015


 Springer Science+Business Media New York and ASM International 2015

Abstract This paper analyzes how a deep cryogenic Introduction


process changes the microstructure and mechanical prop-
erties of a medium-carbon low-alloy steel. The light Cryogenics can be specified as the application of extreme
microscopy, microhardness testing, and transmission elec- low temperatures and its effects on the properties of
tron microscopy reveal that g-carbide and martensite are engineering materials. In physical metallurgy, cryogenic
constituent phases of deep cryogenic treated (DCT) steel, processes can be utilized as a complementary process to
while the microstructure of conventional heat-treated improve the mechanical properties of bulk metals [1].
(CHT) steel consists of cementite, martensite, and retained Medium-carbon low-alloy HY-TUF steel is a structural
austenite. Transmission electron microscopy also shows steel that possesses high resistance to crack propagation
that the particles of g-carbide have the shape of ultra-fine and high strength after conventional quenching and tem-
globules in DCT martensite. The g-carbides grow in a pering heat treatments. This steel, in the heat-treated con-
Hirotsu and Nagakura orientation relationship to the dition, is often used for aerospace and aircraft components,
martensitic matrix that enables highly coherent interphase such as ultra-sensitive screws and landing gears. In these
boundaries. The results of the mechanical tests, including applications, resistance to crack propagation is an impor-
tensile and Charpy impact tests, show that the deep cryo- tant factor because crack propagation could lead to catas-
genic process can improve toughness in terms of elonga- trophic failure. In recent decades, significant attempts have
tion (*12.81%), tensile fracture energy (*266 MPa), and been made to enhance this property by adjusting the che-
the ductile–brittle transition temperature (*-17.2 C). mical composition, optimizing the steel-making process,
The results of fractography are also consistent with the and tuning heat treatment parameters. For example,
improvement in toughness. It is also found that the strength vacuum arc remelting (VAR) is used to produce HY-TUF
and macrohardness values are increased. Unlike CHT steel, steel to remove non-metallic inclusions, which decrease
discontinuous yielding is observed in DCT steel. Moreover, toughness [2]. However, the effect of deep cryogenic
there is no change in Young’s modulus due to the deep treatments on the toughness of HY-TUF steel has not been
cryogenic treatment. investigated to date.
The phase transition of retained austenite to martensite
Keywords Mechanical properties  Deep cryogenic is usually completed for most steels in the temperature
treatment  Conventional heat treatment  Martensite  range of -60 to -110 C, resulting in an increase in
g-Carbide strength, wear resistance, and dimensional stability, and a
decrease in toughness [3, 4]. Studies show that cooling of
martensite below the martensite finish temperature (usually
& Ahmad Zare temperatures below -120 C), a process referred to as a
ahmad273273@gmail.com deep cryogenic treatment, and subsequent tempering
1 resulted in an improvement in the mechanical behavior of
Department of Materials Engineering, Maleke-ashtar
University of Technology, 83145-115 Isfahan, steels [5–9]. The advantage of utilizing a deep cryogenic
Islamic Republic of Iran process is that it enhances the formation of g-carbides,

