Materials Science & Engineering A: Sciencedirect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials Science & Engineering A 735 (2018) 448–455

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Influence of vanadium on the hydrogen embrittlement of aluminized ultra- T


high strength press hardening steel

Lawrence Choa, ,1, Eun Jung Seoa,2, Dimas H. Sulistiyoa, Kyoung Rae Joa, Seong Woo Kimb,
Jin Keun Ohb, Yeol Rae Chob, Bruno C. De Coomana,3
a
Graduate Institute of Ferrous Technology, Pohang University of Science and Technology, Pohang 37673, South Korea
b
Technical Research Laboratories, POSCO, Gwangyang 57807, South Korea

A R T I C LE I N FO A B S T R A C T

Keywords: Improved safety standards and reduced automotive body-in-white weight have led to a strong interest in mar-
Hydrogen embrittlement tensitic press hardening steel (PHS). As the sensitivity to hydrogen embrittlement of martensite increases at
Press hardening steel higher strength level, the very small uptake of diffusible hydrogen by aluminized PHS during the austenitization
Aluminizing stage of the hot press forming process is of concern. The hydrogen uptake was found to reduce the plasticity of
Vanadium
the PHS considerably. The PHS with a higher strength was more susceptible to hydrogen-induced brittle fracture.
The present work reports that vanadium additions, which serve to trap the hydrogen, considerably reduce the
negative impact of the hydrogen uptake in aluminized 1800 MPa and 2000 MPa PHSs. The contribution also
proposes a mechanism for the uptake of hydrogen during the processing of aluminized PHS.

1. Introduction the PHSs is associated with high temperature atmospheric corrosion by


water vapor [4–6]. Al-Si alloy coated (aluminized) PHS is more sensi-
Improved safety standards and reduced automotive body-in-white tive to the hydrogen uptake and resultant hydrogen-induced cracking
weight have led to a strong interest in advanced high strength/ultra- than the uncoated PHS. Fig. 1(b) shows that 0.50 mass ppm of the
high strength steels. The use of ultra-high strength press hardening steel diffusible hydrogen was detected in an aluminized PHS after the in-
(PHS) in structural safety-related parts has experienced a rapid growth dustrial HPF process. A hydrogen desorption peak at 170 °C was clearly
in the automotive industry [1,2]. The requirement of an ultra-high observed in the thermal desorption analysis (TDA) curve for the alu-
strength has, however, led to concerns about the hydrogen embrittle- minized PHS. In contrast, there is no hydrogen-uptake in the uncoated
ment as martensitic steel is increasingly sensitive to hydrogen-induced PHS. The ductility of the aluminized PHS was clearly reduced due to the
cracking at high strength levels. It is now well known that the diffusible hydrogen embrittlement (Fig. 1(c)). The hydrogen-induced cracking
hydrogen which diffuses out of the steel at room temperature, is re- noted for the PHSs exhibits characteristics very similar to the “con-
sponsible for the observation of a transient hydrogen-induced embrit- ventional” embrittlement phenomena in that the presence of hydrogen
tlement [3]. in the PHS causes a transition from ductile microvoid coalescence to
In the direct hot press forming (HPF) process, the press-hardened brittle intergranular fracture, accompanied by a premature failure and
components are produced by heating cold-rolled or coated steel blanks that the resulting loss of ductility is of a time-dependent, transient
to an austenitizing temperature in a furnace and subsequently nature [6].
quenching them between water-cooled quenching dies (Fig. 1(a)). Al- Recent investigations have shown that the bake hardening heat
10%Si alloy coating is the coating system that is most commonly used treatment reduces the diffusible hydrogen content of aluminized PHSs
for PHSs. [4–6]. There is however still a potential risk for the hydrogen-induced
It has been reported that PHSs absorb diffusible hydrogen during cracking if a forming process is applied to the aluminized PHS during
the austenitization in the HPF process [4–6]. The hydrogen uptake of vehicle assembly prior to the paint coating. Georges et al. [4] reported


Corresponding author.
E-mail address: lawrence.cho@nist.gov (L. Cho).
1
Current address: National Institute of Standards and Technology, Boulder, CO 80305, USA.
2
Current address: Advanced Steel Processing and Products Research Center, Colorado School of Mines, Golden, CO 80401, USA.
3
Current address: R&D NLMK Group, Moscow, 119017, Russia.

https://doi.org/10.1016/j.msea.2018.08.027
Received 24 May 2018; Received in revised form 9 August 2018; Accepted 9 August 2018
Available online 10 August 2018
0921-5093/ Published by Elsevier B.V.
L. Cho et al. Materials Science & Engineering A 735 (2018) 448–455

