Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3

Available online at www.sciencedirect.com

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

Temperature-dependent deformation and fracture


properties of low-carbon martensitic steel in
different stress states

€nemann a,
Zhen Zhang a, Fuhui Shen a,*, Hanqing Liu b, Markus Ko
a
Sebastian Münstermann
a
Integrity of Materials and Structures, Steel Institute, RWTH Aachen University, Intzestraße 1, 52072, Aachen,
Germany
b
Department of Mechanical Engineering, Kyushu University, Fukuoka, 819-0395, Japan

article info abstract

Article history: The temperature-dependent ductile-brittle transition (DBT) in low-carbon martensitic steel
Received 26 April 2023 is investigated in different stress states. To capture the effects of the stress state, tensile
Accepted 7 June 2023 tests have been performed with flat specimens and three-point bending tests have been
Available online 10 June 2023 conducted with Charpy specimens at dynamic (impact) and quasi-static loading rates,
covering a broad temperature range of 77 Ke423 K. The results reveal that ductile-brittle
Keywords: transition temperature (DBTT) is significantly affected by stress triaxiality, wherein the
Martensitic steel transition temperature in uniaxial tension is ~142 K lower than that in plane strain con-
Fracture dition of Charpy specimens. The energy absorption capacity in the ductile resistance
DBTT domain during three-point bending tests is sensitive to the loading rate. Moreover, dy-
Loading rate namic strain aging poses a decrease in ductile resistance upon quasi-static loading when
Stress state the temperature exceeds 323 K and thus without a specific upper shelf plateau. Tensile
strength and ductility tend to increase with decreasing temperature, and the associated
failure mode changes from shear-dominated to nearly pure tension due to the specific
discrepancy of temperature influence on the critical normal and shear strengths, wherein
the formation of shear bands that dominate the tensile deformation could be strictly
inhibited by the formation of nanoscale dislocation cells based on dislocation analysis. The
failure mechanisms in tensile tests correlate with the tensile specimens' macroscopic
fracture angle. In the end, a new local parameter is proposed for temperature-dependent
toughness estimation in tensile tests.
© 2023 The Authors. Published by Elsevier B.V. This is an open access article under the CC
BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

martensitic steels used in developing energy-efficient and


1. Introduction wear-resistant structural components in the tool industry [1].
However, one of the long-standing issues in bcc metals is the
Strength and toughness are key mechanical properties of transition of deformation and fracture behaviors with
body-centered cubic (bcc) metals, especially for low-carbon

* Corresponding author. RWTH Aachen University, Germany


E-mail address: fuhui.shen@iehk.rwth-aachen.de (F. Shen).
https://doi.org/10.1016/j.jmrt.2023.06.070
2238-7854/© 2023 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).
1932 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3

temperature [2,3]. Below a critical temperature, referred to as In this work, tensile and three-point bending tests covering
ductile-brittle transition temperature (DBTT), the energy a wide temperature range are carried out to understand DBTT
required to activate dislocation emission at the crack tip at low and high stress triaxiality conditions. The loading rate
manifests a sudden change from below the energy required effect on the transition temperature of low-carbon martens-
for cleavage to above that energy, resulting in a transition itic steel is investigated by comparing the experimental re-
from ductile deformation to brittle fracture [4e6]. This limits sults implemented via three-point bending tests at different
the temperature window for the utilization of bcc metals to loading rates. A fundamental understanding of the mecha-
some extent, and suggests that the temperature-dependent nism dominating the DBT is involved in terms of fracture
ductile-brittle transition (DBT) can be traced back to angle, dimple fraction and dislocation activity, and the stress
the dislocation activities. Since the substructures of state effect on the transition temperature is further discussed.
low-carbon martensitic steels are mainly dislocations, the
dislocation activities are more pronounced compared with
other bcc metals, but tend to be suppressed at lower temper- 2. Experimental procedures
atures [7,8].
Conventionally, the DBTT is evaluated by means of 2.1. Material and specimen preparation
impact tests with standard V-notch Charpy specimens
(55  10  10 mm3 in size) [9e12]. Moreover, the impact The as-received material utilized in this work is an 8 mm thick
toughness degradation can be reflected not only in an upward martensitic steel plate. The plate was austenitized at 1223 K
shift of the transition temperature between brittle and ductile for 15 min, followed by water quenched to room temperature
domains but also in the decrease of crack resistance within and subsequently tempered at 473 K for 60 min to eliminate
the ductile domain [13]. External conditions, likewise loading residual stress that developed during the quenching process.
rate, is a crucial factor affecting DBT. As the loading rate in- The chemical composition (wt%) of the alloy determined by
creases, a high transition temperature can be expected, as wet chemical analysis is tabulated in Table 1.
well as a high upper shelf plateau in the ductile domain At the position corresponding to the 1/3 of the thickness,
[14,15]. The effects of stress state on ductile fracture have been 2 mm thick sheets were cut from the plate, from which flat
intensively investigated for many engineering materials dog-bone-shaped tensile specimens with a gauge section of
[16e19]. At the temperature of liquid nitrogen, stress-state- 50 mm in length, 12.5 mm in width, and 2 mm in thickness
dependent DBT is also observed by Shen et al. [20,21], and were manufactured along the rolling direction (RD). Due to the
they concluded that the failure mode switches from cleavage limited thickness of the plate, V-notch Charpy specimens (a
to shear with decreasing stress triaxiality. So far, the effect of central 45 V-notch of 2 mm depth and a 0.25 mm notch root
stress triaxiality on the temperature-dependent DBT of low- radius) of a 55  10  7 mm3 in size were used for three-point
carbon martensitic steels still remains elusive. bending tests, and crack propagation direction was perpen-
In addition to the primary mechanisms of void nucleation, dicular to the RD of the plate. In addition, all specimens were
growth and coalescence, ductile fracture may also be facili- extracted by wire-cut electrical discharge machining, and the
tated by potential shear localization even with no shear specimen surface was prepared by mechanical grinding (up to
component in the remote loading when experiencing pure 1200 SiC grit sandpaper) to control the surface roughness.
tensile loading [22]. The shear localization leads to a self-
focusing concentration of plastic strain, giving rise to pre- 2.2. Mechanical property tests
mature fracture [23]. On the other hand, cleavage fracture
surface, characterized by profuse facets, tends to be normal to 2.2.1. Tensile test
the maximum principal stress on a macroscopic scale. As Uniaxial tension tests were carried out at various tempera-
such, it can be expected that shear localization can be grad- tures (373 K, 297 K, 233 K, 193 K, 153 K, 133 K and 77 K) under
ually suppressed, accompanying the DBT until being roughly quasi-static loading (strain rate of 1.0  104 s1) by a universal
eliminated. For instance, Ji et al. [24] reported that ductile tensile machine (Zwick/Roell Z250) according to DIN EN ISO
fracture is triggered by shear localization along the maximum 6892-1/2/3 standards. The targeted temperature was achieved
shear stress plane (45 ) at temperatures from 293 K to 123 K, with induction heating or liquid nitrogen cooling, and a
while cleavage fracture takes place along the maximum thermocouple was attached to the specimen to monitor the
normal stress plane (90 ) at 77 K. However, the fracture angle surface temperature. For each temperature, at least two par-
mostly deviates from 45 when the ductile fracture occurs in a allel tests were conducted. All specimens were immersed in
localized shear deformation manner [25e27]. Jiang et al. [28] the targeted temperature environment for 15 min prior to the
studied the tensile properties of a medium-entropy alloy in a experiment. An extensometer along the length direction was
temperature range of 298e1273 K and found that the fracture applied to record the longitudinal strain during the tensile
angle could exceed 45 and varies with temperature, accom- deformation.
panied by the variation of microstructure. The macroscopic
fracture behavior is always sensitive to the underlying defor- 2.2.2. Three-point bending test
mation mechanism. To date, the DBT of low-carbon Dynamic three-point bending tests were conducted over a
martensitic steels has been well documented [5,29e34]. In temperature range of 133 Ke423 K at a loading rate of
contrast, few insightful investigations address the relation- 3.3  105 mm/min by a pendulum impact testing machine
ship between DBT and macroscopic fracture angle and un- (PSW 15/30) with a 300 J capacity following the ASTM E23
derlying microstructural variation. standard. Quasi-static three-point bending tests were carried
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3 1933