123
Metallogr. Microstruct. Anal.

resulting in an enhancement in toughness, wear resistance, Metallographic studies were carried out on specimens
and tensile and fatigue strengths due to fine size of these prepared and etched with two etchants. Vilella’s reagent (1 g
precipitates with a homogenous distribution and spherical picric acid, 5 ml HCl, and 100 ml ethanol) was used to
morphology. Formation of g-carbides is reported to be a reveal retained austenite in martensitic matrix; a mixture of
result of stress-assisted low-temperature diffusion of car- 100 ml H2O, 10 g K3Fe(CN)6, and 10 g NaOH was used to
bon atoms. The stresses are mainly supplied from thermal reveal carbides. Volume fractions of carbides and retained
stresses during cooling and phase transformation stresses of austenite were measured with image analysis software. The
austenite to martensite [10, 11]. microhardness of constituent phases was determined using a
According to many studies [10, 12–15], deep cryogenic load of 300 g force. In order to confirm the presence of
processing at liquid nitrogen temperature enhances the retained austenite in the examined microstructures, the steels
mechanical properties of many tool and die steels. How- were analyzed by means of x-ray diffraction (XRD). Addi-
ever, there are only a few researchers and commercial heat tionally, cross-sectional slices (150 lm) were prepared from
treaters who perform heat treatment process of medium- the end of deep cryo-treated bars for transmission electron
carbon low-alloy steels with deep cryogenic process. Pre- microscopy studies, by the conventional method of wire
vious works on HY-TUF steel showed superior mechanical cutting and mechanical grinding to a thickness of about
properties of deep cryogenic treated (DCT) steel for 80 lm. 2.5-mm-disks were prepared from these specimens
holding time of 48 h at liquid nitrogen in comparison to and then jet polished with 15% perchloric acid and 85%
other holding times (12, 24 and 72 h) [16]. To our acetic acid solution at 25 C until the disks were perforated.
knowledge, no research has been done on the effect of deep Microstructural characterization of cross-sectional samples
cryogenic process on the microstructural changes of was evaluated using a high-resolution transmission electron
medium-carbon low-alloy steels and its simultaneous cor- microscope (HRTEM), operating at 300 kV and camera
relation to mechanical micromechanisms. This study aims length of 20 cm. For indexing carbides, three different pat-
to investigate the effect of deep cryogenic treatment on terns near major low index zone axis were taken.
microstructural aspects such as volume fraction, mor- Charpy V-notched and tensile samples were wire cut from
phology, and type of constituent phases (martensite, car- the as-received bars and semi-finished by machining, prior to
bide, and retained austenite). Moreover, mechanical cryogenic heat treatment. This action was done due to the
properties such as Young’s modulus, yielding behavior, high strength ([1200 MPa) of heat-treated steels. After pre-
micromechanisms of strengthening, increasing toughness, scribed cryogenic heat treatments, all samples were ground to
and mode of fracture are intended to be correlated to final dimensions. For each microstructure, three tensile sam-
microstructural aspects. ples and four Charpy specimens were also prepared. The
Charpy impact specimens were prepared according to the
ASTM-A370 standard. Ductile-to-brittle transition curves
Experimental Method were obtained using impact tests on Charpy V-notched spe-
cimens at different temperatures between -75 and 95 C.
HY-TUF steel bars with chemical composition given in The specimen temperature was obtained by immersing it in a
Table 1 were used in this study. The steel was received in mixture of methanol, water, NaCl, and liquid nitrogen for
the form of 100-mm-diameter forged bars. They were tests below 25 C and in boiling water for tests above 25 C.
austenitized under argon protective atmosphere, DCT at The ductile–brittle transition temperature (DBTT) was de-
liquid nitrogen temperature (-196 C) and tempered fined as the temperature for which the fracture energy is
according to heat treatment schedules listed in Table 2. As 20.4 J (15 ft-lb) [17]. The tensile samples were prepared
shown in this table and recommended by [2], oil quenching according to ASTM E8M-04 [18] and tested with a cross-
was used for hardening machined components made from head speed of 2 mm/min (nominal strain rate of 9.3 9 10-4).
HY-TUF steel. Oil is an ideal quenchant for achieving the The analysis of the failure mechanisms was carried out by
required hardness and strength while keeping residual direct topographical observation of fractured tensile samples,
stresses and distortion, in comparison to water quenching. using a scanning electron microscope (SEM).
No cracking was experimentally observed after oil
quenching.
Results and Discussion
Table 1 Chemical composition of HY-TUF steel (wt%) Microstructure
Ni Mn Si Mo Cr C Fe
Figure 1 shows the representative microstructures con-
1.80 1.40 1.60 0.45 0.30 0.25 Balance
taining carbides (dark regions). A change in the

123
Metallogr. Microstruct. Anal.

Table 2 The used heat


Specimen Heat cycle and cryogenic process
treatment and supplementary
cryogenic process cycles Austenitizing Deep cryogenic process Tempering

