Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Current Opinion in Solid State and Materials Science 22 (2018) 1–7

Contents lists available at ScienceDirect

Current Opinion in Solid State and Materials Science


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

Hydrogen embrittlement in compositionally complex FeNiCoCrMn FCC


solid solution alloy
K.E. Nygren a, K.M. Bertsch b,c, S. Wang a, H. Bei d, A. Nagao c,e, I.M. Robertson a,c,f,⇑
a
Department of Engineering Physics, University of Wisconsin-Madison, Madison, WI 53706, USA
b
Department of Materials Science and Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA
c
International Institute for Carbon-Neutral Energy Research (WPI-I2CNER), Kyushu University, 744 Motooka, Nishi-ku, Fukuoka, Fukuoka 819-0395, Japan
d
Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
e
Material Surface & Interface Science Research Department, Steel Research Laboratory, JFE Steel Corporation, 1-1 Minamiwatarida-cho, Kawasaki-ku, Kawaksai, Kanagawa
210-0855, Japan
f
Department of Materials Science and Engineering, University of Wisconsin-Madison, Madison, WI 53706, USA

a r t i c l e i n f o a b s t r a c t

Article history: The influence of internal hydrogen on the tensile properties of an equi-molar FeNiCoCrMn alloy results in
Received 30 September 2017 a significant reduction of ductility, which is accompanied by a change in the fracture mode from ductile
Revised 6 November 2017 microvoid coalescence to intergranular failure. The introduction of 146.9 mass ppm of hydrogen reduced
Accepted 27 November 2017
the plastic strain to failure from 0.67 in the uncharged case to 0.34 and 0.51 in hydrogen-charged spec-
Available online 1 December 2017
imens. The reduction in ductility and the transition in failure mode are clear indications that this alloy
exhibits the classic signs of being susceptible to hydrogen embrittlement. The results are discussed in
Keywords:
terms of the hydrogen-enhanced plasticity mechanism and its influence on hydrogen-induced intergran-
High entropy alloy
Hydrogen embrittlement
ular failure. Furthermore, a new additional constraint that further promotes intergranular failure is intro-
Mechanical properties duced for the first time.
Grain boundaries Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction ion [12] and electron irradiation [13]. These properties have been
explained by the combined influences of the high entropy (thermo-
Equi-molar and multi-element alloys, including the group of dynamics), low diffusivity (kinetics), and lattice distortion (struc-
five-element alloys referred to as high-entropy alloys, provide ture) of these alloys as well as the sensitivity of the properties to
the opportunity to tailor the properties of the system; see reviews alloy composition [14].
by Miracle and Senkov [1], Zhang et al. [2] and Diao et al. [3]. Recently, it has been reported that the five-element high-
Single-phase FCC high-entropy alloys exhibit high thermal stability entropy alloy FeNiCoCrMn was not susceptible to hydrogen
[4], high malleability [5], slow diffusion [6], low yield strength, embrittlement, which would place it in a unique class of metals
high ductility [7], and high fracture toughness at cryogenic tem- and alloys [15,16]. Zhao et al. reported that the tensile properties
peratures [8]. The electronic, thermal and magnetic properties were similar in the uncharged and hydrogen-charged alloys, with
are also modified by increasing the number of elements [9,10]. It the hydrogen-charged alloy showing only a 5% decrease in ductility
was predicted theoretically and shown experimentally that the [15]. The fracture surfaces of the uncharged and hydrogen-charged
electrical and thermal conductivity is reduced significantly in specimens of the FeNiCoCrMn alloy were similar and consisted of
high-entropy alloys compared to other alloys, as is the electron microvoids, which indicate a ductile failure mode in both cases.
mean free path. These changes result in slower heat dissipation Based on these observations, Zhao et al. concluded that this alloy
from displacement cascades created by heavy-ion irradiation, was not embrittled by hydrogen. However, hydrogen-induced duc-
which results in a greater tolerance for the accumulation of radia- tile failure by microvoid coalescence is common, although it is
tion damage [9–11]. However, it was found that the number of ele- often accompanied by a change in the characteristics of the dim-
ments was not the dominating factor, as specific elements had a ples in terms of size and depth [17]. Luo et al. used electrochemical
different influence on the accumulation of the damage under both charging to introduce hydrogen into an alloy of the same composi-
tion and concluded that hydrogen, rather than being detrimental to
the mechanical properties as it is in most other alloys, was benefi-
⇑ Corresponding author. cial [16]. Hydrogen was reported to enhance strengthening and
E-mail address: irobertson@wisc.edu (I.M. Robertson).