Table 1 e Chemical composition of the martensitic steel, wt %.


C Mn Si Cr Al Ni Ti Nb P Fe
0.173 1.420 0.407 0.347 0.041 0.029 0.027 0.013 0.012 Bal.

out on a 400 kN Instron servohydraulic tester at the same neighboring to the fracture surface of broken tensile speci-
temperatures, and a prescribed cross-head speed of 0.1 mm/ mens were analyzed with a transmission electron micro-
min was applied. Charpy specimens for dynamic and quasi- scope (TEM, FEI Talos F200X) operated at an accelerating
static tests were held in a thermal insulation chamber for voltage of 200 kV.
15 min in advance. Experiments were performed three times
for each testing condition to ensure accuracy.
3. Results
2.3. Microstructure characterization
3.1. Microstructures
Rolling-transverse plane of the specimens was ground,
polished, and etched with Nital reagent (3 wt% HNO3, 97 wt% The as-received low-carbon (<0.6 wt%) martensitic steel has a
ethanol). Metallic microstructure observation and typical fine lath substructure of the sub-millimeter scale, as
fractographic analysis were conducted with a scanning shown in Fig. 1a and b. Pole figure (PF) in Fig. 1c shows that
electron microscope (SEM, Carl Zeiss Microscopy Gmbh). An there is no pronounced crystallographic texture in the inves-
orientation map was collected using the SEM equipped with tigated material. Fig. 1d depicts a general view of kernel
a Channel 5 electron backscatter diffraction (EBSD) average misorientation (KAM) map with a striking feature that
system with a step length of 150 nm at 20 kV. Element rich coarse-lath domains exhibiting relatively lower KAM
identification of the inclusion was executed by an Oxford values are separated from the matrix. Besides, the coarse-lath
Instruments energy dispersive spectroscopy (EDS) detector. domain is softer than the surroundings due to its less dis-
To unravel the deformation mechanism, microstructures facets located, larger size and higher auto-tempering [35].

Fig. 1 e Microstructure of the material characterized by SEM and EBSD analysis: (a) lath martensitic microstructure; (b) IPF
orientation map; (c) (001), (110) and (111) PFs; (d) KAM map. RD and TD are rolling direction and transverse direction,
respectively.
1934 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3

Fig. 2 e Uniaxial tensile engineering stress-strain curves (a) and corresponding fracture angle (b) of the material tested at
373 K, 297 K, 233 K, 193 K, 153 K, 133 K and 77 K.