CHT 850 C, 1 h ? oil-quenched … 300 C, 1 h ? air cooled


DCT 850 C, 1 h ? oil-quenched -196 C, 48 h ? Air warmed 300 C, 1 h ? air cooled

diffraction pattern analysis of TEM samples). This indi-


cates that the mechanism of carbide formation has been
changed by the deep cryogenic process. Two different
mechanisms have been suggested for the formation of
g-carbides. Baldissera [19] attributed the abnormally low
tetragonality of martensite to the migration of carbon atoms
at deep cryogenic temperatures. Tyshchenko et al. [10]
proposed the absorption of carbon atoms by gliding dis-
locations during the deep cryogenic treatment and corre-
sponding decrement in the carbon content in martensitic
matrix, which should result in the formation of g-carbides.
It has been reported that during deep cryogenic treatment,
volume dilatation due to the transformation of austenite to
martensite produces noticeable amount of mobile disloca-
tions which are able to glide at the temperature of deep
cryogenic treatment [10].
The effect of g-carbide formation on the carbon content
of martensitic matrix can also be observed in the micro-
hardness values of this phase. As shown in Table 3, the
microhardness of the martensite in CHT microstructure is
higher than that in DCT microstructure. The hardness of
martensite is directly proportional to its carbon content.
Therefore, the martensitic matrix in DCT steel is assumed
to have lower carbon content, confirming the formation of
g-carbides which have higher carbon content (Fe2C)
compared to that of cementite (Fe3C).
Figure 2 shows representative microstructures of the
materials illustrated in Fig. 1, after the removal of all traces
of the earlier etchant and etching with Vilella’s reagent.
Fig. 1 Optical microstructures of (a) CHT and (b) DCT steels Here the microstructure of DCT steel (Fig. 2b) appears to
showing carbides be fully martensitic, with small amounts of retained
austenite (0.69%), confirming the results of XRD in Fig. 3,
which show no austenite peaks. Table 3 shows the micro-
morphology of carbides is observed after performing deep hardness of the retained austenite phase. It is observed that
cryogenic process. In this case, the martensitic matrix is the microhardness is essentially the same for the DCT and
decorated by a homogeneous distribution of fine spherical CHT samples, considering the standard deviation for this
carbides. The volume fractions of carbides are summarized data. In contrast, the CHT specimen (Fig. 2a), for which
in Table 3, which shows that the volume fraction of carbide the mean microhardness magnitude is 361, shows a non-
is higher in DCT steel when compared to that of the con- uniform distribution of retained austenite, with a higher
ventional heat-treated (CHT) steel. Table 3 also compares volume fraction (16.58%). This is due to the fact that the
the microhardness of carbides in the studied steels and austenite phase is forced to undergo a martensitic trans-
clearly shows that the deep cryogenic process produces a formation when the deep cryogenic process is applied. It is
significant increase in microhardness of the carbide phase. worth mentioning, as Yu and Yang [20] have reported, that
The carbide in DCT steel is g type, and its microhardness austenite can be retained even below the martensite start
is higher when compared to the carbide in CHT steel (as temperature (Ms), because this phase can be constrained by
discussed later, the presence of g-carbide is confirmed with martensite laths. Hence, the thermal stresses of CHT

123
Metallogr. Microstruct. Anal.

Table 3 The resultant microstructure constituents with their microhardness values


Specimen Microstructure constituents
Carbide Retained austenite Martensite
vol.% Microhardness (HV) vol.% Microhardness (HV) vol.% (balance) Microhardness (HV)

CHT 4.41 ± 0.32 900 ± 14.4 16.58 ± 0.50 361 ± 9.7 79.01 ± 0.45 918 ± 17.2
DCT 8.87 ± 0.72 2023 ± 15.6 0.69 ± 0.33 365 ± 11.2 90.44 ± 0.71 753 ± 9.1