https://doi.org/10.1016/j.cossms.2017.11.002
1359-0286/Ó 2017 Elsevier Ltd. All rights reserved.
2 K.E. Nygren et al. / Current Opinion in Solid State and Materials Science 22 (2018) 1–7

toughening mechanisms. This was attributed to hydrogen reducing Samples for examination in a scanning/transmission electron
the stacking-fault energy and therefore the phase stability of the microscope were prepared by focused ion beam machining using
alloy. an FEI Helios G4 PFIB CXe. The electron microscopy was performed
In this paper, the tensile behavior of the five-element alloy in a Tecnai TF-30 S/TEM, which was operated at 300 keV.
FeNiCoCrMn was studied in the absence and presence of hydrogen.
Gas-phase hydrogen charging at elevated temperature was used to 3. Results
introduce hydrogen uniformly throughout the sample. It will be
demonstrated unequivocally that this alloy is susceptible to hydro- The thermal desorption analysis spectra for the uncharged and
gen embrittlement. It undergoes a significant decrease in the strain hydrogen-charged alloy following the completion of the mechani-
to failure in the presence of hydrogen, which is accompanied by a cal property tests are shown in Fig. 1. The hydrogen concentration
transition in the fracture mode from ductile microvoid coalescence of the uncharged sample is defined to be the total hydrogen des-
to a mixed failure mode that includes intergranular failure. This orbed from room temperature to 643 °C, and that of the HC1 sam-
reduction in ductility coupled with the transition in fracture mode ple from room temperature to 660 °C, which corresponds to the
are classic characteristics of hydrogen embrittlement and are first desorption peak. From Fig. 1, the hydrogen concentration
explained in terms of the hydrogen enhanced localized plasticity remaining in the uncharged (open circles) and HC1 samples (filled
mechanism [18,19]. In addition, it will be proposed that the inabil- circles) was 1.4 mass ppm (76 at. ppm) and 146.9 mass ppm (8239
ity of the grain boundaries to realign towards the tensile axis dur- at. ppm), respectively. For comparison, the thermal desorption
ing uniaxial loading in the hydrogen-charged alloy provides an analysis spectra for a 2.0 mm thick Ni sample (filled squares) fol-
additional driving force for the hydrogen-induced transition in lowing charging at 200 °C for a duration of 160 h under a gas pres-
the failure mode. sure of 120 MPa, and for 4.0 mm thick SUS316L austenitic stainless
steel (filled triangles) charged with hydrogen in a 120 MPa hydro-
gen gas environment at 280 °C for 400 h are included in Fig. 1. The
2. Experimental procedure peak 1 hydrogen concentration in Ni and SUS316L were 42.5 mass
ppm (2492 at. ppm) and 104.2 mass ppm (5812 at. ppm), respec-
The near-equi-molar FeNiCoCrMn alloy was produced by arc tively. The rate-determining step for the thermal desorption pro-
melting, rolling, and annealing. Electrical discharge machining cess of these three materials is diffusion due to the low hydrogen
was used to cut flat dog-bone-type specimens with a gauge diffusivity [20]. The observation that the maximum in the desorp-
length of 10 mm and cross-sectional dimensions of approximately tion peak for the FeNiCoCrMn alloy and Ni occurs at 409 °C sug-
2 mm  2 mm from the cold-rolled sheets. The specimens were gests that the hydrogen diffusivity is similar in both materials