Hot press forming process Fig. 1. (a) Schematic of the HPF pro-
cess. (b) Hydrogen TDA curves of the
Uncoated or Furnace Water-cooled Uncoated or aluminized PHS uncoated and the aluminized 35MnB5
Al-10%Si coated steel quenching dies PHSs in the press-hardened (as-deliv-
TA=850-950 ºC
ered) condition, showing that the un-
coated PHS did not show a character-
istic H desorption rate peak in the
20–300 °C temperature range, while
Ferrite + Pearlite Austenite the aluminized PHS exhibited a large
Lath martensite H-uptake. The TDA heating rate was
(a) 100 °C/hour. (c) Room temperature
tensile properties of the uncoated and
0.015 2500 the aluminized 35MnB5 PHSs, in-
Uncoated PHS Uncoated PHS dicative of a loss of the ductility of the
Hydrogen Desorption Rate

Engineering Stress (MPa)


Aluminized PHS Aluminized PHS aluminized PHS due to hydrogen em-
2000
brittlement.
0.010
(ppm·min-1)

1500

0.005 1000

500
0.000
0
0 50 100 150 200 250 300 0 2 4 6 8 10
Temperature ( C) Engineering Strain (%)

(b) (c)

that the control of the dew point of the gas atmosphere of the annealing austenitizing treatment, the specimens were transferred to a laboratory
furnace, i.e. a reduction of the water vapor content in the furnace at- hydraulic HPF simulator where the specimens were quenched between
mosphere was key to suppress the hydrogen uptake of the aluminized two flat water-cooled dies to room temperature. During the die-
PHS. quenching, the austenite was transformed to lath martensite. The me-
Micro-alloying with vanadium (V), which results in the precipita- chanical properties of the specimens were measured in a Zwick/Roell
tion of hydrogen-trapping V carbides (VCs) or nitrides [7–10], can be Z100 universal tensile testing machine.
used as a method to mitigate the hydrogen embrittlement of aluminized The PHS microstructure was observed by means of field emission-
PHS. The present work was initiated to evaluate the effect of V addi- scanning electron microscopy (SEM), electron backscattering diffrac-
tions to the PHS composition, as it does not require the control of the tion (EBSD), and field emission-transmission electron microscopy
gas atmosphere in the industrial austenitizing furnace. This was con- (TEM). The SEM and EBSD analyses were done in an FEI Quanta 3D
sidered to be a more cost-effective approach to improving the hy- FEG. The specimens used for the EBSD analysis were prepared by
drogen-embrittlement resistance of aluminized 1800 MPa and electro-chemical polishing in a solution of 5% perchloric acid and 95%
2000 MPa PHS grades. The contribution also proposes a mechanism for acetic acid. Prior to the SEM observation, the electro-chemically po-
the uptake of hydrogen during the processing of aluminized PHS. lished specimens were etched by using saturated aqueous picric acid at
85 °C to reveal the prior austenite grain boundaries. Thin TEM foils
were prepared at room temperature by twin-jet electro-polishing a
2. Experimental
3 mm disc in a solution of 5% perchloric acid and 95% acetic acid. TEM
specimens were analyzed in a JEOL JEM-2100F field emission-TEM
Four PHSs were used for the present work. A 30MnB5 PHS grade
operating at 200 keV. The precipitates formed in the steels were ana-
(1800 MPa) and a 35MnB5 PHS grade (2000 MPa) were chosen as re-
lyzed by means of energy-dispersive spectroscopy (EDS) using a nan-
ference V-free PHSs. 0.2 wt% of V was added to the composition of each
ometer-sized electron beam. The precipitates formed in the steel matrix
reference PHS to evaluate the effect of V additions on the resistance of
were also extracted by an electrochemical method using a methanol
the PHS to hydrogen embrittlement. Hereafter, the V-added PHSs are
solution containing 10% acetylacetone and 1% tetramethyl ammonium
called as a 0.2V-added 30MnB5 PHS and a 0.2V-added 35MnB5 PHS. In
chloride. The precipitates-containing residue was filtered with a mesh
all cases, 10–20 ppm B was added to the steel composition to enhance
size of 0.2 µm ϕ. Quantitative analysis of the precipitates was carried
their hardenability. The formation of BN was avoided by additions of
out by inductively coupled plasma atomic emission spectroscopy (ICP-
0.03 wt% Ti, which stabilize N as TiN [1].
AES) analysis.
Industrially cold-rolled, full hard PHSs were continuously annealed
In previous investigations, several hydrogen charging methods have
in N2 + 5 vol% H2 gas atmosphere and hot dip aluminized at 680 °C in
been used, including electrolytic charging [8] and gaseous charging
an Al-10%Si bath. The thickness of the Al-Si coating was approximately
[11]. It should be noted that the conventional hydrogen charging
30 µm. The following procedure was used for the HPF simulation of the
methods were not used in the present case because the purpose of this
aluminized PHSs. ASTM E8 standard size tensile specimens with a
study was to simulate the hydrogen uptake of a PHS, which takes place
gauge length of 50 mm were prepared. The initial ferrite-pearlite mi-
during the HPF process. In the present study, hydrogen was naturally
crostructure of the continuously annealed PHS specimen was first
absorbed in the aluminized steel specimens during the austenitizing
austenitized by heating to an austenitizing temperature of 900 °C and
heat treatment as a result of high temperature atmospheric corrosion of
isothermal holding for 5 min in an electrically-heated furnace. No at-
the Al-Si coating due to the water vapor in the furnace atmosphere.
tempt was made to control the furnace atmosphere, i.e. the furnace
Further details about the hydrogen charging method can be found in a
atmosphere contained water vapor. The furnace atmosphere had a dew
recent study by the present authors [6]. Hydrogen analysis of the
point of approximately 5 °C at room temperature. After the