3.2. Tensile properties and fracture angle denotes a better toughness [37,38]. However, as listed in Table
2, the UTS  TE product of the material increases with
The engineering stress-strain curves of as-received low-car- decreasing temperature, which contradicts the well-
bon martensitic steel across a wide temperature range of understood DBT phenomena in bcc steels [2,33]. Therefore,
77e373 K are illustrated in Fig. 2a. The tensile properties, the macroscopic indicator (UTS  TE) is not a suitable variable
including yield strength (YS), ultimate tensile strength (UTS), to describe the temperature-dependent fracture resistance of
uniform elongation (UE) and total elongation (TE), are sum- the material.
marized in Table 2. As shown in Fig. 2a, all curves exhibit high The fracture angle q between the tensile direction and the
elastic limit, continuous yielding and weak strain hardening. fracture plane was examined for all tensile specimens at room
Both strength and uniform elongation are temperature- and cryogenic temperatures. The variation of fracture angle
dependent and tend to increase with decreasing tempera- with temperature is presented in Fig. 2b, along with the
ture, which is in alignment with the experimental findings in representative macroscopic failure modes of tensile samples.
other tempered martensitic steels [13,29,30]. When the testing It is depicted that the specimens tested at temperatures of
temperature decreases from 373 K to 77 K, the YS and the UTS 297 K, 233 K, 193 K and 153 K ruptured by shearing at different
are increased by 27.1% and 23.4%, respectively. The TE is fracture angles of about 65 , 67 , 68.5 and 70 , respectively,
progressively increased from 5.7% to 10.8% as the temperature rather than the maximum shear stress direction (45 ). This
drops from 373 K to 133 K. At 77 K, the TE is declined slightly to implies that both normal stress and shear stress affect the
9.9%. Therefore, no clear evidence of DBT is observed in the fracture behaviors in these cases, and with decreasing tem-
martensitic steel when only the tensile ductility is considered. perature, the normal stress effect is enhanced since the de-
It is also worth noting that the serrated stress-strain curve at viation of the fracture angle from 45 becomes larger.
373 K demonstrates the advent of the dynamic strain aging Furthermore, at temperatures of 133 K and 77 K, the fracture
(DSA) phenomenon, which is prevailing in alloys containing a plane is almost perpendicular to the tensile direction with
low concentration of carbon in the temperature range of fracture angles of about 89.5 and 89 , respectively. This in-
approximately 373 Ke573 K [36]. The toughness of sheet dicates that the tension (mode I), instead of the shear (mode
metals can be represented by a combination of tensile II), dominates the specimen rupture. However, due to the
strength and total elongation, and a larger product of UTS TE significant plastic deformation and slight shear localization, a
small deviation from 90 exists in the fracture angle at 133 K
and 77 K, which can be attributed to the difference of local
deformation behavior such as necking. Additionally, the shear
Table 2 e Mechanical properties of the investigated
material at different temperatures. trace marked with a black dashed line on the specimen at
133 K in Fig. 2b further corroborates the existence of the
Temperature, K YS, MPa UTS, MPa UE, % TE, % UTS  TE,
competition between normal stress and shear stress during
GPa‧%
tensile deformation until fracture. Accordingly, the transition
373 1194 1372 3.3 5.7 7.8
of failure mode from shear-to tension-dominated can be well
297 1240 1411 3.7 7.3 10.3
233 1287 1447 4.0 8.4 12.2
captured by the change in fracture angle of tensile specimens
193 1297 1458 4.6 8.9 13.0 with decreasing temperature. The correlation with underlying
153 1348 1483 4.8 9.5 14.1 failure mechanisms will be discussed in the following. Note
133 1371 1498 5.1 10.8 16.2 that, due to the DSA effect at 373 K, the shear trace along the
77 1518 1693 5.7 9.9 16.8 thickness direction is also visible in Fig. 2b. The detailed and
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3 1935

Fig. 3 e Force-displacement curves (upper) and energy absorption (lower) of dynamic (a, c) and quasi-static (b, d) three-point
bending tests covering the temperature range from 133 K to 423 K.

systematic investigation of this unique spatial failure mode is force reversal site occurs, i.e., the point where dF=ds ¼ 0 (dF
not carried out as the macroscopic failure modes at elevated and ds are the increments of force and displacement,
temperatures are beyond the scope of this study. respectively) [39]. For dynamic tests, the specimens exhibit no
significant plastic deformation up to 273 K and rupture
3.3. Three-point bending properties and energy abruptly, showing a brittle fracture behavior without a force
absorption reversal site. As the temperature rises to 297 K, the curve
preceding an abrupt drop in force depicts a nonlinear ten-
Representative force-displacement curves of three-point dency, which implies the fact that the specimen deforms into
bending tests performed at dynamic and quasi-static loading the ductile domain, and the DBTT range should be between
rates are shown in Fig. 3a and b. Regarding the V-notch Charpy 273 and 297 K. For quasi-static tests, a pronounced bending
specimen, one criterion for distinguishing brittle and ductile curve at 273 K suggests enhanced plastic deformation of the
fracture from the force-displacement curve is whether the specimen. According to the same criterion, the DBTT under
1936 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3