cut such that the longitudinal axis was perpendicular to the roll- and lower than in stainless steels. The calculated initial hydrogen
ing direction. The specimens were annealed at 1200 °C for 24 h. concentration for the charging conditions used for Ni is 41.3 mass
The resulting microstructure consisted of approximately equiaxed ppm (2423 at. ppm) [21] and the thermal desorption analysis of Ni
grains with a mean grain size of 70 mm; annealing twin bound- presented in Fig. 1 was conducted after 11 days of storage at room
aries were frequently observed making the effective grain size temperature. This result suggests that the outgassing of hydrogen
50 mm. from Ni was negligible after 11 days at room temperature. The
The samples were charged with hydrogen in an autoclave at a hydrogen concentration in the FeNiCoCrMn alloy was measured
hydrogen gas pressure of 120 MPa at 200 °C for 160 h. The total following a total time at room temperature of between 8 – 10 days,
time the samples were at room temperature was between 8 and by comparison with the Ni result it is reasonable to conclude that
10 days for transportation to and from the hydrogen-charging the change in the hydrogen concentration due to the time at room
facility. At all other times they were stored in liquid nitrogen. Uni- temperature was negligible. In comparing this hydrogen
axial tensile tests were conducted at a constant displacement rate
of 1.7  10 5 mm s 1 (an initial strain rate of approximately
1.7  10 6 s 1) using a MTS QTest/5 tensile machine with a 5 kN
load cell. Two hydrogen-charged samples were tested and are
referred to as HC1 and HC2. HC1 was tested shortly after it was
removed from the liquid nitrogen storage vessel and brought to
room temperature, whereas HC2 was mechanically polished for
initial electron backscattering diffraction (EBSD) analysis before
testing. The polishing and EBSD analysis took approximately 10 h.
The hydrogen content of the uncharged and hydrogen-charged
samples was measured using gas chromatograph thermal desorp-
tion analysis after the tensile tests were complete. This ensured
that the 8 and 10 days the samples were at room temperature
did not result in loss of hydrogen. A ramp rate of 200 °C h 1 was
used in the thermal desorption test from room temperature to
800 °C.
The fracture surfaces and external surfaces were investigated
using scanning electron microcopy (SEM) with either a LEO
1530–1 FESEM or an FEI Helios G4 PFIB CXe with an ElstarTM
SEM column being used. Electron dispersive spectroscopy (EDS)
and EBSD measurements were performed using EDA TEAMTM
EDS software and instrumentation attached to the FEI Helios G4
PFIB. The EDS measurements were performed using an Octane Elite
Fig. 1. Thermal desorption analysis spectra of the uncharged and hydrogen-
silicon drift detector. EBSD measurements were performed using charged FeNiCoCrMn alloy; heating rate = 200 °C/h. Data for Ni and SUS316L
an Hikari EBSD camera using a step size of 0.5 mm, accelerating austenitic stainless steel (heating rate = 200 °C/h) are included for comparison.
voltage of 30 kV, and current of 26 nA.  HC1; s uncharged; N Ni; j SUS 316 L austenitic stainless steel.
K.E. Nygren et al. / Current Opinion in Solid State and Materials Science 22 (2018) 1–7 3

