You are on page 1of 11

Materials Science & Engineering A 734 (2018) 416–426

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Hydrogen absorption and embrittlement of ultra-high strength aluminized T


press hardening steel
Lawrence Choa,1, Dimas H. Sulistiyoa, Eun Jung Seoa,2, 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: The use of press hardening steel (PHS) in structural safety-related parts has experienced a rapid growth in the
Hydrogen embrittlement automotive industry, due to increased passenger safety standards, which require an improved vehicle intrusion
Press hardening steel resistance. Very small concentrations of diffusible hydrogen in ultra-high strength PHS deteriorate their me-
Hot press forming chanical properties. The focus of the present study was the analysis of the diffusible hydrogen uptake during the
Aluminizing
hot press forming process of the aluminized PHS and its impact on the mechanical properties of PHS. The effect
of the paint baking on the properties of aluminized PHS was also evaluated by mechanical testing and hydrogen
thermal desorption analysis. A mechanism for the hydrogen absorption of aluminized PHS was proposed.

1. Introduction (3) The austenitized steel is subsequently quenched between water-


cooled dies. During the die-quenching stage, the blank is deformed
Hot press forming (HPF) also known as press hardening, hot into the required shape and the austenitic microstructure trans-
stamping, and die quenching, is a process to manufacture ultra-high forms fully into a lath martensitic microstructure.
strength components, i.e. with a tensile strength (TS) > 1 GPa, with a
fully martensitic microstructure. The process also enables the produc- Metallic coating systems are required to protect PHS from corro-
tion of parts with a complex geometry without shape deviations re- sions. In particular, the HPF process leads to significant oxide formation
sulting from spring back. Hardenable boron steels, such as 22MnB5, are on the surface of parts made from uncoated steel during the austeniti-
generally used for the HPF process. Press hardened parts produced by zation stage of the HPF process. This requires the removal of the oxide
the HPF process are currently used in the automotive industry for the scale on the stamped parts by shot blasting. Pre-coated steel blanks are
production of A-pillars, B-pillars, bumpers, roof rails, rockers and tun- therefore preferred for HPF as these do not require an additional scale
nels [1]. The number of press hardening steel (PHS) parts used in au- removal stage. In addition, the protective alloy coating layer on the PHS
tomotive applications has increased sharply, and the percentage of PHS prevents the decarburization of the surface. The coating also improves
parts in state-of-the-art car bodies, which currently stands at about the in-service corrosion resistance. The most common coating for PHS is
26%, could reach 40% in the future [1,2]. the Al-10%Si alloy coating. The Al-10%Si alloy coating is applied to
The HPF process consists of the following stages: cold-rolled steel by hot dip aluminizing. The typical composition of the
aluminizing bath is 87% aluminum, 10% silicon, and 3% iron [3].
(1) A cold-rolled and annealed steel blank is heated to a soaking or The increased use of aluminized, i.e. Al-Si coated, PHS has led to
austenitizing temperature between 850 and 950 °C. concerns about hydrogen (H) embrittlement. The risk of H embrittle-
(2) The blank is held at the austenitizing temperature for 5–10 min. ment is limiting the industrial application of ultra-high strength alu-
During the heating process, the initial ferrite-pearlite micro- minized PHS with TS > 1.5 GPa.
structure transforms to austenite. Martensite, the main constituent of the matrix microstructure of

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.2018.08.003
Received 20 April 2018; Received in revised form 13 July 2018; Accepted 2 August 2018
Available online 07 August 2018
0921-5093/ Published by Elsevier B.V.
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