Fig. 3 XRD results of (a) CHT and (b) DCT steels

shows that the lattice layers of g-carbide have been


nucleated and grown coherently with the martensitic matrix.
Moreover, the interphase boundaries between g-carbide and
martensite phases are found to be continuous. Fig-
Fig. 2 Optical microstructures of (a) CHT and (b) DCT steels
ure 4(c) shows a selected area diffraction pattern and its
showing retained austenite indexing. To obtain this diffraction pattern, the incident
electron beam was selected to be parallel to the [131] and
[121] directions of martensite and g-carbide, respectively.
quenching cannot supply sufficient activation energy (shear
The resultant diffraction pattern (Fig. 4c) is conformed
stresses) for completing martensitic transformation in some
with that which was obtained by Hirotsu and Nagakura (H–
instances.
N). Their TEM studies on martensite and g-carbide phases
The microstructure of the DCT steel as observed in the
show that the [001] direction of g-carbide is parallel to the
TEM is shown in Fig. 4(a). In most regions of the
[100] direction of martensite; also the (110) plane of g-
martensitic matrix, ultra-fine spherical particles of g-car-
carbide is parallel to the (010) plane of the martensite
bide developed within boundaries of mechanical twins.
phase. The H–N orientation relationship between marten-
According to Fig. 4(a), g-carbide particles vary in size
site (M) and g-carbide (g) phases can be summarized as
from *200 to 300 nm. Figure 4(b) is a higher magnifica-
below:
tion view of the rectangular region of the
g-carbide/martensite boundary in Fig. 4(a). This figure ð110Þg jjf010gM ; ½001g jjh100iM ; ð1Þ

123
Metallogr. Microstruct. Anal.

Fig. 5 Representative engineering stress–strain curves of (a) CHT


and (b) DCT steels

This orientation relationship could also confirm that


g-carbides are coherent with martensitic matrix in DCT
steel.

Mechanical Properties

Figure 5 illustrates typical engineering tension stress–strain


curves of the investigated steels at room temperature. As
shown in this figure, the stress–strain curves of both steels
exhibit the same Young’s modulus (E = 206 GPa), which
has been interpreted to correspond to the independency of
Young’s modulus to microstructure. It should be men-
tioned that Young’s modulus only depends on the com-
position and temperature [21]. Figure 5 also compares
yielding behaviors of CHT and DCT steels. The CHT steel
exhibits a continuous yielding behavior, as opposed to the
discontinuous yielding of the DCT steel, which can be
explained as follows. Sakaki et al. [22] calculated the
distribution of the internal stresses in the tough phase
(austenite) adjacent to martensitic matrix due to the volume
dilatation of martensitic transition based on the continuum
mechanics. They found that there are many preferential
yielding regions within the tough phase (austenite), and
initial yielding begins simultaneously from these regions
under a very low applied force, compared to the typical
yield strength of the austenite, because the internal stresses
decrease the initial yielding strength. Thus the CHT steel,
with 16.58% austenite, yields continuously. The DCT steel
exhibits discontinuous yielding because this mechanism is
strongly dependent on the simultaneous presence of a
deformable soft phase (austenite) and a deformable hard
Fig. 4 (a) TEM micrograph of g-carbide precipitation in DCT steel,
phase (martensite) in microstructure; with only 0.69%
(b) HRTEM image of g-carbide/martensite boundary, and (c) indexed austenite, this mechanism is not possible in the DCT steel.
diffraction pattern illustrating martensitic matrix and g-carbide The mechanism of discontinuous yielding can be related to

123
Metallogr. Microstruct. Anal.

Table 4 Mechanical properties of studied steels


Mechanical property Specimen
CHT DCT

Macrohardness (HV) 498 ± 35.8 536 ± 43.6


Yield strength (MPa) 1441 ± 12.3 1584 ± 17.5
Ultimate tensile strength (MPa) 1531 ± 19.2 1697 ± 25.7
Percent elongation 9.94 ± 1.1 12.81 ± 0.6
Tensile fracture energy (MPa) 95 ± 8.9 266 ± 11.4
Strength difference (MPa) 90 ± 11.2 113 ± 4.8