concentration to that in Ni and SUS 316L austenitic stainless steel, Closer inspection of the faceted surfaces of both HC1 and HC2
it is found that the concentration in the FeNiCoCrMn alloy is indicated that they were not flat but showed evidence of slip traces
greater. This suggests that the FeNiCoCrMn alloy has a higher sol- or steps; these are common features on a hydrogen-induced inter-
ubility of hydrogen. Preliminary microstructural analysis showed granular facet [21,22]. On the same facet, these steps could lie in
no evidence for the formation of hydrides in the hydrogen- three directions. On some surfaces these steps were curved signif-
charged material, indicating that hydrogen is retained in the lattice icantly, which is an indication that a high degree of deformation
and in other traps such as dislocations, grain boundaries, and had occurred. These steps could result from the disruptions of
precipitates. the grain boundaries caused by the intersection and transmission
The stress-strain curves for the uncharged and hydrogen- of dislocation slip systems or deformation twins [23]. An example
charged samples are compared in Fig. 2. Both HC1 and HC2 exhib- of this structure is shown in the fractograph presented in Fig. 5.
ited higher yield strengths than the uncharged material; the 0.2% The disruptions on the free surfaces caused by the intersections
offset yield strengths were 152, 214, and 207 MPa in the of dislocation slip systems and deformation twins were examined
uncharged, HC1, and HC2 samples, respectively. The ultimate ten- along the gauge length. For the uncharged sample, intersections on
sile strengths were 507, 528, and 475 MPa in the uncharged, HC1, the surface occurred in up to three different directions and the
and HC2 samples, respectively. Both hydrogen-charged samples steps were distinct and high. In contrast, in the hydrogen-
failed at a significantly lower strain than the uncharged sample. charged material, activity was observed on multiple slip systems
The strain at failure was 67% in the uncharged sample, whereas but in most grains the steps were less distinct and shallower in
it was 51% and 34% for HC1 and HC2, respectively. It is noted that appearance. An additional difference between the uncharged and
Gludovatz et al. reported a range of failure strains for this alloy hydrogen-charged samples was the presence of microcracks along
composition in the absence of hydrogen, suggesting that variations the grain boundaries away from fracture surface in the hydrogen-
in internal defects, oxides, grain size or grain boundary misorienta- charged samples only. Examples of these microstructural features
tion distribution may be the root cause of the scatter in the failure are compared for the uncharged and hydrogen-charged samples
strains [8]. However, the difference of strain to failure between the in Fig. 6(a) and (b), respectively. In Fig. 6, the images are not from
uncharged and hydrogen-charged materials in the current study is the same distance from the fracture surface; Fig. 6(a) is approxi-
statistically significant and indicates this material is embrittled by mately 1.6 mm from the fracture surface in the uncharged sample
hydrogen. and Fig. 6(b) is approximately 300 lm from the fracture surface in
The fracture surfaces of the uncharged, HC1, and HC2 samples HC2. Examples of these traces and steps are indicated with arrows
are compared in the fractographs presented in Fig. 3. The in Fig. 6(a) and (b), and examples of microcracks with arrowheads
uncharged sample exhibited a ductile microvoid coalescence fail- in Fig. 6(b). To confirm that these microcracks lie along grain
ure mode, Fig. 3(a), with precipitates evident in the microvoids, boundaries and the fracture mode is appropriately labelled inter-
see Fig. 3(d). EDS analysis of a precipitate within a void, the results granular, an EBSD map of the same area prior to loading is pre-
of which are presented in Fig. 4 as elemental maps, confirmed that sented as Fig. 6(c). Superimposing the EBSD map on the
the precipitates were rich in Cr and Mn and deficient in Fe, Co and deformed image confirms the microcracks lie along the grain
Ni. In contrast, both hydrogen-charged samples exhibited predom- boundaries; the locations of the microcracks are indicated by
inantly intergranular failure as evidenced by the fractographs pre- arrowheads in Fig. 6(c). Another important feature revealed in
sented in Fig. 3(b) and (c). However, there were subtle differences Fig. 6 is that the grains in the uncharged alloy rotate towards
between HC1 and HC2. HC1 showed extensive secondary cracking, and elongate in the direction of the tensile axis, which runs from
as well as dimples and ductile fracture features in multiple regions, top to bottom in this figure. In contrast, the grains in the
including on some surfaces that appeared macroscopically flat, see hydrogen-charged alloy do not show the same rotation or elonga-
Fig. 3(e). Similar ductile features were identified on the fracture tion. This can be seen by comparing Fig. 6(b) and (c). This is the
surface of the HC2 sample, although with a lower frequency. This first report of hydrogen having an influence of the shape of the
observation of a change in fracture mode is also consistent with grain and restricting its ability to rotate and elongate in response
this material being susceptible to hydrogen embrittlement. to the applied load.
To determine the deformation processes associated with the
change in the fracture mode, the microstructural state needs to
be assessed. Preliminary observations of the evolved microstruc-
ture in the uncharged and hydrogen-charged samples beneath
the fracture surfaces revealed that the deformation occurred by
dislocation slip and deformation twinning. Examples of the
deformed microstructure adjacent to the fracture surfaces of the
uncharged and hydrogen-charged samples are compared in
Fig. 7. The region beneath a microvoid in the uncharged specimen
and the faceted fracture surface of the HC1 sample are shown in
the STEM diffraction contrast images presented in Fig. 7(a) and
(b), respectively. Here it is important to remember that these frac-
ture surfaces are different and they were formed at different
strains, 67% and 51% for the uncharged and hydrogen-charged
samples, respectively. In the uncharged alloy, a region with sub-
grains elongated parallel to the void surface exists immediately
below the fracture surface; this region is indicated with an arrow-
head in Fig. 7(a). The presence of this sub-surface region is consis-
tent with observations made in other systems [24,25]. Beyond this
layer, the microstructure transitions into a dislocation cell struc-
ture on which deformation twins are superimposed; an example
of a deformation twin is indicated by the double arrow. Beneath
Fig. 2. Engineering stress-strain curves for uncharged, HC1, and HC2 specimens. the intergranular facet in HC1, a dislocation cell structure exists
4 K.E. Nygren et al. / Current Opinion in Solid State and Materials Science 22 (2018) 1–7