PHS, has a pronounced susceptibility to H-related failure, as compared iii) BH stage: During the bake hardening (BH) simulation, a press-
to ferritic, bainitic, austenitic, and precipitation-hardened steel [4]. hardened specimen was isothermally held at 170 °C for 20 min. This
Previous studies have shown that the H-uptake occurs during the pro- low-temperature tempering treatment simulates the paint baking of
duction of PHS [5]. This H-uptake causes a severe, albeit transient, an automotive part.
reduction of the plasticity of PHS [5].
Aluminized PHS is considerably more sensitive to H-uptake and the The uncoated specimens obtained after the CA, HPF, and BH stages
H-induced loss of plasticity than uncoated PHS [5]. The H-embrittle- are labelled CA, CA+HPF, and CA+HPF+BH, respectively.
ment sensitivity of aluminized PHS has not been studied in detail. In The schematic shown in Fig. 1(b) illustrates the sequence of pro-
particular, the influence of the Al-Si coating on the H-uptake of the PHS cessing steps for the aluminized PHS. The specimen was continuously
during the HPF process has not yet been studied systematically. annealed at 800 °C for 1 min in an N2 + 15% H2 gas atmosphere. The
The present contribution aims to address the following questions: specimen was then aluminized by hot dipping in an Al-10% Si liquid
alloy at 680 °C prior to cooling to room temperature in a hot dip alu-
(1) Which process in the industrial processing of PHS is responsible for minizing simulator. The aluminizing stage is referred to as “Al”. The
the H-uptake of the PHS? thickness of the Al-Si coating was approximately 30 µm. The thermal
(2) How does the aluminized coating accelerate the H-uptake of the cycles for the HPF and BH stages of the aluminized PHS used the same
PHS? thermal cycles as those used for the uncoated PHS (Fig. 1(a)). The
(3) What is the impact of the H-uptake of the aluminized PHS on their aluminized specimens obtained after CA and Al, HPF, and BH are la-
mechanical properties? belled CA+Al, CA+Al+HPF, and CA+Al+HPF+BH, respectively.
The panels of the PHS after each stage (CA, CA+HPF, CA+HPF
The present study focused on the analysis of the H-uptake during the +BH, CA+Al, CA+Al+HPF, and CA+Al+HPF+BH) were provided
industrial processing of an ultra-high strength aluminized PHS and its by POSCO, Gwangyang. The dimensions of the industrially processed
effect on the mechanical properties of this PHS. The H-uptake was PHS panels were 500 mm in width, 550 mm in length, and 1.5 mm in
quantified by means of H thermal desorption analysis (H-TDA). The H- thickness. It is important to note that, after the HPF, it took three days
uptake of uncoated PHS during the industrial processing was also in- to prepare the PHS panels for H-TDA and the mechanical testing, i.e. the
vestigated, for comparison purpose. Finally, a mechanism for the H- H-TDA and the mechanical testing were, in all cases, carried out 72 h
uptake of the aluminized PHS was proposed. after each processing step.
In order to evaluate the H trapping states in both the uncoated PHS
2. Experimental procedure and the aluminized PHS, a second set of H-TDA tests were carried out,
using a heating rate in the range of 4–30 °C/min. Since H has an ex-
An industrially produced, press-hardenable 35MnB5 steel was used tremely low solubility in α-Fe, weakly trapped H or diffusible H will
in the present study. The final cold-rolled thickness was 1.5 mm. The naturally escape from steel at room temperature. For these experiments,
steel in the press-hardened state has the ultimate tensile strength (UTS) a laboratory-scale HPF simulation was carried out. Prior to the la-
of approximately 2 GPa. The schematic shown in Fig. 1(a) illustrates the boratory-scale HPF simulation, the small coupons of 20 mm in width,
sequence of processing stages for the uncoated PHS: 100 mm in length, and 1.5 mm in thickness were taken from the con-
tinuously annealed PHS panels and the aluminized PHS panels. The
i) CA stage: A continuous annealing (CA) or recrystallization an- PHS specimens were heated to 900 °C and isothermally held for 5 min in
nealing simulation was carried out at 800 °C for 1 min in an N2 an electrically-heated furnace. The specimens were then transferred to a
+ 15% H2 gas atmosphere. laboratory hydraulic HPF simulator where the specimens were quen-
ii) HPF stage: During the HPF process simulation, a continuously an- ched between two flat water-cooled dies to room temperature. The die
nealed specimen was austenitized at 900 °C for 5 min and quenched temperature was maintained at room temperature by cooling the water
between the water-cooled dies to room temperature. Note that no flowing through channels machined in the die. The thermal cycle used
attempt was made to control the furnace atmosphere, i.e. the fur- for the laboratory-scale HPF simulation was same as the one used for
nace atmosphere contained water vapor. the HPF of the larger industrial PHS panels (Fig. 1). No attempt was
made to control the furnace atmosphere in the laboratory-scale HPF
simulations.
H analysis was carried out using a combination of two different H-
TDA instruments. The first H-TDA set up consisted of an infrared fur-
nace attached to an Agilent 7890A Gas Chromatograph. The dimensions
of the specimens for the H-TDA were 1.5 mm in thickness, 20 mm in
width, and 100 mm in length. The H desorption rate of the specimens
was measured during heating to 300 °C using a heating rate of 1.67 °C/
min. The specimens were analyzed at three-minute intervals using He as
carrier gas. The second H-TDA set up was the BRUKER G4 Phoenix
diffusible H analyzer. The H desorption rate of the specimens was
measured during heating to 800 °C using a constant heating rate. The
heating rate could be varied in the range of 4–30 °C/min. The carrier
gas was pure N2.
Uniaxial tensile tests and three-point bending tests were carried out
to observe the effect of H on the mechanical properties of the uncoated
and the aluminized PHS. The tensile tests were done with a Zwick/Roell
2100 universal tensile testing machine using a strain rate of 5 × 10−3
s−1. The ASTM E8 standard sized tensile specimens with a gauge length
of 50 mm used for this purpose were machined with their length par-
Fig. 1. Thermal cycles used for (a) uncoated and (b) aluminized PHS (CA: allel to the rolling direction. For each processing condition, three tensile
continuous annealing, HPF: hot press forming, BH: bake hardening, and Al: tests were performed. The three-point bending test equipment had a
aluminizing). punch radius of 0.4 mm. The punch speed was 20 mm/min. The

417
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

Fig. 2. SEM images of the microstructure of the PHS specimens obtained after (a) continuous annealing, (b) hot press forming, and (c) bake hardening.