the decrease in carbon content of DCT martensite, which


was investigated by Tyshchenko et al. [10] with TEM
observation, Mössbauer spectroscopy, and internal friction
studies. They reported retained austenite is capable of time-
dependent transformation into martensite with holding at
the temperatures of deep cryogenic processing. This phe-
nomenon, which is known as isothermal martensitic Fig. 6 Absorbed Charpy impact energy vs. testing temperature
curves of (a) CHT and (b) DCT steels
transformation, can produce mobile dislocations in DCT
martensite due to the volume expansion of c ? M. Carbon
atoms are captured by the mobile dislocations. This will which have a low activation energy for nucleation of
result in dislocation pinning, which leads to the discon- microcracks [26]. This is due to the crystallographic mis-
tinuous yielding behavior of DCT steel. match between the structure of large-sized cementite and
The results of macrohardness and tensile tests are given high-carbon martensite [27]. On the contrary, in DCT steel
in Table 4. In this table, the tensile fracture energies as a as given in Eq. (1) and also reported by Jack and Jack [28],
criterion of fracture toughness were digitally determined by there is a crystallographic correlation between crystals of
calculating the areas under curves of Fig. 5. Variations of g-carbides and low-carbon martensite, which causes the
Charpy impact energy with temperature for the studied formation of coherent boundaries. It should be mentioned
CHT and DCT steels are shown in Fig. 6. The determined that nucleation of microcracks is retarded at coherent
ductile-to-brittle transition temperatures (DBTT) are boundaries (as shown in Fig. 4b) and within regions adja-
approximately -12.2 C for the CHT steel and -29.4 C cent to boundaries.
for the DCT steel. DBTT is referenced as a temperature for From Table 4, it can be seen that the DCT steel has
transition of fracture mode from ductile to brittle. As the higher macrohardness, yield strength, and ultimate tensile
CHT steel has a higher DBTT, it is more prone to fracture strength in comparison to the CHT steel, which can be due
at a given temperature in comparison with the DCT steel to the presence of g-carbides in the vicinity of martensite.
which has a lower DBTT. It can be seen from Table 4 and The parameters which cause lower strength and macro-
Fig. 6 that DCT steel has greater toughness in terms of hardness of CHT steel are (1) lower volume fraction of
elongation (*12.81%), tensile fracture energy martensite; (2) lower microhardness (or strength) of
(*266 MPa), and DBTT (*-29.4 C). The increase of cementite in comparison to g-carbides. g-Carbides are
toughness observed here is in agreement with the previous stronger than cementite, due to the orthorhombic structure
works, which were conducted on medium-carbon low-alloy of g-carbides, which have stronger atomic bonds and are
steels [23, 24]. It is reported that DCT martensite is tougher coherent with martensitic matrix [28]; (3) lower volume
than CHT martensite in medium-carbon low-alloy steels as fraction of carbides (Table 3); and (4) higher volume
explained in ‘‘Microstructure’’ section. The observed im- fraction of retained austenite.
provement in toughness of the DCT steel in comparison to The difference between the ultimate tensile strength and
the CHT steel can be related to the fact that DCT yield strength can be considered as a measure of the work
martensite has lower concentration of carbon, and conse- hardening occurring in the sample [29]. Table 4 also gives
quently higher ductility in comparison to CHT martensite. the difference between ultimate tensile and yield strengths
The DCT martensite retards propagation of microcracks, of the studied steels. From this it can be concluded that the
which consequently increases resistance to brittle fracture clustering of carbon atoms in the vicinity of dislocations
[25]. In addition, in the CHT steel, cementite/martensite (formation of g-carbides) during the deep cryogenic pro-
boundaries and regions within them are high-energy sites cess causes a noticeable decrease in the carbon

123
Metallogr. Microstruct. Anal.

concentration of the martensitic matrix and consequently


an increase in the work hardening of DCT steel.