Fig. 3. Fractographs of (a) uncharged, (b) HC1, and (c) HC2 with magnified views of the fracture surfaces (d) uncharged, (e) HC1, and (f) HC2.

Fig. 4. Elemental map showing the composition of the particle associated with the void in uncharged material. (a) SEM image of the void, (b) Co (c) Ni, (d) Fe, (e) Cr, and
(f) Mn.

as shown in Fig. 7(b). The presence of deformation twins was less


evident in the hydrogen-charged alloy. Despite the smaller strain
to failure, the dislocation cell size (300–500 nm) was similar to
that in the uncharged samples and the cell walls were thicker. This
type of dislocation cell structure has been observed beneath inter-
granular facets in other hydrogen embrittled systems [22,26]. The
observation of a comparable cell size and thicker cell walls despite
the differences in strain at failure are indicative of hydrogen having
enhanced the evolution of the microstructure through the HELP
mechanism [19].

4. Discussion

It is demonstrated for the first time that the FCC equi-molar,


single-phase FeNiCoCrMn alloy is susceptible to hydrogen embrit-
tlement. This embrittlement was accompanied by a change in frac-
ture mode from ductile microvoid coalescence to intergranular
failure. In this regard, this alloy behaves similar to Ni, which
undergoes a transition in the failure mode from transgranular to
intergranular with increasing hydrogen concentration [22]. A tran-
sition in failure mode also occurs in BCC Fe [26]. For both Ni [22]
and Fe [26], it has been proposed that the presence of hydrogen
Fig. 5. Fractograph of a facet in HC1. on the grain boundary, while a necessary condition for the ultimate
K.E. Nygren et al. / Current Opinion in Solid State and Materials Science 22 (2018) 1–7 5

Fig. 6. Deformation on the free surface (a) uncharged specimen (b) HC2. (c) EBSD of the free surface region shown in (b) prior to loading. Arrows indicate the traces and steps,
arrowheads microcracks. The tensile axis is vertical.

Fig. 7. Bright-field STEM diffraction contrast micrographs of the deformed microstructure beneath fracture surfaces of (a) the uncharged specimen and (b) HC1. Arrowhead
marks the near-surface region with subgrains and the double arrow marks the deformation twins.

failure of the grain boundary, is by itself not sufficient. It was pro- dynamic exchange of hydrogen with the grain boundary and this
posed that the additional driving force was derived from the local was considered to play an important role in establishing conditions
stresses associated with the work-hardening that occurred in the that promoted grain boundary decohesion. Initial evaluation of the
grain interior and the changes in the grain boundary structure that deformed microstructure in the hydrogen-charged FeNiCoCrMn
accompany the incorporation and emission of dislocations as strain samples shows an extensive work-hardened microstructure
was transferred across the grain boundary in hydrogen. Support for accompanying the intergranular fracture, Fig. 7.
these additional mechanisms was found in the microstructure For the first time it is reported that hydrogen influences the
beneath the intergranular facets in Ni and Fe, which was further ability of the grains to elongate and rotate in the direction of the
developed than suggested by the magnitude of the macroscopic tensile axis. A similar observation has been made in hydrogen-
strain to failure [22,26]. The redistribution of hydrogen through charged Ni following uniaxial loading [27]. These observations sug-
transport by glissile dislocations was considered to create a gest that an additional compatibility constraint will be established
6 K.E. Nygren et al. / Current Opinion in Solid State and Materials Science 22 (2018) 1–7