dimensions of the specimen used for the bending tests were 20 mm in process. The martensite structure of the bulk PHS matrix is character-
width, 80 mm in length, and 1.5 mm in thickness. One bending test per ized by its characteristic lath martensite microstructure (Fig. 3(b) and
each processing condition was performed. (c)).
The microstructure of the specimens was investigated by means of
field emission (FE) scanning electron microscopy (SEM), electron
3.2. Binding state of H absorbed in the PHS
backscattering diffraction (EBSD), and FE electron probe microanalysis
(EPMA). The SEM analysis was carried out in a ZEISS ULTRA-55 FE-
Fig. 4 shows results of H-TDA tests for the uncoated PHS and the
SEM. The EBSD analysis was conducted in an FEI Quanta 3D FEG FE-
aluminized PHS after each processing step. Note that, for the results
SEM. The EPMA analysis was conducted in a JEOL JXA-8530F FE-
shown in Fig. 4, the H-TDA specimens were taken from the industrially
EPMA operated at 15 keV using a 50 nA beam current.
processed PHS panels and that the H-TDA test was performed three
days after each process. The uncoated specimens did not show a char-
3. Results and discussion acteristic H desorption rate peak in the 20–300 °C temperature range
(Fig. 4(a)). The total amounts of diffusible H measured for the uncoated
3.1. Microstructure specimens after the CA, CA+HPF, and CA+HPF+BH stages were
0.036 ppm, 0.019 ppm and 0.020 ppm, respectively (Fig. 4(a)). The
Fig. 2 shows the PHS microstructure after CA, HPF, and BH. After TDA data for the aluminized specimen after aluminizing (CA+Al) did
CA, the specimen microstructure consisted of a ferrite matrix and also not show a H desorption rate peak in the 20–300 °C temperature
pearlite islands, formed by alternating lamellae of ferrite and cementite range (Fig. 4(b)). However, a H desorption rate peak was observed at
(Fig. 2(a)). The specimen after the HPF (CA+HPF) had a lath mar- approximately 150 °C in the specimen after the HPF stage (CA+Al
tensite microstructure. This lath martensite microstructure contained +HPF) and after the BH stage (CA+Al+HPF+BH). The total contents
fine carbide precipitates, which likely formed as a result of auto-tem- of diffusible H for the aluminized specimens after the CA+Al, CA+Al
pering during the die-quenching (Fig. 2(b)). The micrographs in +HPF, and CA+Al+HPF+BH stages were 0.055 ppm, 0.502 ppm and
Fig. 2(c) show the tempered martensitic microstructure of the specimen 0.369 ppm, respectively. This result implies that the aluminized PHS
after the BH process (CA+HPF+BH). Tempering resulted in the for- exhibited a large H-uptake during the HPF. After the BH, the H content
mation of relatively coarse carbide particles as can be seen in the inset in the specimen decreased, which suggests that the BH treatment acted
of Fig. 2(c). as a low temperature H out diffusion treatment.
The hot dip Al-Si alloy coating is affected by the austenitization and In order to evaluate the H binding states in both the uncoated PHS
the die-quenching stages in the HPF process. During heating to the and the aluminized PHS, a second set of H-TDA tests were carried out
austenitizing temperature, the Al-Si coating close to the surface is li- using different heating rates. Fig. 5 shows the H-TDA curves of the
quefied while the Al-Si coating on the side of the steel/coating interface uncoated PHS and the aluminized PHS, both austenitized at 900 °C for
reacts with Fe to form a phase mixture consisting of Fe-Al-Si inter- 5 min and die-quenched to room temperature. In this experimental set,
metallic phases and the Fe2Al5 phase. During the austenitizing at the H-TDA could be done immediately after the laboratory-scale HPF
900 °C, the liquid Al-Si phase is fully replaced by a solid reaction layer simulation. A H desorption rate peak was clearly observed in the range
due to the growth of the Fe-Al-Si intermetallic phases and the Fe2Al5 of 50–350 °C in the H-TDA curves of the uncoated PHS immediately
phase. after the die-quenching (Fig. 5(a)). This low-temperature peak is la-
Cross-sectional analysis of aluminized PHS after the HPF process belled “Peak 1” in Fig. 5(a). The intensity of the H peak increased and
(CA+Al+HPF) was carried out by means of SEM, EBSD, and FE-EMPA. the temperature of the peak maximum shifted to higher temperatures
Fig. 3 shows the surface and substructure of the aluminized PHS spe- with increasing heating rate. A second H desorption peak was observed
cimen in cross-section. Three regions are visible: a reacted surface in the range of 350–650 °C in the TDA curves of the uncoated PHS
coating layer, a diffusion layer, and the lath martensitic matrix. The (Fig. 5(a)). This high-temperature peak is labelled “Peak 2” in Fig. 5(a).
reacted coating layer consisted of an Fe2Al5 and complex Fe-Al-Si in- The intensity of this H peak increased and the temperature for the peak
termetallic phases. Inside the reacted alloy coating area, the regions maximum shifted to higher temperatures with increasing heating rate.
with a black contrast in the EBSD phase map (Fig. 3(c)) correspond to a Enomoto et al. [7] analyzed the TDA curve for high strength martensitic
complex Fe-Si-Al intermetallic phase. Al and Si were richer in the Fe-Si- steel by means of numerical simulation and experiments. They also
Al intermetallic phase than in the Fe2Al5 phase (Fig. 3(d)). This Si-rich observed two peaks in the TDA curve and demonstrated that the first
intermetallic constituent has been identified as Al2Fe3Si3 (τ1) phase in a peak appearing at around 100 °C, which corresponds to Peak 1 in
previous study [6]. The coarse-grained Al-diffused ferritic layer was Fig. 5(a), is due to H-atoms trapped in the stress field surrounding
formed at the steel/coating interface. The diffusion layer was α-Fe(Al), dislocation. They reported that the second peak appearing at about
i.e. a solid solution of Al in BCC α-Fe. This region was not austenitized 300 °C is due to H-atoms trapped in the dislocation cores. This peak
due to its high Al and Si contents (Fig. 3(d)). The austenitized steel corresponds to Peak 2 in Fig. 5(a) of the present study.
matrix was transformed into martensite during the die-quenching The diffusible H content of the uncoated PHS was approximately

418
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

Fig. 3. Cross-section of the aluminized PHS after hot press forming. (a) SEM image. (b) EBSD inverse pole figure map. (c) EBSD phase map. (d) SEM image and the
corresponding FE-EPMA elemental maps for Fe, Al, and Si.

Fig. 4. (a) H-TDA curves for the uncoated PHS


after continuous annealing (CA), hot press
forming (CA+HPF), and bake hardening
(CA+HPF+BH). (b) H-TDA curves for the
aluminized PHS after continuous annealing
and aluminizing (CA+Al), hot press forming
(CA+Al+HPF), and bake hardening (CA+Al
+HPF+BH). TDA was done three days after
each process using an Agilent 7890A Gas
Chromatograph. The TDA heating rate was
1.67 °C/min.

0.10 ppm. The term “diffusible H” refers to H that desorbs from a H- The trap activation energy of H (Ea) was obtained from the heating
charged specimen during aging at the moderately elevated temperature rate dependence of the peak maximum temperature, using the fol-
of approximately 300 °C [8] or even at room temperature [9]. In the lowing equation [10]:
present study, the diffusible H is defined as the H which desorbs at low
∂ ln(ϕ/ Tp2) Ea
temperature. It corresponds to the H-atoms which give rise to the Peak =−
1 desorption peak in the H-TDA curves. ∂ (1/ Tp) R (1)

419
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

Fig. 5. H-TDA curves for (a) the uncoated PHS and (b) the aluminized PHS, both austenitized at 900 °C for 5 min and die-quenched to room temperature. TDA was
done immediately after the HPF simulations using a Bruker G4 Phoenix Diffusible H analyzer. TDA heating rate ranged from 4° to 30 °C/min.