449
L. Cho et al. Materials Science & Engineering A 735 (2018) 448–455

hydrogen-charged specimens was carried out by means of TDA with a PHSs is shown in Fig. 3. The SEM and EBSD results indicate that the
carrier gas consisting of He and N2 using an Agilent 7890 A gas chro- prior austenite grain size, packet size, and block size of the lath mar-
matograph. The dimensions of the aluminized PHS specimen for the tensite were reduced with the V additions. VC is known to inhibit the
TDA were 10 mm x 100 mm x 1.0 mm. The hydrogen desorption rate of growth of the prior austenite grain [14]. This results in the refinement
the specimens was measured during heating to 300 °C using a heating of the martensite substructure.
rate of 100 °C/hour.
3.2. Hydrogen thermal desorption analysis

3. Results and discussion Fig. 4 shows the hydrogen TDA curves of the aluminized V-free
35MnB5 PHS and aluminized 0.2V-added 35MnB5 PHS, both austeni-
3.1. Microstructure tized at 900 °C, die-quenched, and aged at room temperature. A hy-
drogen desorption peak was clearly observed at about 140 °C in the
The hot dip Al-Si alloy coating is affected by the austenitization and TDA curves for both PHSs immediately after quenching. The value of
the die-quenching. The evolution of the microstructural features of the the activation energy obtained from this desorption peak was de-
Al-Si alloy coating at high temperature has been reported earlier [12]. termined to be 22.1 kJ/mol in the previous study [6]. Table 2 lists the
During the heating to the austenitizing temperature, the Al-Si coating literature data for the hydrogen trapping energies associated with dif-
close to the surface is liquefied while the Al-Si coating on the side of the ferent trap sites in ferritic steels and martensitic steels [8,15–23]. The
steel/coating interface reacts with Fe to form a phase mixture consisting activation energy values of hydrogen-desorption for grain boundaries,
of Fe-Al-Si intermetallic phases and the Fe2Al5 phase. During the aus- dislocations, and micro-voids in steels are reported to be 17.2, 26.8 and
tenitizing, which lasts 5–10 min, the liquid Al-Si phase is fully replaced 35.2 kJ/mol, respectively [15]. Based on these values, the desorption
by a solid reaction layer due to the growth of the Fe-Al-Si intermetallic peak observed at 140 °C is associated with the hydrogen-trapping at
phases and the Fe2Al5 phase. either grain boundaries or dislocations. The intensity of the hydrogen
The precipitation behavior of Ti and V during the processing of the peak decreased and the temperature for the peak maximum shifted to
V-added PHSs was investigated by quantitative analysis using the higher temperature with aging at room temperature. These observa-
electrochemical extraction of the precipitates (Table 1). In both the V- tions indicate that hydrogen was trapped at reversible traps and that it
added PHSs (0.2V-added 30MnB5 and 0.2V-added 35MnB5 PHSs), TiN out-diffused during aging at room temperature. The hydrogen weakly
and TiC precipitates were formed during the hot rolling process. These trapped at martensitic matrix should, in principle, naturally escape
precipitates do not dissolve in the subsequent heat treatment processes. from the specimen in several days [8], but the observed desorption peak
Some of the V precipitated during hot rolling, but most of the V was did not completely disappear even after the room temperature aging for
found to remain in solid solution. It is reasonable to assume that the V 4 weeks (Fig. 4(a)). The slow kinetics of the hydrogen desorption during
containing precipitate was VC, since the N was stabilized by TiN for- the room temperature aging are clearly due to the alloy coating layer
mation. The density of the VC increased during the continuous an- that acts as a hydrogen diffusion barrier, as was discussed in the pre-
nealing. During heating of the continuously annealed PHS, cementite in vious study [6]. It should be noted that the Al-Si coating also influences
the pearlite regions first dissolved and the VC precipitated at this re- the hydrogen thermal desorption behavior. It was shown that the hy-
gion. The VC was also partially redissolved during the austenitization in drogen desorption peak obtained for an aluminized PHS specimen was
the HPF process due to its higher solubility in austenite [13]. 20–30% of observed at slightly higher temperature as compared to the corre-
the added V was precipitated as VC while most of the added Ti was sponding peaks of an uncoated PHS specimen [6]. The possible me-
precipitated in the PHSs in the press-hardened state (Table 1). Fig. 2(a) chanism of hydrogen absorption in the aluminized PHS during the HPF
shows scanning TEM (STEM) micrographs of Ti(C,N) and VC pre- process will be discussed in the Section 3.3.
cipitates formed in the as-cold rolled microstructure of the 0.2V-added For all conditions, the intensity of the hydrogen desorption peak in
35MnB5 PHS. Fig. 2(b) to (d) show the precipitates formed in the lath the TDA curve of the aluminized 0.2V-added PHS was higher as com-
martensitic microstructure of the 0.2V-added 35MnB5 PHS in the as- pared to that of the aluminized V-free PHS (Fig. 4), indicating that the V
quenched condition. In particular, fine VC precipitates were uniformly addition resulted in the absorption of more hydrogen. The activation
distributed in the lath martensitic matrix (Fig. 2(c)). Their diameter was energy values of hydrogen-desorption for VC in steels are reported to be
approximately 30 nm. A coarse Ti(C,N) precipitate of about 100 nm in 29–35 kJ/mol, slightly higher than that for grain boundaries (17.2 kJ/
diameter was also present in the lath martensitic matrix (Fig. 2(d)). mol) or the strain field of dislocations (26.8 kJ/mol), as indicated in
The effect of V additions on the microstructure of the aluminized Table 2. For a TDA heating rating of 100 °C/hour, the temperature for
the peak maximum associated with VC is reported to range from 110° to
Table 1 220°C, which is slightly higher compared to those related to grain
Results of quantitative analysis of the Ti and V precipitation (in wt%) in the V- boundaries or dislocations (approximately 100 °C) (Table 2). Wei et al.
added PHSs after cold rolling (CR), continuous annealing (CA), and hot press [24] showed that a desorption peak associated with VC does not appear
forming (HPF). above 250 °C. Therefore, the observed increase in the intensity of the
Alloy Precipitate Thermo-Mechanical Treatment hydrogen desorption peak (Fig. 4(b)) is associated with the hydrogen
trapping at VC precipitates. In addition, the refinement of martensite
After CR After After CR+CA+HPF
substructure, which creates more boundaries of prior austenite grain,
CR+CA
packet, and block, could have contributed to the increased amount of
0.2V-added Precipitated Ti 0.027 0.030 0.030 the hydrogen trapping sites.
30MnB5 (wt%)
PHS Precipitated V 0.091 0.13 0.052 3.3. Influence of V on hydrogen absorption and embrittlement
(wt%)
0.2V-added Precipitated Ti 0.029 0.029 0.031
35MnB5 (wt%) The V addition clearly suppressed the hydrogen embrittlement
PHS Precipitated V 0.089 0.13 0.055 during the tensile straining of the aluminized PHS at room temperature.
(wt%) The aluminized V-free 30MnB5 PHS fractured at an engineering strain
of 4.1% (Fig. 5(a)). The aluminized V-free 35MnB5 PHS fractured
earlier during tensile straining, i.e. at an engineering strain of 2.3%, as
expected due to its higher strength. This clearly illustrates that PHSs