Fractography of Tensile-Tested Specimens

Microscopic observations were carried out on the fracture


surfaces of tensile-tested specimens to understand the effect of
the deep cryogenic process on the fracture mode of HY-TUF
steel. For the DCT specimens, observation of the fracture
surface indicates that necking has occurred during tensile
testing. The fracture surface exhibits typical ‘cup-cone’ mode
of fracture, which includes the fiber pattern in the central zone,
the shear lip pattern in outside zone, and the radial pattern. An
example of these patterns is shown in Fig. 7, which corre-
sponds to that for the samples deep cryo-treated at -196 C.
The fracture surface of CHT sample is shown in
Fig. 7(a). There are two parts which can be distinctly
observed from the fracture feature: (1) central zone where
the material fractured through cleavage, and (2) an outer
zone where fracture occurred through growth of dimples
and their subsequent coalescence. Comparison of the ap-
pearance of the fractographs in Fig. 7 shows that the area
of brittle fracture (central zone) for the CHT steel is larger
than that in the DCT steel, while the area of shear lip zone
(outside zone) for the DCT steel is larger than that in the
CHT steel; there is no sign of radial marks for the CHT
steel. This evidence indicates a lower ductility of the CHT
steel compared to that of the DCT steel [30], which
matches the data in Table 4.
SEM micrographs of the central area of fracture surfaces
are shown in Fig. 8. In this figure, mean values of dimple
sizes are also shown on the lower right corners. The frac-
ture surface of the DCT steel (Fig. 8b) exhibits a distri-
bution of dimples with varying sizes, showing a general
ductile fracture mode. There is a direct correlation between
microvoid (dimple) coalescence and fracture strain. Since
larger and deeper dimples are joined together at greater
strains, they are representative of larger fracture strains. On
the contrary, smaller and shallower dimples indicate less
fracture toughness in steels [31]. The mean value of dimple
size at fracture surface of the DCT steel (Fig. 8b) is greater Fig. 7 Stereographs captured from tensile-tested fracture surfaces of
(a) CHT and (b) DCT steels
than that of the CHT steel (Fig. 8a), indicating an
enhancement in ductility (*2.87%). In the CHT steel
(Fig. 8a), a mixture of bright facets and dimples with dominant coherent boundaries of g-carbide/martensite (as
smaller size are observed; these are distinguished as fea- shown in Fig. 4b) limit the nucleation of microcracks as a
tures of semi-cleavage fracture. The reason for the semi- toughening mechanism. On the other hand, the g-carbide
cleavage mode of fracture is related to the formation of prevents crack propagation in the DCT martensitic matrix
large-sized cementite in the high-carbon martensitic by bridging the crack forming ligaments behind the
matrix. Since crack nucleation can easily occur in the high- microcrack tip, or by creating a stress field in front of the
energy boundaries of large cementite/martensite, and crack microcrack tip [33]. This causes higher elongation in
propagation is relatively easy at the high-carbon marten- necking stage for the DCT steel. Hence, with the formation
sitic islands, these result in a decrease in ductility of the of g-carbides, the semi-cleavage mode of fracture is
CHT steel compared to that of DCT steel [32]. The replaced with the dimpled mode.