Here, ϕ is the heating rate used during the H-TDA test in K/s, Tp is laboratory-scale HPF simulation (Fig. 5(a)), while it did not appear in
the peak maximum temperature in K, and R is the gas constant. The the H-TDA curve for the uncoated PHS samples after the industrial HPF
values of the activation energy obtained from Peak 1 and Peak 2 are process (CA+HPF) (Fig. 4(a)). This is due to that the fact that, in the
22.1 kJ/mol and 74.7 kJ/mol, respectively. These values indicate that case of the PHS panels produced industrially, the H-TDA tests were
the H-atoms giving rise to the desorption Peak 1 are weakly bound. The performed three days after the HPF process. This implies that most of
H-atoms giving rise to the desorption Peak 2 are more strongly bound. the absorbed H diffused out of the steel at room temperature. This is
The “Peak 1” H-atoms diffuse easily at room temperature. The “Peak 2” consistent with the result of Hirakami et al. [13] that most of the dif-
H-atoms remain bound at the trap site at room temperature. Table 1 fusible H diffused out of a H-charged, 0.35 wt% C martensitic steel at
lists the data of the present study and literature data [7,10–12] for the room temperature in 2–4 days.
trapping energies for H in ferritic steels and martensitic steels. The
activation energy values of H-desorption for grain boundaries, dis- 3.3. Effect of microstructure and H on the mechanical and fracture behavior
locations, and micro-voids in steels are reported to be 17.2, 26.8 and
35.3 kJ/mol [10]. Based on these values the desorption “Peak 1” is The results of the tensile tests, bending tests, and the TDS analysis
associated with the H-trapping at either grain boundaries or the strain show a clear correlation between the deterioration of the plasticity of
field of dislocations. the aluminized PHS and the H content. Table 2 indicates the room
Fig. 5(b) shows the H-TDA curves of aluminized PHS immediately temperature tensile properties of the uncoated PHS and aluminized PHS
after die-quenching. Both the H-desorption rate peak “Peak 1” and the specimens after the different process stages. Each value listed in Table 2
H-desorption rate peak “Peak 2” were observed at slightly higher is the average value measured from three tensile tests. Fig. 6 shows the
temperatures as compared to the corresponding peaks of the uncoated representative engineering stress strain curves. Fig. 6(a) shows the
PHS. Peak 1 was in the 100–400 °C temperature range. Peak 2 was in stress–strain curves of the H-free, uncoated PHS. The uncoated spe-
the 400–650 °C temperature range. The shift of the H desorption peaks cimen after the CA stage had the ferrite-pearlite microstructure. Its
is likely caused by the presence of the Al coating, i.e. diffusion of H in tensile properties exhibited a low yield strength (YS), a low UTS, and a
solid Al is very slow and thus, the Al coating can slow down the des- high total elongation (TE). The average YS, UTS, and TE for the spe-
orption process of H trapped at martensitic matrix during the TDA. The cimen after CA were 410 MPa, 637 MPa, and 23.3%, respectively
diffusible H content of the aluminized PHS was approximately (Table 2). After the HPF simulation (CA+HPF), the steel microstructure
0.30 ppm, i.e. much higher than that of the uncoated PHS (~ 0.10 ppm). was fully martensitic. The uncoated specimen after the CA+HPF stage
The intensity of both Peak 1 and Peak 2 was higher for the aluminized had a high YS (1234 MPa), a high UTS (1986 MPa), and a low TE
PHS (Fig. 5(b)) as compared to the uncoated PHS (Fig. 5(a)), confirming (6.3%). After the BH process (CA+HPF+BH), the YS increased to
that the aluminized PHS is more sensitive to H-uptake than the un- 1439 MPa, the UTS decreased to 1911 MPa and the TE increased
coated PHS. slightly to 7.2%. The increase in the YS by approximately 200 MPa
It is noteworthy that an H desorption peak (Peak 1) was clearly clearly indicates a strong BH response of the PHS. The BH behavior of
visible in the H-TDA curve of the uncoated PHS samples after the lath martensitic steel has been explained by dislocation locking caused

Table 1
Literature data overview for the trapping activation energies for H in ferritic steels and martensitic steels. Ferrite and martensite are labelled α and α′, respectively.
Type of trap Peak temperature, °C (heating rate, °C/min) Activation energy, kJ/mol Microstructure Reference

Reversible H traps
Grain boundary 105–162 (3.06–8.18) 17.2 Pure iron (α) Choo and Lee [24]
Grain boundary and dislocation strain field ~ 100 (1.67) 21.9 Ti-added steel (α) Wei et al. [25]
Dislocation strain field 198–311 (1.39–8.44) 26.8 Pure iron (α) Choo and Lee [24]
Dislocation strain field ~ 100 (1.67) 27 Martensitic steel (α′) Enomoto et al. [21]
Microvoids 267–379 (1.39–8.13) 35.2 Pure iron (α) Choo and Lee [24]
Irreversible H traps
Dislocation core ~ 300 (1.67) 45 Martensitic steel (α′) Enomoto et al. [21]
Ferrite/Fe3C interface ~ 370 (3.33) 65.0 Medium carbon steel (α) Takai and Watanuki [26]
Peak 1: dislocation strain field 110–250 (4–30) 22.1 PHS (α′) Present study
Peak 2: dislocation core 480–500 (4–30) 74.7 PHS (α′) –
Peak 1: dislocation strain field 200–300 (4–30) – Aluminized PHS (α′) –
Peak 2: dislocation core 480–520 (4–30) – Aluminized PHS (α′) –

420
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

Table 2 (a) (b)


CA CA+HPF CA+HPF+BH CA+Al CA+Al+HPF CA+Al+HPF+BH
Summary of tensile properties of the PHS after each process (CA: continuous
annealing, HPF: hot press forming, BH: bake hardening, and Al: aluminizing).
The reported results are averages of three repeated tests for each condition.
Process YS (MPa) UTS (MPa) TE (%)
62°

Uncoated PHS CA 410 ± 8.9 637 ± 8.4 23.3 ± 0.3


CA+HPF 1234 ± 2.0 1986 ± 7.5 6.3 ± 1.3
CA+HPF+BH 1439 ± 12 1911 ± 7.7 7.2 ± 0.6
Aluminized PHS CA+Al 441 ± 0.7 585 ± 3.3 24.1 ± 0.1 62°
CA+Al+HPF 1236 ± 20 1896 ± 46 2.8 ± 0.3
CA+Al+HPF+BH 1447 ± 7.7 1927 ± 17 6.9 ± 0.6