450
L. Cho et al. Materials Science & Engineering A 735 (2018) 448–455

Fig. 2. (a) STEM micrographs of the ferrite and pearlite


microstructure of the 0.2V-added 35MnB5 PHS in the as-
cold rolled state. Ti(C, N) and VC precipitates present in a
ferrite grain are shown in (a). (b) STEM micrographs of the
fully martensitic microstructure after the HPF of the 0.2V-
added 35MnB5 PHS. (c) STEM micrograph and re-
presentative EDS spectrum of nanometer-size VC particles
in the lath martensite. (d) STEM micrograph and EDS
spectrum of a coarse Ti(C, N) precipitate in the lath mar-
tensite.

V-free 30MnB5 V-free 35MnB5 0.2V-added 30MnB5 0.2V-added 35MnB5

2 µm 2µm 2µ m 2 µm

TD 111
RD
001 101

5µm 5µ m 5 µm 5µ m

(a) (b) (c) (d)


Fig. 3. SEM micrographs (top row) and EBSD inverse pole figure maps (bottom row) of the fully martensitic microstructures after the HPF of the aluminized PHS. (a)
V-free 30MnB5 PHS, (b) V-free 35MnB5 PHS, (c) 0.2V-added 30MnB5 PHS, and (d) 0.2V-added 35MnB5 PHS. The martensitic microstructure is clearly refined in the
V-added PHSs.

with a higher strength are more susceptible to hydrogen embrittlement. embrittlement than the V-free PHSs despite the fact that their strength
The 0.2V-added PHSs exhibited the normal ductile fracture behavior was increased by up to 40 MPa (Fig. 5) as a result of the VC pre-
with a significant improvement of the ductility. The total elongations of cipitation hardening (Fig. 2(c)) and the grain refinement (Fig. 3). In the
the 0.2V-added PHSs were approximately 5.5% (Fig. 5(a)). An addi- case of the aluminized V-free 30MnB5 PHS, the hydrogen embrittle-
tional advantage of the V addition is the precipitate strengthening ef- ment was clearly a time-dependent temporary degradation because the
fect. The 0.2V-added PHSs were more resistant to hydrogen mechanical properties were recovered as the diffusible hydrogen

451
L. Cho et al. Materials Science & Engineering A 735 (2018) 448–455

Fig. 4. Changes in the hydrogen TDA curves with aging at room temperature of (a) the aluminized V-free 35MnB5 PHS and (b) the aluminized 0.2V-added 35MnB5
PHS austenitized for 5 min at 900 °C followed by water-cooled die-quenching. The TDA heating rate was 100 °C/hour.