quasi-static loading is also preliminarily determined to be specimens in Fig. 4aec, and as the temperature decreases
273e297 K. In addition, insets a1 and b1 indicate the force from 297 K to 193 K, the equiaxed dimples decrease in size.
required for crack propagation reaches its peak value at Note that the shear dimples are generally regarded as evi-
temperatures close to the transition temperature domain. To dence of shear fracture [22]. In the more cryogenic case
the best of the authors’ knowledge, increasing loading rates (Fig. 4d), the fracture surface is mainly characterized by dim-
can significantly shift the DBTT toward higher values ples, while a small number of cleavage facets are also visible.
[14,40,41]. However, no drastic effect of the loading rate on the In contrast, at 133 K in Fig. 4e, the fracture surface is domi-
DBTT is found in the investigated material, which may be nated by equiaxed dimples and numerous cleavage facets.
ascribed to the weak strain rate sensitivity on failure mecha- The fracture surface at 77 K is composed of almost entirely
nisms of the full lath martensite structure. Further fracture cleavage facets (Fig. 4f), despite that a large amount of plastic
morphology inspection will be presented in section 3.4 to gain flow occurs at this extremely low temperature. In light of the
a better insight into the transition temperature. fractography studies, pure cleavage fracture occurs at 77 K,
An important perspective in evaluating the macroscopic and pure ductile fracture occurs at temperatures above 193 K.
mechanical properties of steels is energy absorption. On the 133 K and 153 K are located within the DBTT domain in tensile
basis of the maximum force, fracture total energy (Et ) can be tests. The observed change in macroscopic shear to tension
divided into two parts, fracture initiation energy (Ei ) and failure modes, characterized by the fracture angle as depicted
fracture propagation energy (Ep ), which are calculated by in Fig. 2b, is consistent with the transition of underlying fail-
integrating the areas under the force-displacement curves ure mechanisms. Hence, instead of relying on tensile ductility,
up to the maximum force and from the maximum force to the macroscopic fracture angle is a simple and reliable indi-
failure, respectively. The obtained energy absorption at cator revealing the temperature-dependent failure mecha-
different temperatures is reported in Fig. 3c and d. The ratios nisms in tensile tests.
of Ei =Et and Ep =Et are also shown for different temperatures In the case of the Charpy specimens ruptured under dy-
under both loading conditions. Apparently, well-defined namic and quasi-static loadings, the insets in Fig. 5aeh shows
lower and upper shelf plateaus and a transition tempera- the overview fractographs of specimens ruptured at different
ture range are observed at the dynamic loading rate. With temperatures. Fig. 5a and e exhibit ductile fracture with deep
raising temperature, Et and Ep increase, while Ei = Et ratio dimples at 373 K, whilst the specimens tested at 297 K mainly
decreases. Under quasi-static loading, the transition tem- show ductile fracture with shallow dimples in Fig. 5b and f.
perature range is ambiguous from the viewpoint of energy Besides, the fractographs of the specimens tested at 273 K and
absorption. Additionally, it is to be expected that a lower 133 K for both loading conditions are similar, showing brittle
upper shelf plateau shall be obtained in comparison with the fracture with cleavage facets surrounded by tear ridges with
dynamic loading case. However, no specific upper shelf different lengths (Fig. 5c, d, g and h). It is known that the
plateau is observed here because of the decrease in Et at deeper or larger the dimple, the more energy is consumed
temperatures above 323 K. This is mainly associated with the during fracture, which suggests better toughness. Thus, as can
occurrence of DSA at elevated temperatures, rendering the be seen in Fig. 5a and e, the toughness under dynamic loading
progressive embrittlement of the material [42]. Specifically, is superior to that under quasi-static loading because deeper
in Fig. 3d, Ep value increases and then decreases, whilst Ei = Et and larger dimples are observed in the former case, which can
ratio declines from 1 to 0.61 as the temperature increases be attributed to the synergistic effects of loading rate and DSA
from 133 K to 323 K, and then gradually increases to 0.82 at elevated temperatures. Note that a low loading rate can
when the temperature further increases to 423 K, indicating weaken ductile resistance capacity in the upper shelf plateau
that the DSA effect accelerates the crack propagation and domain, and DSA has a negative influence on toughness
eventually deteriorates the impact toughness. The higher properties at temperatures above 323 K. In Fig. 5c, an isolated
crack propagation velocity is actually caused by the deteri- dimple region surrounded by profuse cleavage facets is trig-
oration of the average dislocation slip rate owing to the gered by a small globular inclusion. Further analysis by EDS
interaction of mobile dislocations with C atoms in a highly measurement reveals that the inclusion combines two oxides
dynamic way in the DSA domain [43]. Noteworthy, the (Al2O3, MgO2) and a sulfide (CaS), as demonstrated in Fig. 5i,
peculiar feature of DSA is high temperature and negative indicating that such inclusions in the investigated material
strain rate sensitivities, and this is the reason why the DSA are active sites for void nucleation below the transition tem-
effect is pronounced at temperatures above 323 K and pretty perature. Based on the variation of fracture morphology, the
feeble under dynamic loading in this study. DBTT of the investigated material in Charpy specimens is
further confirmed between 273 K and 297 K, regardless of the
3.4. Fracture morphologies loading rate.

To determine the DBTT of low-carbon martensitic steel in 3.5. Microstructural variation


terms of failure mechanisms, SEM fractographs of tensile
specimens tested over a wide temperature range of Post-mortem TEM foils were extracted near the fracture sur-
297 Ke77 K are reported in Fig. 4aef. At each temperature, faces to study the microstructural variation of the tensile
high-magnification images are taken from different positions specimens at different temperatures. Fig. 6a shows a typical
(1 and 2), where position 1 is in the vicinity of the crack initi- bright-field (BF) TEM lath martensite structure with a low
ation site, as indicated by the white circles in Fig. 4. Both shear dislocation density of the specimen tensile deformed at 297 K.
dimples and equiaxed dimples are observed in tensile A well-defined narrow shear band with ~660 nm in width is
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3 1937

Fig. 4 e SEM images of tensile specimen ruptured at (a) 297 K, (b) 233 K, (c) 193 K, (d) 153 K, (e) 133 K and (f) 77 K. Images in (1)
and (2) are the corresponding regions at a high magnification.

also visible along the direction inclined to the lath boundaries, profuse dislocation cells with hundreds of nanometers in
as marked by the parallel white dashed line in Fig. 6b. The Fig. 6g, and the dislocation cells present a homogeneous dis-
formation of the shear band traverses the original lath tribution over the entire area, which is similar to the other
martensite structure, introducing a significant heterogeneity observations [45,46]. It is evident from Fig. 6h that the dislo-
within the associated martensite colony by generating a high cation cell is characterized by a cell interior with a low dislo-
density of dislocations around the shear band (Fig. 6b). This cation density, whereas a cell wall is piled up with plentiful
will, in turn, induce the building of large strain gradients along dislocations. Also, one can observe Moire  fringes with sharp
the shear band and the vindicated domains against further interfaces tangential to the dislocation cell, as marked by the
deformation, giving rise to the formation of strain localization red dashed lines, which demonstrates local translational and
and thus the catastrophic shear fracture [44]. When the tem- rotational changes of the lattice fringes during the formation
perature drops to 193 K, as indicated in Fig. 6c, the lath of the dislocation cell [47].
structure is still discernible, with extensive dislocation pile-
ups at lath boundaries. Furthermore, to accommodate the
plastic deformation, intersecting shear bands are observed in 4. Discussion
Fig. 6d, resulting in a kinked section with ~1000 nm in length
at the intersection domain length. At 133 K, most of the lath 4.1. Temperature dependence of fracture properties
boundaries become fuzzy and cannot be distinguished from
the inner tangled dislocations and dislocation cells, as shown The true fracture strain (εf ), derived from the area reduction at
in Fig. 6e and f. A further decrease in temperature to 77 K could fracture can be evaluated according to Eq. (1) for all the flat
result in a change in dislocation morphology by generating tensile specimens,
1938 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3