across the grain boundaries as the adjoining grains cannot rotate 4. To understand the influence of hydrogen on the mechanical
and elongate in response to the deformation. This additional con- properties of this class of alloys, further assessment of the frac-
straint would further favor failure of a grain boundary that has ture toughness and fatigue-crack growth as well as determina-
been weakened by the presence of hydrogen. tion of fundamental properties such as the diffusivity,
These results are in marked contrast to those that have been permeability and solubility of hydrogen are required.
published [15,16]. Both previous studies claimed this alloy was
immune to hydrogen embrittlement, which would place it in a
unique group of metals that exhibit negligible hydrogen embrittle- Acknowledgements
ment. Zhao et al. reported that the FeNiCoCrMn alloy, despite con-
taining a higher hydrogen concentration than 304 and 316L, was SW, KEN, KMB and IMR gratefully acknowledge the support of
more resistant to hydrogen embrittlement than these austenitic the National Science Foundation through Award No. CMMI-
stainless steels, and did not exhibit a transition in fracture mode 1406462. AN acknowledges the support from JFE Steel Corporation.
at a hydrogen concentration of 76.5 mass ppm [15]. Here it is KB and KN acknowledge partial support from the World Premier
important to note several characteristics of hydrogen embrittle- International Research Center Initiative (WPI), MEXT, Japan,
ment. It is well known that hydrogen embrittlement occurs within through the International Institute for Carbon-Neutral Energy
a well-defined window of strain rate and temperature [28]. The Research (I2CNER) of Kyushu University. HB was supported by
embrittling potency of hydrogen is generally greater for decreasing the US Department of Energy, Office of Science, Basic Energy
strain rates [28]. Zhao et al. used a strain rate of 1.7  10 4 s 1 Sciences, Materials Sciences and Engineering Division. The electron
whereas an initial strain rate of 1.7  10 6 s 1 was used in this microscopy was carried out using facilities and instrumentation in
study. Additionally, the hydrogen content in their alloy was 76.5 the Materials Science Center at the University of Wisconsin-
mass ppm compared to 146.9 mass ppm, which might not be suf- Madison, which is partially supported by NSF through the Materi-
ficient to cause a change in the fracture mode from transgranular als Research Science and Engineering Center (DMR-1121288). This
to intergranular. Such a dependence of the failure mode transition manuscript has been authored by UT-Battelle, LLC under Contract
on the hydrogen concentration has been reported in Ni [29]. Fur- No. DE-AC05-00OR22725 with the U.S. Department of Energy.
ther, the dichotomy presented by the observations by Luo et al. The United States Government retains and the publisher, by
[16] can be explained by the hydrogen-charging method accepting the article for publication, acknowledges that the United
employed, the sluggish diffusion of hydrogen in this alloy, and States Government retains a non-exclusive, paid-up, irrevocable,
the use of a strain rate of 1.0  10 4 s 1. Electrochemical charging world-wide license to publish or reproduce the published form of
conditions at room temperature can create a high surface concen- this manuscript, or allow others to do so, for United States Govern-
tration of trapped hydrogen, which may include the formation of a ment purposes. The Department of Energy will provide public
near-surface hydride. Luo et al. reported intergranular fracture in access to these results of federally sponsored research in accor-
the near-surface region and this may be explained by a locally high dance with the DOE Public Access Plan (http://energy.gov/down-
hydrogen concentration. The proposal that the beneficial effects loads/doe-public-access-plan).
are attributable to hydrogen reducing the stacking-fault energy
does not agree with observations of the impact of hydrogen on