Table 2
Literature data overview for the hydrogen trapping activation energies associated with different trapping sites in ferritic and martensitic steels.

Type of trap Peak temperature, °C (heating rate, °C/hour) Activation energy, kJ/mol Microstructure Reference

Reversible H traps
Grain boundary 105-162 (184-491) 17.2 Pure iron [15]
Grain boundary and dislocation strain field ~100 (100) 21.9 Ti-added ferritic steel [16]
Dislocation strain field 198-311 (83-506) 26.8 Pure iron [15]
Dislocation strain field ~100 (100) 27 Martensitic steel [17]
Microvoids 267-379 (83-488) 35.2 Pure iron [15]
VC (coherent) ~220 (100) 30 V-added martensitic steel [18]
VC (coherent) 110-150 (100) 33-35 V-added tempered martensitic steel [8]
VC (coherent) 190-210 (100) 29.2 V-added martensitic steel [19]
VC (coherent) 219 (100) - V-added high carbon martensitic steel [20]
TiC (semi-coherent) 220 - Ti-added ferritic steel [21]
Irreversible H traps
Dislocation core ~300 (100) 45 Martensitic steel [17]
Ferrite/Fe3C interface ~370 (200) 65.0 Medium carbon ferritic steel [22]
TiC (incoherent) 718-787 (150-360) 86.9 Ti-added ferritic steel [23]

partially out-diffused from the steel during the room temperature aging and the aluminized V-added PHS is described in Fig. 8. During the HPF
(Fig. 5(b)). The plasticity of the aluminized V-free 30MnB5 PHS clearly process of aluminized V-free PHS (Fig. 8(a)), hydrogen is formed by the
improved after the aging treatment, i.e. the aluminized V-free 30MnB5 reduction of the water vapor in the furnace atmosphere by the Al of the
PHS exhibited a normal tensile behavior. On the other hand, the alu- alloy coating layer according to the following reaction: 2Al + 6H2O →
minized V-free 35MnB5 PHS aged for one week still ruptured in the 2Al(OH)3 + 3H2 or 2Al + 3H2O → Al2O3 + 3H2 [25]. It is known that
early stages of tensile straining, confirming an increase of the suscept- this reaction results in the formation of hydrogen-filled blisters in pure
ibility to hydrogen embrittlement with increasing strength in PHSs. Al metal at elevated temperatures [26]. The increased hydrogen uptake
The V addition also had a significant influence on the failure me- in the aluminized PHS is due, in part, to the fact that the thermo-
chanism of the PHSs, as shown in Figs. 6 and 7. A distinct flat fracture dynamic driving force for the reduction of water vapor by Al is much
region was shown at the central region of the fracture surface of the higher than the water vapor reduction by Fe (Table 3). This may result
aluminized V-free 30MnB5 PHS in the as-quenched state (Fig. 6(a)). in the faster generation of hydrogen on aluminized PHS as compared to
The central flat fracture region displayed mainly quasi-cleavage frac- uncoated PHS. At elevated temperatures, hydrogen diffuses much faster
ture, a mixed mechanism of ductile fracture and brittle cleavage frac- in liquid Al as compared to γ-Fe [27,28], while both the solubility and
ture (Fig. 6(b)). Intergranular cracks were also observed in the flat the diffusivity of hydrogen in solid Al are extremely low at room tem-
fracture region. The region outside of the flat fracture exhibited entirely perature [27,29]. This suggests that hydrogen can be rapidly absorbed
ductile microvoid coalescence, or dimpled, fracture surface appearance in the liquid coating and transferred to the γ-Fe matrix during the
(Fig. 6(c)). The aluminized V-free 35MnB5 PHS specimen had a larger austenitization. The hydrogen is then prevented from diffusion out of
area fraction of the flat fracture, consisting of quasi-cleavage feature the martensite after the die-quenching to room temperature by the re-
and intergranular cracks, compared to the aluminized V-free 30MnB5 acted coating layer.
PHS (Fig. 6(a) and (d)), which is also associated with the influence of Comparatively, aluminized V-added PHS has a larger density of
the material's strength. The observation of quasi-cleavage feature and hydrogen traps and retains slightly more hydrogen (Fig. 8(b)). A pre-
intergranular cracks clearly indicates that the sudden fracture without vious study showed that when V is added in martensitic steels, the in-
any post uniform elongation observed for the aluminized V-free PHSs tensity of a hydrogen desorption peak in the temperature range of
(Fig. 5(a)) was due to the hydrogen-induced brittle fracture. In contrast, 70–150 °C is increased due to the hydrogen trapping at VC precipitates
the fractures of the aluminized 0.2V-added PHSs were ductile, as can be [8]. The refinement of martensite substructure also increases the
inferred from the dimpled fracture surface appearances shown in Fig. 7. amount of the hydrogen reversibly trapped to grain, packet and block
No evidence of quasi-cleavage feature or intergranular cracks was ob- boundaries.
served for the aluminized 0.2V-added PHSs. The V micro-alloying addition considerably improved the hydrogen-
A mechanism for the hydrogen uptake in the aluminized V-free PHS embrittlement resistance of the aluminized PHSs despite the fact that it