123
Metallogr. Microstruct. Anal.

4. The resultant engineering stress–strain curves show


that the Young’s modulus (*206 GPa) is not affected
by the performance of deep cryogenic treatment.
5. Continuous yielding is observed in the CHT steel. This
behavior can be attributed to dislocations which are
unlocked and generated during low-temperature
martensitic transformations within retained austenite
phase. In contrast, the DCT steel has discontinuous
yielding behavior resulting from the lack of any
appreciable austenite phase during deep cryogenic
treatment.
6. DCT steel has greater macrohardness, yield strength,
and ultimate tensile strength than the CHT steel. This
is attributed to the lower martensite volume fraction in
the CHT steel, the lower strength of cementite in
comparison to g-carbides, a higher volume fraction of
carbides in DCT steel, and a lower volume fraction of
retained austenite in DCT steel.
7. Toughness, in terms of elongation, tensile fracture
energy, and Charpy impact energy, is greater in the
DCT steel than in the CHT steel. This is related to the
lower carbon content of the martensitic matrix due to
the precipitation of g-carbides which have higher
carbon content in their stoichiometry formula, com-
pared to cementite. Crack nucleation and propagation
Fig. 8 SEM micrographs captured from central part of tensile-tested
fracture surfaces of (a) CHT and (b) DCT steels (numbers on the can also occur easily in the high-energy boundaries of
lower right corners are average dimple size) large-sized cementite/martensite. This results in a
decrease in the toughness of the CHT steel compared
to that of DCT steel.
Conclusion
8. The tensile fracture investigations confirm an increase
in toughness of the HY-TUF steel after deep cryogenic
The CHT and DCT steels with different microstructural
treatment as a supplemental process to conventional
constituents were produced through conventional heat
heat treatment. The fractography also shows that
treatment and supplementary deep cryogenic process at
dimples have a tendency to appear shallower and
liquid nitrogen temperature. The mutual relationship
thinner, and the character of the fracture surface
between the microstructure and mechanical behavior can
switches from flat to cup-cone while performing deep
be summarized as below:
cryogenic process.
1. Formation of fine homogenized g-carbides, disappear-
ance of retained austenite, and decrease in carbon
content of martensite take place during the deep
cryogenic treatment at -196 C.
References
2. TEM studies show the existence of ultra-fine g-carbide
globules which are coherent with the DCT martensite. 1. R.B. Scott, Cryogenic Engineering, vol. 282 (Van Nostrand,
The diffraction pattern of the DCT steel validates the Princeton, 1959)
H–N orientation relationship between martensite and 2. W. Klopp, Aerospace Structural Metals Handbook, 39th edn.
g-carbide phases. (Purdue Research Foundation, West Lafayette, 1992), pp. 1–12
3. A. Bensely, A. Prabhakaran, D. Mohan Lal, G. Nagarajan, En-
3. Microhardness values of carbides show cementite and hancing the wear resistance of case carburized steel (En 353) by
g-carbides form in CHT and DCT steels, respectively. cryogenic treatment. Cryogenics 45, 747–754 (2005)
Moreover, lower microhardness value of martensitic 4. D. Candane, N. Alagumurthi, K. Palaniradja, Effect of cryogenic
matrix in DCT steel confirms its lower carbon content. treatment on microstructure and wear characteristics of AISI M35
HSS. Int. J. Mater. Sci. Appl. 2(2), 56–65 (2013)
This is attributed to the formation of g-carbide during 5. V. Firouzdor, E. Nejati, F. Khomamizadeh, Effect of deep
the deep cryogenic process, which absorbs the carbon cryogenic treatment on wear resistance and tool life of M2 HSS
atoms of martensite. drill. J. Mater. Process. Technol. 206, 467–472 (2008)