by the formation of Cottrell atmospheres, precipitation of iron carbides,


and softening of martensitic matrix due to the tempering [14].
Fig. 6(b) shows the stress–strain curves of the aluminized PHS. The
fracture behavior during the tensile test was clearly associated with the
amount of the diffusible H in the aluminized PHS. The average YS, UTS, Uncoated PHS Aluminized PHS
and TE for the aluminized specimen after the aluminizing (CA+Al)
Fig. 7. (a) Sample fracture after tensile test of uncoated PHS after continuous
stage were 441 MPa, 585 MPa, and 24.1%, respectively (Table 2). The
annealing (CA), hot press forming (CA+HPF), and bake hardening (CA+HPF
tensile properties of the aluminized specimen after the CA+Al stage
+BH) and (b) aluminized PHS after continuous annealing and aluminizing
were similar to those for the uncoated PHS after the CA stage because it
(CA+Al), hot press forming (CA+Al+HPF), and bake hardening (CA+Al
did not contain diffusible H (Fig. 4(b)). The aluminized specimen after +HPF+BH).
the HPF stage (CA+Al+HPF) had a YS of 1236 MPa and a UTS of
1896 MPa. The specimen fractured abruptly at an elongation of 2.8%.
martensite is associated with its high density of dislocations (~ 1015
Approximately 0.5 mass ppm of diffusible H was present in the alu-
m−2), which serve principal traps for H. Therefore, it is also possible
minized specimen after the CA+Al+HPF stage (Fig. 4(b)). The ducti-
that a slight reduction in the dislocation densities resulting from the
lity of the aluminized PHS was clearly reduced by H embrittlement. The
tempering of the martensitic PHS contributed to the improvement in
aluminized specimen after the BH treatment (CA+Al+HPF+BH) had a
the ductility shown in Fig. 6(b). While the BH treatment clearly resulted
YS of 1447 MPa, and a UTS of 1927 MPa. The TE was 6.9%, i.e. much
in a decrease of the diffusible H content of the aluminized PHS
larger than for the specimen after CA+Al+HPF. This improvement in
(Fig. 4(b)), the TE for the aluminized specimen after the CA+Al+HPF
the ductility is due partly to the fact that the BH process partially re-
+BH remained smaller than for the uncoated specimen after the
leased the diffusible H present in the aluminized PHS (Fig. 4(b)). The H
CA+HPF+BH stage. This implies that the BH process can only partially
susceptibility of lath martensitic steel is known to depend on the degree
mitigate the damage caused by H. It does not fully prevent the H em-
of tempering [15] and the dislocation density [16] as well as the dif-
brittlement.
fusible H content. In general, as-quenched martensitic steel is more
Fig. 7 shows macro images of the fractured tensile specimen. A
sensitive to H-induced brittle fracture than tempered martensitic steel
ductile fracture with diffuse necking occurred in the cases of the
[15]. Lee et al. [15] reported that the high H sensitivity of as-quenched

(a) i) CA ii) CA+HPF iii) CA+HPF+BH


2500
Engineering Stress, MPa

2000

1500

1000

500

0
0 5 10 15 20 0 2 4 6 8 0 2 4 6 8 10
Engineering Strain, % Engineering Strain, % Engineering Strain, %
(b) i) CA+Al ii) CA+Al+HPF iii) CA+Al+HPF+BH
2500
Engineering Stress, MPa

2000

1500

1000

500

0
0 5 10 15 20 0 2 4 6 8 0 2 4 6 8 10
Engineering Strain, % Engineering Strain, % Engineering Strain, %
Fig. 6. (a) Tensile properties of the uncoated PHS after continuous annealing (CA), hot press forming (CA+HPF), and bake hardening (CA+HPF+BH). (b) Tensile
properties of the aluminized PHS after continuous annealing and aluminizing (CA+Al), hot press forming (CA+Al+HPF), and bake hardening (CA+Al+HPF+BH).

421
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

uncoated specimen after the CA (Fig. 7(a)). Both the uncoated speci- BH treatment (CA+Al+HPF+BH). A sudden fracture with a θF = 90°
mens after the HPF stage (CA+HPF) and after the BH treatment due to limited plastic deformation was observed only for specimens
(CA+HPF+BH) had a shear-type fracture with a fracture angle, θF, of containing diffusible H. Therefore, this sudden fracture can be con-
62°. A shear-type fracture is usually attributed to a very high strength, sidered to be related to the H embrittlement.
which results in a decreased ratio of the average critical normal fracture Fig. 8 shows micrographs of the fracture surface of specimens after
stress to shear fracture stress. Zhang and Eckert [17] have shown the tensile tests. It shows the region of the fractured surface up to 40 µm
dependence of the tensile shear fracture angle on the ratio of the below the surface. Fig. 8(a) reveals that for the uncoated specimens, the
average critical normal fracture stress to shear fracture stress. In high- fracture was always ductile, as can be inferred from dimpled fracture
strength materials or nanostructured materials, θF is in the range of surface appearance obtained in the CA, CA+HPF and CA+HPF+BH
45°–90°. In brittle materials with a very high hardness, the maximum specimens. In the case of aluminized specimens (CA+Al), the fracto-
normal fracture stress controls the failure mode, resulting in a fracture graphic observation also revealed a dimpled fracture surface (Fig. 8(b)).
with a θF = 90°. The appearance of the fracture area for the aluminized The aluminized specimen after the HPF stage (CA+Al+HPF) had a
specimen after the aluminizing (CA+Al) was similar to that for the mixed ductile and brittle fracture surface. This fracture surface con-
uncoated specimen after the CA, i.e. they both exhibited diffuse necking tained inter-granular cracks close to the coating. The presence of inter-
characteristics. However, a fracture with a θF = 90° occurred for the granular cracks is usually considered as direct evidence for brittle
aluminized specimens after the HPF stage (CA+Al+HPF) and after the fracture. Closer to the steel matrix, the fracture plane consisted of

Fig. 8. SEM images of the fracture surface of tensile specimens. (a) The uncoated PHS after continuous annealing (CA), hot press forming (CA+HPF), and bake
hardening (CA+HPF+BH). (b) The aluminized PHS after continuous annealing and aluminizing (CA+Al), hot press forming (CA+Al+HPF), and bake hardening
(CA+Al+HPF+BH).

422
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

Fig. 9. (a) Bending test results of the uncoated PHS after continuous annealing (CA), hot press forming (CA+HPF), and bake hardening (CA+HPF+BH). (b) Bending
test results of the aluminized PHS after continuous annealing and aluminizing (CA+Al), hot press forming (CA+Al+HPF), and bake hardening (CA+Al+HPF+BH).