452
L. Cho et al. Materials Science & Engineering A 735 (2018) 448–455

As-quenched trapped in the dislocation strain field or the grain boundaries. In the
2000 0.2V-added 30MnB5 meanwhile, there have been questions on the remarkable hydrogen
Engineering Stress (MPa)

0.2V-added 35MnB5 trapping capacity of VC despite the exceptionally low trapping energy
1600 [30], which is only slightly higher compared to other weak traps such
V-free 30MnB5
1200 V-free 35MnB5 as dislocations or grain boundaries. In a recent study of Turk et al. [30],
the authors attributed this high hydrogen trapping capacity to the effect
800 of the dense dislocation structures surrounding VC particles. They based
400 their idea on the finding that the hydrogen trapping capacity of VC was
more pronounced in martensitic steels, which contain a high density of
0 dislocations, than in low carbon ferritic steels. It should however be
0 1 2 3 4 5 6 7 1 2 3 4 5 6 7
Engineering Strain (%) pointed out that no physical evidence of the synergistic effect of VCs
(a) and dislocations has been provided.
After 1 week Another important metallurgical effect of the V micro-alloying ad-
dition is the refinement of martensite substructure. This refinement
2000 0.2V-added 30MnB5
Engineering Stress (MPa)

creates many more block and packet boundaries, leading to the de-
0.2V-added 35MnB5
1600 V-free 30MnB5
pletion of the hydrogen content per unit boundary area [7,31]. As a
V-free 35MnB5 result, V micro-alloying effectively increases the critical diffusible hy-
1200 drogen concentration below which the hydrogen embrittlement does
800 not occur. These results are in agreement with a previous report by
Krutikova et al. [10] that V additions resulted in improvement of the
400 hydrogen resistance of high strength quenched and tempered bolt
0
steels. Krutikova et al. also attributed the improved hydrogen resistance
0 1 2 3 4 5 6 7 1 2 3 4 5 6 7 of the material to both the hydrogen trapping effect of V carbonitride
Engineering Strain (%) precipitates and the grain refinement achieved by the micro-alloying
(b) with V. It has similarly been reported that Nb additions effectively
Fig. 5. Influence of the V addition on the hydrogen embrittlement of the alu-
suppress the hydrogen-induced delayed fracture of a press hardened
minized PHSs. (a) Room temperature tensile properties of the aluminized V-free 22MnB5 steel because of the benefits associated with the micro-alloying
PHSs and the aluminized 0.2V-added PHSs in the as-quenched state, i.e. im- precipitates and the resultant grain refinement [7,32].
mediately after the HPF process. (b) Room temperature tensile properties of the
aluminized V-free PHSs and the aluminized 0.2V-added PHSs after aging at
4. Conclusions
room temperature for one week.

Diffusible hydrogen is introduced in aluminized PHS during the


also resulted in a slightly increased hydrogen uptake. This observation austenitization stage of the HPF process. This is due to the reduction of
suggests that the susceptibility to hydrogen embrittlement depends on water vapor in furnace atmosphere by Al. The hydrogen is trapped in
the distribution of the solute hydrogen in the microstructure rather than the lath martensite and inhibited from escaping from the martensite
the overall concentration of hydrogen. The key role of VC on the hy- matrix by the presence of the reacted alloy coating layer at the surface.
drogen-embrittlement behavior of steel is the enhancement of the hy- This causes an early hydrogen-induced fracture of the aluminized PHS
drogen trapping capacity. The hydrogen trapping capacity of VC is at room temperature. PHS with a higher strength is more susceptible to
higher as compared to the dislocation strain field or the grain bound- hydrogen-induce brittle fracture. While the V additions result in the
aries [3]. De-trapping of hydrogen is also slower in V-added steel than absorption of slightly more hydrogen due to an increased density of
V-free steel [3,8]. The hydrogen trapped to VC is therefore considered hydrogen trapping sites and interface boundaries in the finer lath
to be less harmful to its mechanical properties than the hydrogen martensite substructures, it considerably reduces the negative impact of

V-free 30MnB5 V-free 35MnB5

Microvoid coalescence fracture Microvoid coalescence fracture

Quasi-cleavage fracture

0.5mm Quasi-cleavage fracture 0.5mm

(a) (d)

5µm 5µ m 5 µm 5 µm

(b) (c) (e) (f)


Fig. 6. Fracture surfaces of the tensile specimens of the aluminized V-free PHSs in the as-quenched state. (a) Overview SEM micrograph of the fracture surface of
aluminized V-free 30MnB5 PHS. (b) Enlargement of the quasi-cleavage and intergranular fracture zone surrounded by the dashed line in (a). (c) Enlargement of the
microvoid coalescence fracture zone in (a). (d) Overview SEM micrograph of the fracture surface of aluminized V-free 35MnB5 PHS. (e) Enlargement of the quasi-
cleavage and intergranular fracture zone surrounded by the dashed line in (d). (f) Enlargement of the microvoid coalescence fracture zone in (d).