123
Metallogr. Microstruct. Anal.

6. S. Li, L. Deng, X. Wu, Y.A. Min, H. Wang, Influence of deep 19. P. Baldissera, Deep cryogenic treatment of AISI 302 stainless
cryogenic treatment on microstructure and evaluation by internal steel: Part I—Hardness and tensile properties. Mater. Des. 31,
friction of a tool steel. Cryogenics 50, 754–758 (2010) 4725–4730 (2010)
7. F. Meng, K. Tagashira, R. Azuma, H. Sohma, Role of eta-carbide 20. T. Yu, J. Yang, Effect of retained austenite on GPM A30 high-
precipitations in the wear resistance improvements of Fe-12Cr- speed steel. J. Mater. Eng. Perform. 16, 500–507 (2007)
Mo-V-1.4C tool steel by cryogenic treatment,’’. ISIJ Int. 34, 21. M.A. Meyers, K.K. Chawla, Mechanical Behavior of Materials,
205–210 (1994) vol. 547 (Cambridge University Press, Cambridge, 2009)
8. D. Mohan Lal, S. Renganarayanan, A. Kalanidhi, Cryogenic 22. T. Sakaki, K. Ohnuma, K. Sugimoto, Y. Ohtakara, Plastic ani-
treatment to augment wear resistance of tool and die steels. sotropy of dual-phase steels. Int. J. Plast 6, 591–613 (1990)
Cryogenics 41, 149–155 (2001) 23. S. Zhirafar, A. Rezaeian, M. Pugh, Effect of cryogenic treatment
9. A. Molinari, M. Pellizzari, S. Gialanella, G. Straffelini, K. Sti- on the mechanical properties of 4340 steel. J. Mater. Process.
asny, Effect of deep cryogenic treatment on the mechanical Technol. 186, 298–303 (2007)
properties of tool steels. J. Mater. Process. Technol. 118, 350–355 24. D. Senthilkumar, I. Rajendran, M. Pellizzari, J. Siiriainen, In-
(2001) fluence of shallow and deep cryogenic treatment on the residual
10. A. Tyshchenko, W. Theisen, A. Oppenkowski, S. Siebert, O. state of stress of 4140 steel. J. Mater. Process. Technol. 211,
Razumov, A. Skoblik et al., Low-temperature martensitic trans- 396–401 (2011)
formation and deep cryogenic treatment of a tool steel. Mater. 25. G.E. Totten, Steel Heat Treatment: Metallurgy and Technologies,
Sci. Eng., A 527, 7027–7039 (2010) vol. 1 (CRC Press, Boca Raton, 2007)
11. V. Alexiades, E.C. Aifantis, Singular problems in the theory of 26. M. Sauzay, K. Vor, Influence of plastic slip localization on grain
stress-assisted diffusion. SIAM J. Math. Anal. 14, 925–933 boundary stress fields and microcrack nucleation. Eng. Fract.
(1983) Mech. 110, 330–349 (2013)
12. A. Akhbarizadeh, M. Golozar, A. Shafeie, M. Kholghy, Effects of 27. R. Padmanabhan, W. Wood, Precipitation of e carbide in
austenizing time on wear behavior of D6 tool steel after deep martensite. Mater. Sci. Eng. 65, 289–297 (1984)
cryogenic treatment. J. Iron. Steel Res. Int. 16, 29–32 (2009) 28. D. Jack, K. Jack, Invited review: carbides and nitrides in steel.
13. K. Amini, A. Akhbarizadeh, S. Javadpour, Investigating the ef- Mater. Sci. Eng. 11, 1–27 (1973)
fect of the quench environment on the final microstructure and 29. A. Zare, A. Ekrami, Effect of martensite volume fraction on work
wear behavior of 1.2080 tool steel after deep cryogenic heat hardening behavior of triple phase (TP) steels. Mater. Sci. Eng.,
treatment. Mater. Des. 45, 316–322 (2012) A 528, 4422–4426 (2011)
14. M. Koneshlou, K. Meshinchi Asl, F. Khomamizadeh, Effect of 30. ASM Handbook, Fractography, vol. 12 (2nd printing, 1992)
cryogenic treatment on microstructure, mechanical and wear (ASM International, Materials Park, 1987), p. 217
behaviors of AISI H13 hot work tool steel. Cryogenics 51, 55–61 31. A. Zare, A. Ekrami, Influence of martensite volume fraction on
(2011) tensile properties of triple phase ferrite–bainite–martensite steels.
15. A. Oppenkowski, S. Weber, W. Theisen, Evaluation of factors Mater. Sci. Eng., A 530, 440–445 (2011)
influencing deep cryogenic treatment that affect the properties of 32. R.W.K. Honeycombe, H.K.D.H. Bhadeshia, Steels: Microstruc-
tool steels. J. Mater. Process. Technol. 210, 1949–1955 (2010) ture and Properties (Edward Arnold, London, 1981)
16. A. Hadi, MS thesis, Maleke-ashtar University of Technology, 33. P. Novak et al., A statistical, physical-based, micro-mechanical
Isfahan, Iran (2014) (in Persian) model of hydrogen-induced intergranular fracture in steel. Mech.
17. ASTM International, Iron and Steel Products, Annual Book of Phys. Solids 58, 206–226 (2010)
ASTM Standards, A370-B, 01.01 (ASTM International, West
Conshohocken, 1989)
18. ASTM Standard E8M-04, Standard Test Methods for Tension
Testing of Metallic Materials (ASTM International, West Con-
shohocken, 2004)

123

You might also like