Fig. 5 e SEM images of Charpy specimens tested under (a)e(d) dynamic and (e)e(h) quasi-static loadings at 373 K, 297 K,
273 K and 133 K, respectively; (i) EDS analysis of the inclusions in the area enclosed by the red dashed line in (c).

    et al. [48] proposed that the sf can be obtained by extrapolating


A0 w0 ,t0
εf ¼ ln ¼ ln (1) the flow curve for before-necking strain with the Swift strain
Af wf ,tf
hardening law of sðεp Þ ¼ Aðεp þ ε0 Þn . The symbols sðεp Þ, A, εp ,
where A0 and Af are the tensile specimen cross-section areas ε0 and n denote true stress, strengthening coefficient, true
before and after testing, w0 and t0 are the initial width and strain, material constant and strain hardening exponent,
thickness of the gauge section, and wf and tf are the specimen respectively. Fig. 7aef displays the experimental results and
width and thickness at fracture, respectively. To compensate the true strain-stress curves up to fracture strain extrapolated
for the necking effects, the tf is obtained by a multiple- by the Swift strain hardening law upon different tempera-
thickness approach according to the ASTM E8 standard. The tures. The calibrated parameters A, ε0 and n are given in Fig. 7.
averaged εf value at each temperature is listed in Table 3. It is Both A and n increase with declining temperature, while ε0
found that true fracture strain decreases with decreasing keeps constant. Accordingly, the sf can be obtained and is also
temperature because of the lessening of the shear localiza- summarized in Table 3.
tion, which is a trend opposite to that of the global parameter, Within the temperature range of 77 Ke297 K, the fracture
total elongation, in Table 2. This suggests the material tends to angles are in the range between 45 and 90 . Following the
deform more uniformly with decreasing temperature. Ellipse criterion proposed by Zhang et al. [49], macroscopic
At such a large true fracture strain, it is inaccurate to obtain failure modes of high-strength materials (bulk metallic
the fracture strength (sf ) by dividing the fracture load by the glasses, steels and nanostructured materials) can be described
final cross-section area (Af ) due to necking. Hence, Kamaya [25,26,50e52], which is expressed as
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3 1939

Fig. 6 e BF-TEM images showing the microstructural variation: (a) lath martensite structure with low dislocation density and
(b) shear band formed at 297 K; (c) lath martensite structure with high dislocation density and (d) intersecting shear bands
formed at 193 K; (e)e(f) ambiguous lath boundaries and dense dislocation tangles formed at 133 K; (g) dislocation cells and
 fringes deformed at 77 K.
(h) a typical dislocation cell with Moire

remarkably. The results revealed in Figs. 2 and 8 are also


 2  2 illustrated schematically in the first quadrant in the inset of
s t
þ ¼1 (2) Fig. 8a for clarity, elucidating the evolution of the Mohr's circle
s0 t0
(blue circles) and the Ellipse fracture locus (solid curve), i.e.,
where s0 and t0 are the critical normal and shear stresses, and Eq. (2), with decreasing temperature under uniaxial tension.
s and t are the normal and shear stresses on the fracture For instance, the Mohr's circle touches the right boundary of
plane. Considering the uniaxial tension condition, the critical the Ellipse fracture locus at low temperature (at T3), while the
fracture condition assumed by the Ellipse criterion is that the failure point is more close to the shear stress axis at high
Mohr's circle touches the fracture locus in the first quadrant. temperature (at T1). As the discrepancy between s0 and t0
Thus, the critical Mohr's circle can be formulated as decreases with decreasing temperature, the continuously
 2 enhanced normal stress sensitivity leads to a shift of the
s ¼ sf sin q (3) macroscopic failure modes from the shear fracture to tension
t ¼ sf sin q cos q
fracture.
In addition, the failure mode factor is defined as a ¼ t0 =s0 .
Combining Eqs. (2) and (3), a can be deduced as a function of 4.2. Temperature dependence of deformation
fracture angle q and critical shear stress t0 is related to sf and mechanisms and failure modes
a,
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Temperature has an essential effect on the failure mode. In
1 
the present work, the ductile fracture in low-carbon
a¼ 1  cot2 q (4)
2
martensitic steel is promoted by shear localization
sf
t0 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi (5)
2 1  a2
In light of the measured fracture angle q (Fig. 2b) and Table 3 e True fracture strain and stress of the tensile
calculated tensile fracture strength sf (Fig. 7), the change in a specimens upon different temperatures.
with temperature is shown in Fig. 8a. Note that the value of a Fracture strain and stress Temperature, K
becomes significantly large below the transition point (153 K).
297 233 193 153 133 77
The variation of s0 and t0 as a function of temperature is
further plotted in Fig. 8b. It is explicit that s0 and t0 exhibit the εf , - 0.885 0.852 0.804 0.760 0.603 0.501
sf , MPa 1610 1670 1706 1735 1753 1998
same trend, and below 153 K, both s0 and t0 increase
1940 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3

Fig. 7 e Experimental true stress-strain curves of the material and fitted curves (by Swift strain hardening law) at different
temperatures: (a) 297 K, (b) 233 K, (c) 193 K, (d) 153 K, (e) 133 K and (f) 77 K.