453
L. Cho et al. Materials Science & Engineering A 735 (2018) 448–455

0.2V-added 30MnB5 0.2V-added 35MnB5

Microvoid coalescence Microvoid coalescence


fracture fracture
0.5mm 0.5mm

(a) (c)

2 µm 2 µm

(b) (d)
Fig. 7. Fracture surface of the tensile specimens of the aluminized 0.2V-added PHSs in the as-quenched state. (a) Overview SEM micrograph of the fracture surface of
the aluminized 0.2V-added 30MnB5 PHS. (b) Enlargement of the half thickness region of the fractured specimen of the aluminized 0.2V-added 30MnB5 PHS. (c)
Overview SEM micrograph of the fracture surface of the aluminized 0.2V-added 35MnB5 PHS. (d) Enlargement of the half thickness region of the fractured specimen
of the aluminized 0.2V-added 35MnB5 PHS.

Fig. 8. (a) Schematic illustrating the me-


chanism of hydrogen absorption in the alumi-
nized PHS during the HPF process. The hy-
drogen uptake is accelerated due to the Al
reduction of the water vapor in the furnace gas
atmosphere during the austenitization. The
diffusible hydrogen is prevented from escaping
from the martensitic matrix by the reacted
coating layer at room temperature. (b)
Schematic illustrating the influence of the V
addition on the hydrogen absorption in the
aluminized PHS. The VC precipitates inhibit
grain growth during austenitization and pro-
vide an effective trap site for hydrogen. This
leads to the absorption of slightly more hy-
drogen in the martensitic matrix.

Table 3 the hydrogen uptake on the mechanical properties of aluminized PHS.


Possible reduction reaction of a water vapor molecule (H2O) in the furnace
atmosphere during heating and austenitizing in the HPF process. FactSage Acknowledgement
(version 6.4) was used for the calculations [6].
Possible H2O reduction reactions T (°C) ΔG (kJ) The authors gratefully acknowledge the support of the POSCO
Technical Research Laboratories, Gwangyang, South Korea.
2 Al (l) + 3H2O (g) → Al2O3 (s) + 3H2 (g) 900 − 755.070
Fe (s) + H2O (g) → FeO (s) + H2 (g) 900 − 5.684
2 Fe (s) + 3H2O (g) → Fe2O3 (s) + 3H2 (g) 900 − 8.068
References
3 Fe (s) + 4H2O (g) → Fe3O4 (s) + 4H2 (g) 900 − 6.905
[1] H. Karbasian, A. Tekkaya, A review on hot stamping, J. Mater. Process. Technol.
210 (2010) 2103–2118.
[2] D.W. Fan, H.S. Kim, B.C. De Cooman, A review of the physical metallurgy related to
the hot press forming of advanced high strength steel, Steel Res. Int. 80 (2009)
241–248.