(a) CA CA+HPF CA+HPF+BH each processing step. The specimens were bent to a bending angle of
90° or 140°. The 140° bending was used for specimens after the CA stage
and after the CA+Al stage. The specimens after the HPF stage and after
the BH treatment were bent to a smaller bending angle (90°) since they
were less ductile as compared to the samples with a ferrite+pearlite
microstructure after the CA stage or the CA+Al stage. The bending
(b) CA+Al CA+Al+HPF CA+Al+HPF+BH angle at maximum bending load, θB, for the uncoated specimens after
the CA, CA+HPF, and CA+HPF+BH stages was 97°, 67°, and 73°,
respectively (Fig. 9(a)). The value of θB for the aluminized specimens
after the CA+Al, CA+Al+HPF, and CA+Al+HPF+BH stages was 97°,
28°, and 46°, respectively (Fig. 9(b)). The aluminized specimen after the
20 mm CA+Al stage exhibited a bending performance similar to that of the
uncoated specimen after the CA stage. However, the bending perfor-
Fig. 10. Specimens after bending test. (a) The uncoated PHS after continuous mance of the uncoated and the aluminized specimens after the HPF
annealing (CA), hot press forming (CA+HPF), and bake hardening (CA+HPF stage was clearly different. After the BH treatment, the θB for the alu-
+BH). (b) The aluminized PHS after continuous annealing and aluminizing minized specimen increased to 46°, but it was smaller as compared to
(CA+Al), hot press forming (CA+Al+HPF), and bake hardening (CA+Al the uncoated specimen after the CA+HPF+BH stage (73°). These re-
+HPF+BH).
sults mirror those obtained by the tensile tests in that the BH process
partially mitigated the effect of the H embrittlement.
dimples. River marks, indicative of trans-granular fracture, were pre- The effect of the H was also visible in the appearance of the speci-
sent in the coating area of the fracture surface of the CA+Al+HPF mens after the bending tests. Fig. 10 shows images of the specimens
specimen. The trans-granular fracture feature is evidence that the re- after the bending test. The uncoated specimen after the CA+HPF stage
acted surface alloy coating was brittle. The dimple size for the alumi- had a single surface crack in the bend area. The aluminized specimen
nized specimens after the HPF stage (CA+Al+HPF) and after the BH after the CA+Al+HPF stage fully ruptured in the early stages of
treatment (CA+Al+HPF+BH) was much smaller than that for the bending. The aluminized specimen after the CA+Al+HPF+BH stage
uncoated specimens. had multiple cracks in the bend area.
The origin of the smaller dimple size observed on the fracture sur- It could be argued that the reduction of the ductility is caused by the
face on the aluminized specimens after HPF is very likely associated lower ductility of the reacted surface layer that consists mostly of the
with the presence of diffusible H. Decrease of the dimple size is, in brittle Fe2Al5 intermetallic compound, rather than the uptake of dif-
general, due to a very high void nucleation rate and a very small void fusible H during the austenitization stage of the HPF process. The em-
growth rate during the void coalescence process [18]. Lynch [19] has brittlement phenomena shown in the present study can be attributed
shown that a dimpled fracture can be caused by adsorbed H atoms fa- mainly to H, because the degree of the embrittlement was dependent on
cilitating dislocation formation at crack tips, which promotes the coa- the amount of the diffusible H in the specimen. In addition, the BH
lescence of cracks with voids. treatment, i.e. low-temperature tempering treatment, partially removed
The H-uptake during the HPF stage of the aluminized PHS also af- the diffusible H, and it suppressed the embrittlement of the steel.
fected the bending performance of the steel. Fig. 9 shows the bending Moreover, the degradation of the mechanical properties is shown to be
test results of the uncoated PHS and aluminized PHS specimens after transient phenomena (Supplementary materials).

423
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

Table 3 water generates large amounts of H [20]. Al an Al-alloys have been


Possible reduction reaction of a water vapor molecule (H2O) in the furnace used as H-generation materials [20]. The reaction of pure Al metal with
atmosphere during heating and austenitizing in the HPF process. water at elevated temperatures is known to result in the formation of H-
Possible H2O reduction reactions T (°C) ΔG (kJ) filled blisters [21]. The thermodynamic driving force for the reduction
of water vapor by Al is much higher than the water vapor reduction by
2 Al (l) + 3H2O (g) → Al2O3 (s) + 3H2 (g) 900 − 755.070 Fe. The thermodynamic analysis shown in Table 3 illustrates the fact
Fe (s) + H2O (g) → FeO (s) + H2 (g) 900 − 5.684
that the water reduction reaction by Al, i.e. 2Al (s) + 3H2O (g) → Al2O3
2 Fe (s) + 3H2O (g) → Fe2O3 (s) + 3H2 (g) 900 − 8.068
3 Fe (s) + 4H2O (g) → Fe3O4 (s) + 4H2 (g) 900 − 6.905 (s) + 6 H (g), has the lower free energy as compared to the reaction
between Fe and water, e.g. Fe (s) + H2O (g) → FeO (s) + H2 (g). During
the austenitization stage, the Al of the alloy coating will therefore
3.4. Mechanism of H-uptake promote the generation and absorption of H.
In addition to the accelerated H formation, the presence of the Al-Si
In the present study, the industrial process that is responsible for the coating influences the permeability of H through the surface of the
H-uptake of the PHS was determined by means of the H-TDA. The alloy. Fig. 11(a) and (b) compare the temperature-dependent solubility
diffusible H was not introduced into the steel in the CA stage via the and diffusivity of H in pure Fe and pure Al [22–26]. The solubility data
atmosphere which has a relatively high H2 content, i.e. 15 vol% (Fig. 4). revels that, at elevated temperature, H is significantly more soluble in
The specimen after the (CA+Al) stage did not contain diffusible H. The solid austenite (γ-Fe) as compared to the liquid Al. In the austenitiza-
H-uptake of the PHS took place during the austenitization stage of the tion stage of the HPF process, the H that is absorbed in the Al coating
HPF process (Figs. 4 and 5). The furnace atmosphere during the HPF will, therefore, be partitioned preferentially to the austenite. The
process did not contain gaseous H2. However, water vapor in the fur- transfer of the H through liquid Al is very efficient because H diffuses
nace gas reacted with the steel surface to form Fe oxides and H2 gas very rapidly in liquid Al. At temperatures lower than 600 °C, both the H
during the heating to the austenitizing temperature [5]. This high solubility and the H diffusivity are extremely low in solid Al as com-
temperature atmospheric corrosion by water vapor results in the H pared to ferrite, α-Fe. This implies that the Al coating layer acts as a
absorption. The water vapor in the furnace is formed by the combustion strong diffusion barrier for H, keeping the H effectively in the Fe matrix
of natural gas (mainly CH4) used to heat the furnace. at room temperature.
The aluminized PHS was more sensitive to H-uptake than the un- It is important to note that during the austenitization stage of the
coated PHS (Figs. 4 and 5). It is known that the reaction between Al and HPF process, the liquid Al phase rapidly reacts with Fe in the matrix to

Fig. 11. Temperature-dependence of (a) the solubility and (b) the diffusivity of H in Fe and Al (PH2 = 1 atm). Literature data of the H solubility in Fe and Al provided
by San-Martin and Manchester [22] and Qui et al. [23], respectively. The H diffusion coefficients in Fe and Al were calculated by the equations provided by Fast [24],
Kikuch and McLellan [25], and Jakse and Pasturel [26]. (c) Room temperature diffusivity of H in Fe-Al alloys as a function of Al content of the alloys [25,26,29–35].