associated with shear bands (Fig. 6b and d). Martensite has dynamic and quasi-static bending loadings due to the loading
an inherently high ratio of ~18% of mobile screw dislocations rate and DSA effects. However, the transition of underlying
from displacive transformations [53]. In Fig. 6, the absence of failure mechanisms in Fig. 9a exhibits a weak strain rate
the lath feature indicates that the catastrophic release and sensitivity. Likewise, the void-controlled ductile fracture re-
movement of screw dislocations in the dislocation sub- mains the dominant failure mechanism even with super-
structures from the lath walls lead to the complete collapse imposed DSA effects. The significant difference in the
and break of the original lath boundaries and eventually transition temperature between tensile and bending tests is
refine the martensitic laths into the stable dislocation thus ascribed to the specimens with distinguished intrinsic
cells at 77 K for minimizing the total system energy. stress states.
Prior studies have revealed that dislocation activity is The stress state has a significant impact on the fracture
sensitive to temperature, and the stable dislocation cell is a properties of steels. In this study, compared with the uniaxial
result of the balance between dislocation nucleation and tension (h ¼ 0:33), the thicker V-notch Charpy specimen has
dislocation movement at lower temperatures [54,55]. plane strain condition with a much higher stress triaxiality
From the current results, these changes within the micro- (h ¼ 1:48) [20], which can greatly restrict the plastic defor-
structure can restrict the effect of shear instability on the mation at the crack tip, inducing the higher local principal
failure of the investigated material, rendering a failure mode tensile stress. Therefore, the DBT in Charpy specimens shifts
from shear to tension, i.e., the advent of DBT with decreasing upward to a higher temperature. However, for quantitatively
temperature. shedding light on the relationship between stress triaxiality
and DBTT, more specimens with various geometries need to
4.3. Stress-state-dependent transition of failure be involved, which is the subject of our ongoing research.
mechanisms
4.4. Toughness indicators under uniaxial tension
The change in area fraction of ductile domain with tempera-
ture under different testing configurations is further quanti- It is known that global tensile ductility (uniform elongation or
tatively analyzed in Fig. 9a. The DBTT can be determined by total elongation) is not a reliable index for toughness assess-
the temperature at which the dimple area occupies 50% of the ment because the toughness primarily depends on the local
fracture surface [56]. Therefore, the DBTT is approximately fracture resistance. Since the UTS  TE product mentioned in
determined as 140 K in tensile tests, while the DBTT is 282 K in section 3.2 is derived from global tensile properties, it is
three-point bending tests. At temperatures above 323 K, the also not a good indicator to evaluate the toughness. Some
fracture total energy shows a clear difference between investigations suggested that the true fracture strain obtained
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3 1941

Fig. 8 e (a) Failure mode factor a versus temperature; (b) critical normal stress s0 and critical shear stress t0 versus
temperature.

Fig. 9 e (a) Change in area fraction of ductile domain versus temperature under different testing configurations; (b) Change
in normalized energy absorption versus temperature under different stress states.

from tensile specimens can be used as an index for toughness


estimation [57,58]. However, the true fracture strain only 5. Conclusions
considers the local ductility while overlooking the aspects of
strength and energy absorption capacity. Thus, from the en- The deformation and fracture behaviors of a low-carbon
ergy absorption viewpoint, the local fracture energy El can be martensitic steel were multiscale studied. The effects of
Rε loading rate and stress state on the DBTT were clarified. The
defined as El ¼ 0 f st dε, where εf , st and dε are true fracture
main conclusions can be summarized as follows.
strain, true stress and the increment of true strain, respec-
tively. The temperature dependence of the normalized tensile
 Both tensile strength and ductility of the low-carbon
energy is shown in Fig. 9b, which is calculated by taking El at
martensitic steel are improved with a decrease in testing
RT as reference. It is clear that the local fracture energy
temperature from 373 K to 77 K.
parameter captures the temperature dependence of the
 The deformation mechanisms show a clear
toughness. When compared with the normalized energy ab-
temperature dependence, which is reflected in
sorption in Charpy specimens under dynamic loading, the
microstructural variation from the formation of shear
pronounced stress state dependence of the DBTT is also
bands to stable dislocation cells with decreasing temper-
observed. In summary, the macroscopic fracture angle, the
ature. The change in defect structures is correlated to the
area fraction of dimples/facets on the fracture surface, and the
transition of macroscopic failure mode from shear to
proposed local fracture energy are all efficient indicators to
tension.
describe the toughness properties under uniaxial tension.
1942 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3