454
L. Cho et al. Materials Science & Engineering A 735 (2018) 448–455

[3] H.K.D.H. Bhadeshia, Prevention of hydrogen embrittlement in steels, ISIJ Int. 56 [18] S. Yamasaki, H. Bhadeshia, M4C3 precipitation in Fe–C–Mo–V steels and relation-
(2016) 24–36. ship to hydrogen trapping, Proceedings of the Royal Society of London A:
[4] C. Georges, T. Sturel, P. Drillet, J.-M. Mataigne, Absorption/desorption of diffusible Mathematical, Physical and Engineering Sciences, The Royal Society, 2006, pp.
hydrogen in aluminized boron steel, ISIJ Int. 53 (2013) 1295–1304. 2315-2330.
[5] S. Salmi, M. Rhode, S. Jüttner, M. Zinke, Hydrogen determination in 22MnB5 steel [19] T. Tsuchida, T. Hara, K. Tsuzaki, Relationship between microstructure and hy-
grade by use of carrier gas hot extraction technique, Weld. World 59 (2015) drogen absorption behavior in a V-bearing high strength steel, Tetsu-to-Hagané 88
137–144. (2002) 771–778.
[6] L. Cho, D.H. Sulistiyo, E.J. Seo, K.R. Jo, S.W. Kim, J.K. Oh, Y.R. Cho, B.C. De [20] B. Szost, R. Vegter, P. Rivera-Díaz-del-Castillo, Developing bearing steels combining
Cooman, Hydrogen absorption and embrittlement of ultra-high strength aluminized hydrogen resistance and improved hardness, Mater. Des. 43 (2013) 499–506.
press hardening steel, Mater. Sci. Eng.: A 734 (2018) 416–426. [21] J. Takahashi, K. Kawakami, Y. Kobayashi, T. Tarui, The first direct observation of
[7] J. Bian, H. Mohrbacher, S. Zhan, H. Lu, W. Wang, Y. Zhang, L. Wang, Proceedings of hydrogen trapping sites in TiC precipitation-hardening steel through atom probe
the 5th International Conference on Hot Sheet Metal Forming of High-Performance tomography, Scr. Mater. 63 (2010) 261–264.
Steel (CHS2), Canada, 2015, pp. 65-74. [22] K. Takai, R. Watanuki, Hydrogen in trapping states innocuous to environmental
[8] H. Asahi, D. Hirakami, S. Yamasaki, Hydrogen trapping behavior in vanadium- degradation of high-strength steels, ISIJ Int. 43 (2003) 520–526.
added steel, ISIJ int. 43 (2003) 527–533. [23] H. Lee, J.-Y. Lee, Hydrogen trapping by TiC particles in iron, Perspect. Hydrog.
[9] C. Zhang, Y. Liu, C. Jiang, J.-f Xiao, Effects of niobium and vanadium on hydrogen- Metals (1986) 421–426 Elsevier.
induced delayed fracture in high strength spring steel, J. Iron Steel Res. Int. 18 [24] F.-G. Wei, T. Hara, K. Tsuzaki, Nano-preciptates design with hydrogen trapping
(2011) 49–53. character in high strength steel, Adv. Steels (2011) 87–92 Springer.
[10] I. Krutikova, L. Panfilova, L. Smirnov, Study of the tendency towards delayed failure [25] H. Wang, D. Leung, M. Leung, M. Ni, A review on hydrogen production using
of high-strength bolt steels microalloyed with vanadium and nitrogen, Metallurgist aluminum and aluminum alloys, Renew. Sustain. Energy Rev. 13 (2009) 845–853.
54 (2010) 48–56. [26] P. Blackburn, E. Gulbransen, Aluminum reactions with water vapor, dry oxygen,
[11] R.L. Amaro, N. Rustagi, K.O. Findley, E.S. Drexler, A.J. Slifka, Modeling the fatigue moist oxygen, and moist hydrogen between 500 and 625 C, J. Electrochem. Soc.
crack growth of X100 pipeline steel in gaseous hydrogen, Int. J. Fatigue 59 (2014) 107 (1960) 944–950.
262–271. [27] N. Jakse, A. Pasturel, Hydrogen diffusion in liquid aluminum from ab initio mole-
[12] M. Windmann, A. Röttger, W. Theisen, Phase formation at the interface between a cular dynamics, Phys. Rev. B 89 (2014) 174302.
boron alloyed steel substrate and an Al-rich coating, Surf. Coat. Technol. 226 [28] V. Olden, C. Thaulow, R. Johnsen, Modelling of hydrogen diffusion and hydrogen
(2013) 130–139. induced cracking in supermartensitic and duplex stainless steels, Mater. Des. 29
[13] S. Zajac, T. Siwecki, W. Hutchinson, R. Lagneborg, Strengthening mechanisms in (2008) 1934–1948.
vanadium microalloyed steels intended for long products, ISIJ int. 38 (1998) [29] C. Qiu, G.B. Olson, S.M. Opalka, D.L. Anton, Thermodynamic evaluation of the Al-H
1130–1139. system, J. Ph. Equilibria Diffus. 25 (2004) 520–527.
[14] G. Yang, X. Sun, Z. Li, X. Li, Q. Yong, Effects of vanadium on the microstructure and [30] A. Turk, D. San Martín, P.E. Rivera-Díaz-del-Castillo, E.I. Galindo-Nava, Correlation
mechanical properties of a high strength low alloy martensite steel, Mater. Des. 50 between vanadium carbide size and hydrogen trapping in ferritic steel, Scr. Mater.
(2013) 102–107. 152 (2018) 112–116.
[15] W. Choo, J.Y. Lee, Thermal analysis of trapped hydrogen in pure iron, Metall. [31] M. Matsumoto, Y. Takemoto, T. Senuma, Proceedings of the 5th International
Mater. Trans. A 13 (1982) 135–140. Conference on Hot Sheet Metal Forming of High-Performance Steel (CHS2), Canada,
[16] F. Wei, T. Hara, K. Tsuzaki, Precise determination of the activation energy for 2015, pp. 55-63.
desorption of hydrogen in two Ti-added steels by a single thermal-desorption [32] S. Zhang, Y. Huang, B. Sun, Q. Liao, H. Lu, B. Jian, H. Mohrbacher, W. Zhang,
spectrum, Metall. Mater. Trans. B 35 (2004) 587–597. A. Guo, Y. Zhang, Effect of Nb on hydrogen-induced delayed fracture in high
[17] M. Enomoto, D. Hirakami, T. Tarui, Thermal desorption analysis of hydrogen in strength hot stamping steels, Mater. Sci. Eng.: A 626 (2015) 136–143.
high strength martensitic steels, Metall. Mater. Trans. A 43 (2012) 572–581.

455

You might also like