424
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

form Fe-Al or Fe-Al-Si intermetallics. In fact, Fe-Al intermetallics such which has transformed to martensite. The H atoms are tapped in the
as Fe3Al and FeAl are also susceptible to the moisture-induced H up- martensite, which has the high density of dislocation (1015–1016 m−2).
take, which can be explained by the same water reduction reaction The H is prevented from diffusing out of the martensite by the reacted
mechanism suggested for pure Al [27,28]. Therefore, the Fe-Al inter- coating layer that acts as a H-diffusion barrier.
metallics in the reacted coating should also have contributed to the H
absorption of the Al-Si coated PHS during the austenitization stage. 4. Conclusions
However, temperature-dependence of the H behavior in Fe-Al or Fe-Al-
Si intermetallics is not known and thus, further investigation will be The main conclusions of the present contribution are as follows:
needed to fully understand the role of the intermetallics in the H ab-
sorption and desorption of Al-Si coated PHS. Fig. 11(c) shows the lit- 1. Diffusible H is introduced in both uncoated PHS and aluminized PHS
erature data of the H diffusivity in Fe-Al alloys at room temperature, during the HPF process. The H-uptake is due to the high temperature
indicating that the H diffusion is generally slower in Fe-Al alloys than in atmospheric corrosion of the steel or the Al-Si coating in the aus-
pure Fe or lath martensitic steel [25,26,29–35]. This suggests that the tenitization stage of the HPF process.
reacted coating layer of the Al-Si coated PHS, consisting mainly of the 2. The H-atoms introduced in the uncoated PHS naturally escape from
Fe-Al intermetallics, likely lowers the permeability of H through the the steel at room temperature. Uncoated PHS aged at room tem-
surface of the alloy, preventing H from diffusing out of the martensite. perature exhibits a ductile tensile and bending performance.
Fig. 12 illustrates the mechanism for the H-uptake in aluminized 3. The H-uptake is larger in aluminized PHS than in uncoated PHS.
PHS. After the aluminizing process, the PHS surface consists of Al-10% Much larger amounts of diffusible H are absorbed in aluminized PHS
Si alloy coating layer and an inhibition layer at the coating/steel in- as compared to uncoated PHS. The Al-Si coating of PHS promotes
terface. The inhibition layer is formed during the hot dip aluminizing the generation of H. The water vapor in the furnace atmosphere
process. Water vapor in the furnace is formed during the combustion of reacts with the Al in the alloy coating, generating H. The high H
the natural gas used to heat the furnace. During the heating to the solubility in austenite and the high H diffusivity in Al result in an
austenitizing temperature, the water vapor in the furnace reacts with efficient partitioning of H to the PHS matrix during austenitization.
the Al of the Al-Si coating to form Al2O3 and H2. The H produced from H is inhibited from diffusing out of the steel matrix by the reacted
the reduction of H2O by the Al adsorbed into the coating layer. A me- Al-Si coating, which acts as a H-diffusion barrier.
chanism for the transport of H through the Al oxide film has been 4. Aluminized PHS is more sensitive to H embrittlement than uncoated
proposed by Draley and Ruther [36]. In this model, the H protons (H+) PHS. The H embrittlement of aluminized PHS is caused by diffusible
diffuse through the oxide film to the metal. Their diffusion is balanced H-atoms weakly bound to dislocation or grain boundary. This em-
by the diffusion of Al cations (Al3+) to the gas-oxide interface. The brittlement is characterized by a pronounced reduction of both the
protons are reduced to H atoms at the oxide-metal interface and H- tensile elongation and the bending angle at maximum load in three
atoms diffuse into the metal. During the austenitization of the PHS at point bending. The presence of diffusible H also causes inter-gran-
900 °C for 5–10 min, the Al-10%Si alloy coating reacts with the steel ular cracking and a refinement of the dimple size at the fracture
substrate. This results in the formation of Fe2Al5 and Fe-Al-Si inter- surface.
metallic phases. After the austenitization, the PHS specimens are 5. The BH treatment partially releases the H present in aluminized
transferred into water-cooled dies for HPF process. During cooling to PHS, and it suppresses the brittle fracture of aluminized PHS.
room temperature, most of the H is transferred into the steel matrix,

Fig. 12. Mechanism of the H-uptake of the aluminized PHS.