 Toughness cannot be directly inferred from tensile tests [3] Lu Y, Zhang YH, Ma E, Han WZ. Relative mobility of screw
by means of the conventional global indicator (UTS  TE), versus edge dislocations controls the ductile-to-brittle
whilst the local fracture energy can capture the transition in metals. Proc Natl Acad Sci USA
2021;118(37):e2110596118.
temperature-dependent toughness properties. Furthermore,
[4] Griffith AA. The phenomena of rupture and flow in solids.
the macroscopic fracture angle is proven efficient in Philos Trans R Soc A 1921;221(582e593):163e98.
revealing the underlying failure mechanisms in tensile tests. [5] Sharma T, Kumar NN, Mondal R, Krishna KVM, Samajdar I,
 The activation of the DSA at elevated temperatures has a Kain V. Ductile-to-brittle transition in low-alloy steel: a
negative influence on energy absorption in quasi-static combined experimental and numerical investigation. J Mater
three-point bending tests, rendering a decrease in energy Eng Perform 2019;28(7):4275e88.
absorption with increasing temperature and thus without [6] Gao XS, Zhang GH, Srivatsan TS. Prediction of cleavage
fracture in ferritic steels: a modified Weibull stress model.
an upper shelf plateau.
Mater Sci Eng A 2005;394(1e2):210e9.
 The DBTT is significantly affected by stress states, while [7] Krauss G. Martensite in steel: strength and structure. Mater
the strain rate effect is weak. The transition temperature Sci Eng A 1999;273:40e57.
is shifted toward higher values with increasing stress [8] Tao NR, Wang ZB, Tong WP, Sui ML, Lu J, Lu K. An
triaxiality. investigation of surface nanocrystallization mechanism in Fe
induced by surface mechanical attrition treatment. Acta
Mater 2002;50(18):4603e16.
Author contributions [9] Orowan E. Fracture and strength of solids. Rep Prog Phys
1949;12(1):185e232.
[10] Benzerga AA, Tvergaard V, Needleman A. Size effects in the
Zhen Zhang: Data curation, Formal analysis, Investigation,
Charpy V-notch test. Int J Fract 2002;116(3):275e96.
Methodology, Software, Validation, Visualization, Writing e [11] Pineau A. Modeling ductile to brittle fracture transition in
original draft. steelsdmicromechanical and physical challenges. Int J Fract
Fuhui Shen: Conceptualization, Supervision, Project 2008;150(1):129e56.
administration, Formal analysis, Methodology, Validation, [12] Duan QQ, Qu RT, Zhang P, Zhang ZJ, Zhang ZF. Intrinsic
Visualization, Writing e review & editing. impact toughness of relatively high strength alloys. Acta
Hanqing Liu: Formal analysis, Methodology, Validation, Mater 2018;142:226e35.
[13] Spa€ tig P, Bonade R, Odette GR, Rensman JW, Campitelli EN,
Visualization, Writing e review & editing.
Mueller P. Plastic flow properties and fracture toughness
Markus Ko € nemann: Resources, Funding acquisition,
characterization of unirradiated and irradiated tempered
Writing e review & editing. martensitic steels. J Nucl Mater 2007;367:527e38.
Sebastian Münstermann: Supervision, Resources, Funding [14] Barsom JM, Rolfe ST. Correlation between KIC and Charpy V-
acquisition, Writing e review & editing. notch test results in the transition temperature range. In:
Impact testing of metals, STP 466. Philadelphia: ASTM; 1970.
p. 281e302.
Declaration of competing interest [15] Zhu XK. State-of-the-art review of fracture control
technology for modern and vintage gas transmission
The authors declare that they have no known competing pipelines. Eng Fract Mech 2015;148:260e80.
[16] Shen FH, Münstermann S, Lian JH. Investigation on the
financial interests or personal relationships that could have
ductile fracture of high-strength pipeline steels using a
appeared to influence the work reported in this paper. partial anisotropic damage mechanics model. Eng Fract
Mech 2020;227:106900.
[17] Pütz F, Shen FH, Ko € nemann M, Münstermann S. The
Acknowledgment differences of damage initiation and accumulation of DP
steels: a numerical and experimental analysis. Int J Fract
2020;226:1e15.
The financial support for the project (IGF.-Nr. 21094N/FOSTA
[18] Mu L, Jia Z, Ma ZW, Shen FH, Sun YK, Zang Y. A theoretical
P1359) by the German Federal Ministry for Economic Affairs prediction framework for the construction of a fracture
and Energy based on a decision taken by the German Bun- forming limit curve accounting for fracture pattern
destag is greatly acknowledged. The authors would like to transition. Int J Plast 2020;129:102706.
thank the “Forschungsvereinigung Stahlanwendung e.V. [19] Shen FH, Wang HS, Liu ZJ, Liu WQ, Ko € nnemann M, Yuan G,
(FOSTA)” and the “Arbeitsgemeinschaft industrieller For- Wang G, Münstermann S, Lian JH. Local formability of
schungsvereinigungen e.V.” for their support. Z. Zhang would medium-Mn steel. J Mater Process Technol 2022;299:117368.
[20] Shen FH, Münstermann S, Lian JH. A unified fracture
like to acknowledge the financial support from China Schol-
criterion considering stress state dependent transition of
arship Council (Grant No. 202006240048). failure mechanisms in bcc steels at e196 C. Int J Plast
2022;156:103365.
[21] Shen FH, Münstermann S, Lian JH. Cryogenic ductile and
references
cleavage fracture of bcc metallic structureseInfluence of
anisotropy and stress states. J Mech Phys Solid
2023;176:105299.
[1] Tomita Y, Okabayashi K. Effect of microstructure on strength [22] Pineau A, Benzerga AA, Pardoen T. Failure of metals I: brittle
and toughness of heat-treated low alloy structural steels. and ductile fracture. Acta Mater 2016;107:424e83.
Metall Trans A 1986;17(7):1203e9. [23] Azizi H, Zurob HS, Embury D, Wang X, Wang K, Bose B. Using
[2] Christian JW. Some surprising features of the plastic architectured materials to control localized shear fracture.
deformation of body-centered cubic metals and alloys. Acta Mater 2018;143:298e305.
Metall Trans A 1983;14(7):1237e56.
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 1 9 3 1 e1 9 4 3 1943

[24] Ji XL, An Q, Xia YP, An R, Zheng R, Wang CQ. Maximum shear precracked Charpy specimens. Int J Pres Ves Pip
stress-controlled uniaxial tensile deformation and fracture 2008;85(11):752e61.
mechanisms and constitutive relations of SnePb eutectic [42] Kim IS, Kang SS. Dynamic strain aging in SA508-class
alloy at cryogenic temperatures. Mater Sci Eng A 3 pressure vessel steel. Int J Pres Ves Pip 1995;62(2):123e9.
2021;819:141523. [43] Caillard D, Bonneville J. Dynamic strain aging caused by a
[25] Dong FY, Zhang P, Pang JC, Ren YB, Yang K, Zhang ZF. new Peierls mechanism at high-temperature in iron. Scripta
Strength, damage and fracture behaviors of high-nitrogen Mater 2015;95(1):15e8.
austenitic stainless steel processed by high-pressure torsion. [44] Xu YB, Zhang JH, Bai YL, Meyers MA. Shear localization in
Scripta Mater 2015;96:5e8. dynamic deformation: microstructural evolution. Metall
[26] Qu RT, Zhang ZF. A universal fracture criterion for high- Mater Trans A 2008;39(4):811e43.
strength materials. Sci Rep 2013;3(1):1e6. [45] Morito S, Ohba T, Maki T. Comparison of deformation
[27] Jiang MQ, Wilde G, Chen JH, Qu CB, Fu SY, Jiang F, Dai LH. structure of lath martensite in low carbon and ultra-low
Cryogenic-temperature-induced transition from shear to carbon steels. Mater Sci Forum 2007;558:933e8.
dilatational failure in metallic glasses. Acta Mater [46] Chatterjee A, Sk MB, Ghosh A, Mitra R, Chakrabarti D.
2014;77:248e57. Combining crystal plasticity and electron microscopy to
[28] Jiang W, Yuan SY, Cao Y, Zhang Y, Zhao YH. Mechanical elucidate texture dependent micro-mechanisms of tensile
properties and deformation mechanisms of a deformation in lath martensitic steel. Int J Plast
Ni2Co1Fe1V0.5Mo0.2 medium-entropy alloy at elevated 2022;153:103251.
temperatures. Acta Mater 2021;213:116982. [47] Zhang Z, Hu ZF, Schmauder S, Zhang BS, Wang ZZ. Low cycle
[29] Dudko V, Fedoseeva A, Kaibyshev R. Ductile-brittle transition fatigue properties and microstructure of P92 ferritic-
in a 9% Cr heat-resistant steel. Mater Sci Eng A martensitic steel at room temperature and 873 K. Mater Char
2017;682:73e84. 2019;157:109923.
[30] Morris Jr JW. On the ductile-brittle transition in lath [48] Kamaya M, Kawakubo M. A procedure for determining the
martensitic steel. ISIJ Int 2011;51(10):1569e75. true stressestrain curve over a large range of strains using
[31] Morris Jr JW, Kinney C, Pytlewski K, Adachi Y. Microstructure digital image correlation and finite element analysis. Mech
and cleavage in lath martensitic steels. Sci Technol Adv Mater 2011;43(5):243e53.
Mater 2013;14:014208. [49] Zhang ZF, Eckert J. Unified tensile fracture criterion. Phys Rev
[32] Wang CF, Wang MQ, Shi J, Hui WJ, Dong H. Effect of Lett 2005;94(9):094301.
microstructural refinement on the toughness of low carbon [50] Wang B, Zhang P, Duan QQ, Zhang ZJ, Yang HJ, Pang JC,
martensitic steel. Scripta Mater 2008;58(6):492e5. Tian YZ, Li XW, Zhang ZF. High-cycle fatigue properties and
[33] Sakamaki T, Tanaka M, Morikawa T, Hidenori Nako, damage mechanisms of pre-strained Fe-30Mn-0.9C twinning-
Nanba S. Brittle-to-ductile transition in martensiteebainite induced plasticity steel. Mater Sci Eng A 2017;679:258e71.
Steel. ISIJ Int 2021;61(7):2167e75. [51] Zhao YH, Liu JZ, Topping TD, Lavernia EJ. Precipitation
[34] Chatterjee A, Ghosh A, Moitra A, Bhaduri AK, Mitra R, and aging phenomena in an ultrafine grained Al-Zn alloy
Chakrabarti D. Role of hierarchical martensitic by severe plastic deformation. J Alloys Compd
microstructure on localized deformation and fracture of 9Cr- 2021;851:156931.
1Mo steel under impact loading at different temperatures. [52] Gao WJ, Zhang WW, Zhang T, Yang C, Huang XS, Liu ZY,
Int J Plast 2018;104:104e33. Wang Z, Li WH, Li WR, Li L, Liu LH. Large tensile plasticity in
[35] Morsdorf L, Jeannin O, Barbier D, Mitsuhara M, Raabe D, Zr-based metallic glass/stainless steel interpenetrating-
Tasan CC. Multiple mechanisms of lath martensite plasticity. phase composites prepared by high pressure die casting.
Acta Mater 2016;121:202e14. Compos B Eng 2021;224:109226.
[36] Caillard D, Bonneville J. Dynamic strain aging caused by a [53] Sandvik BPJ, Wayman CM. Characteristics of lath martensite:
new Peierls mechanism at high-temperature in iron. Scripta Part I. Crystallographic and substructural features. Metall
Mater 2015;95:15e8. Trans A 1983;14(4):809e22.
[37] Fro meta D, Lara A, Grife L, Dieudonne  T, Dietsch P, Rehrl J, [54] Khantha M, Pope DP, Vitek V. Dislocation screening and the
Suppan C, Casellas D, Calvo J. Fracture resistance of brittle-to-ductile transition: a Kosterlitz-Thouless type
advanced high-strength steel sheets for automotive instability. Phys Rev Lett 1994;73(5):684e7.
applications. Metall Mater Trans A 2021;52(2):840e56. [55] Akama D, Tsuchiyama T, Takaki S. Change in dislocation
[38] Xiong ZP, Jacques PJ, Perlade A, Pardoen T. characteristics with cold working in ultralow-carbon
Characterization and control of the compromise martensitic steel. ISIJ Int 2016;56(9):1675e80.
between tensile properties and fracture toughness in a [56] Jia QX, Wang YX, Mei RX, Chen L, Hao S, Zhang HY, Ma XC,
quenched and partitioned steel. Metall Mater Trans A Zou ZY, Jin M. The dependences of deformation temperature
2019;50(8):3502e13. on the strain-hardening characteristics and fracture
[39] Bruchhausen M, Holmstro € m S, Lapetite JM, Ripplinger S. On behavior of MneN bearing lean duplex stainless steel. Mater
the determination of the ductile to brittle transition Sci Eng A 2021;819:141440.
temperature from small punch tests on Grade 91 ferritic- [57] Hance BM. Advanced high strength steel: deciphering local
martensitic steel. Int J Pres Ves Pip 2017;155:27e34. and global formability. Int Automot Body Congr 2016;1(2).
[40] Rolfe ST, Barsom JM. Fracture and fatigue control in Dearborn, MI.
structures: applications of fracture mechanics. ASTM [58] Gao C, Wang YC, Cheng XW, Li ZY, Zhang HC, Cai HN.
International; 1977. Excellent tensile local deformability and toughness of a
[41] Chaouadi R, Puzzolante JL. Loading rate effect on tempered low carbon-low alloy steel with multilevel
ductile crack resistance of steels using boundaries. Mater Sci Eng A 2022;844:143195.

You might also like