425
L. Cho et al. Materials Science & Engineering A 734 (2018) 416–426

Acknowledgment on the properties of press hardened boron steels, J. Mater. Process. Technol. 252
(2018) 90–104.
[15] S.-J. Lee, J.A. Ronevich, G. Krauss, D.K. Matlock, Hydrogen embrittlement of har-
The authors sincerely acknowledge the support of the POSCO dened low-carbon sheet steel, ISIJ Int. 50 (2) (2010) 294–301.
Technical Research Laboratories, Gwangyang. [16] K. Takasawa, R. Ikeda, N. Ishikawa, R. Ishigaki, Effects of grain size and dislocation
density on the susceptibility to high-pressure hydrogen environment embrittlement
of high-strength low-alloy steels, Int. J. Hydrog. Energy 37 (3) (2012) 2669–2675.
Appendix A. Supplementary material [17] Z. Zhang, J. Eckert, Unified tensile fracture criterion, Phys. Rev. Lett. 94 (9) (2005)
094301.
Supplementary data associated with this article can be found in the [18] T. Neeraj, R. Srinivasan, J. Li, Hydrogen embrittlement of ferritic steels: observa-
tions on deformation microstructure, nanoscale dimples and failure by nano-
online version at doi:10.1016/j.msea.2018.08.003 voiding, Acta Mater. 60 (13) (2012) 5160–5171.
[19] S. Lynch, A fractographic study of gaseous hydrogen embrittlement and liquid-
References metal embrittlement in a tempered-martensitic steel, Acta Metall. 32 (1) (1984)
79–90.
[20] Z.Y. Deng, J.M. Ferreira, Y. Sakka, Hydrogen‐generation materials for portable
[1] H. Karbasian, A. Tekkaya, A review on hot stamping, J. Mater. Process. Technol. applications, J. Am. Ceram. Soc. 91 (12) (2008) 3825–3834.
210 (15) (2010) 2103–2118. [21] P. Blackburn, E. Gulbransen, Aluminum reactions with water vapor, dry oxygen,
[2] J. Bian, H. Mohrbacher, Novel alloying design for press hardening steels with better moist oxygen, and moist hydrogen between 500 and 625 °C, J. Electrochem. Soc.
crash performance, in: Proceedings of International Symposium on New 107 (12) (1960) 944–950.
Developments in Advanced High-Strength Sheet Steels, 2013, pp. 23–27. [22] A. San-Martin, F. Manchester, The Fe-H (iron-hydrogen) system, Bull. Alloy Phase
[3] F. Borsetto, A. Ghiotti, S. Bruschi, Investigation of the high strength steel Al-Si Diagr. 11 (2) (1990) 173–184.
coating during hot stamping operations, Key Eng. Mater. Trans. Tech. Publ. (2009) [23] C. Qiu, G.B. Olson, S.M. Opalka, D.L. Anton, Thermodynamic evaluation of the Al-H
289–296. system, J. Phase Equilib. Diffus. 25 (6) (2004) 520–527.
[4] J. Woodtli, R. Kieselbach, Damage due to hydrogen embrittlement and stress cor- [24] J. Fast, Gases in Metals, Philips Technical Library, Macmillan, 1976.
rosion cracking, Eng. Fail. Anal. 7 (6) (2000) 427–450. [25] K. Kiuchi, R. McLellan, The solubility and diffusivity of hydrogen in well-annealed
[5] C. Georges, T. Sturel, P. Drillet, J.-M. Mataigne, Absorption/desorption of diffusible and deformed iron, Acta Metall. 31 (7) (1983) 961–984.
hydrogen in aluminized boron steel, ISIJ Int. 53 (2013) 1295–1304. [26] N. Jakse, A. Pasturel, Hydrogen diffusion in liquid aluminum from ab initio mole-
[6] M. Windmann, A. Röttger, W. Theisen, Phase formation at the interface between a cular dynamics, Phys. Rev. B 89 (17) (2014) 174302.
boron alloyed steel substrate and an Al-rich coating, Surf. Coat. Technol. 226 [27] C. Liu, E. Lee, C. McKamey, An environmental effect as the major cause for room-
(2013) 130–139. temperature embrittlement in FeAl, Scr. Metall. 23 (6) (1989) 875–880.
[7] M. Enomoto, D. Hirakami, T. Tarui, Thermal desorption analysis of hydrogen in [28] C. Liu, C. McKamey, E. Lee, Environmental effects on room-temperature ductility
high strength martensitic steels, Metall. Mater. Trans. A 43 (2) (2012) 572–581. and fracture in Fe3Al, Scr. Metall. Mater. 24 (2) (1990) 385–389.
[8] M. Wang, E. Akiyama, K. Tsuzaki, Hydrogen degradation of a boron-bearing steel [29] Y. Sakamoto, T. Mantani, Effect of quenching and tempering on diffusion of hy-
with 1050 and 1300MPa strength levels, Scr. Mater. 52 (5) (2005) 403–408. drogen in carbon steel, Trans. Jpn. Inst. Met. 17 (11) (1976) 743–748.
[9] A.-M. Brass, F. Guillon, S. Vivet, Quantification of hydrogen diffusion and trapping [30] M. Luppo, J. Ovejero-Garcia, The influence of microstructure on the trapping and
in 2.25 Cr-1Mo and 3Cr-1Mo-V steels with the electrochemical permeation tech- diffusion of hydrogen in a low carbon steel, Corros. Sci. 32 (10) (1991) 1125–1136.
nique and melt extractions, Metall. Mater. Trans. A 35 (5) (2004) 1449–1464. [31] H. Hosoda, K. Mizuuchi, K. Inoue, The effect of hydrogen on the hardness of Fe−Al
[10] W. Choo, J.Y. Lee, Thermal analysis of trapped hydrogen in pure iron, Metall. alloys, JOM 49 (8) (1997) 56–59.
Mater. Trans. A 13 (1) (1982) 135–140. [32] Y. Yang, S. Hanada, Absorption and desorption of hydrogen in Fe-40Al inter-
[11] F. Wei, T. Hara, K. Tsuzaki, Precise determination of the activation energy for metallic, Scr. Metall. Mater. 32 (11) (1995).
desorption of hydrogen in two Ti-added steels by a single thermal-desorption [33] U. Prakash, N. Parvathavarthini, R. Dayal, Effect of composition on hydrogen
spectrum, Metall. Mater. Trans. B 35 (3) (2004) 587–597. permeation in Fe–Al alloys, Intermetallics 15 (1) (2007) 17–19.
[12] K. Takai, R. Watanuki, Hydrogen in trapping states innocuous to environmental [34] Y. Chu, H. Yin, 10th International Congress on Metallic Corrosion, Vol. 3: Sessions
degradation of high-strength steels, ISIJ Int. 43 (4) (2003) 520–526. 10-13, Madras, India, 1987.
[13] D. Hirakami, S. Yamasaki, T. Tarui, K. Ushioda, Competitive phenomenon of hy- [35] T. Ishikawa, R. McLellan, The diffusivity of hydrogen in aluminum, Acta Metall. 34
drogen trapping and carbon segregation in dislocations introduced by drawing or (6) (1986) 1091–1095.
martensitic transformation of 0.35 mass% and 0.8 mass% C steels, ISIJ Int. 56 (2) [36] J. Draley, W. Ruther, Some unusual effects of hydrogen in corrosion reactions, J.
(2016) 359–365. Electrochem. Soc. 104 (6) (1957) 329–333.
[14] H. Järvinen, M. Honkanen, M. Järvenpää, P. Peura, Effect of paint baking treatment

426

You might also like