Appendix A. Supplementary material
the mechanical properties of other FCC alloys in which the magni-
tude of the reduction in the stacking-fault energy has been mea-
Supplementary data associated with this article can be found, in
sured directly [30]. These alloys remained susceptible to
the online version, at https://doi.org/10.1016/j.cossms.2017.11.002.
hydrogen embrittlement despite the reduction in the stacking-
fault energy. Finally, they proposed that hydrogen affected the
References
propensity for a system to deform by twinning. However, there
has been no evidence to support that a hydrogen-induced reduc- [1] D.B. Miracle, O.N. Senkov, A critical review of high entropy alloys and related
tion in the stacking-fault energy results in an increased propensity concepts, Acta Mater. 122 (2017) 448–511.
of the system to deform by twinning or that deformation twinning [2] Y. Zhang, T.T. Zuo, Z. Tang, M.C. Gao, K.A. Dahmen, P.K. Liaw, Z.P. Lu,
Microstructures and properties of high-entropy alloys, Prog. Mater. Sci. 61
is enhanced by hydrogen [31]. Indeed, initial assessment of the (2014) 1–93.
evolved microstructure does not show hydrogen increasing the [3] H.Y. Diao, R. Feng, K.A. Dahmen, P.K. Liaw, Fundamental deformation behavior
propensity for deformation and if anything it may show the oppo- in high-entropy alloys: an overview, Curr. Opin. Solid State Mater. Sci. 21 (5)
(2017) 252–266.
site effect. However, additional analysis is needed to confirm this
[4] A. Gali, E.P. George, Tensile properties of high-and medium-entropy alloys,
statement. Nygren has reported deformation twinning in the near Intermetallics 39 (2013) 74–78.
fracture surface region of 316L. However, these twins were super- [5] F. Otto, Y. Yang, H. Bei, E.P. George, Relative effects of enthalpy and entropy on
imposed on a dislocation cell structure and it was shown that the the phase stability of equiatomic high-entropy alloys, Acta Mater. 61 (7)
(2013) 2628–2638.
twins appeared only after the dislocation cell structure could not [6] S.-Y. Chang, C.-E. Li, Y.-C. Huang, H.-F. Hsu, J.-W. Yeh, S.-J. Lin, Structural and
continue to support the strain [32]. thermodynamic factors of suppressed interdiffusion kinetics in multi-
component high-entropy materials, Scient. Rep. 4 (2014).
[7] F. Otto, A. Dlouhý, C. Somsen, H. Bei, G. Eggeler, E.P. George, The influences of
5. Conclusions temperature and microstructure on the tensile properties of a CoCrFeMnNi
high-entropy alloy, Acta Mater. 61 (15) (2013) 5743–5755.
[8] B. Gludovatz, A. Hohenwarter, D. Catoor, E.H. Chang, E.P. George, R.O. Ritchie, A
1. The high-entropy FCC alloy FeNiCoCrMn is embrittled in the
fracture-resistant high-entropy alloy for cryogenic applications, Science 345
presence of hydrogen and this is accompanied by a change in (6201) (2014) 1153–1158.
fracture mode from ductile microvoid coalescence to intergran- [9] Y. Zhang, G.M. Stocks, K. Jin, C. Lu, H. Bei, B.C. Sales, L. Wang, L.K. Beland, R.E.
Stoller, G.D. Samolyuk, M. Caro, A. Caro, W.J. Weber, Influence of chemical
ular failure.
disorder on energy dissipation and defect evolution in concentrated solid
2. The solubility of hydrogen in this alloy is higher than in other solution alloys, Nat. Commun. 6 (2015).
FCC alloys such as types 304 and 316L austenitic stainless steels [10] Y. Zhang, K. Jin, H. Xue, C. Lu, R.J. Olsen, L.K. Beland, M.W. Ullah, S. Zhao, H. Bei,
or Ni. D.S. Aidhy, G.D. Samolyuk, L. Wang, M. Caro, A. Caro, G.M. Stocks, B.C. Larson, I.
M. Robertson, A.A. Correa, W.J. Weber, Influence of chemical disorder on
3. The hydrogen diffusivity in this alloy appears to be similar to energy dissipation and defect evolution in advanced alloys, J. Mater. Res. 31
that in Ni and lower than that in austenitic stainless steels. (16) (2016) 2363–2375.
K.E. Nygren et al. / Current Opinion in Solid State and Materials Science 22 (2018) 1–7 7

[11] Y. Zhang, S. Zhao, W.J. Weber, K. Nordlund, F. Granberg, F. Djurabekova, [22] M.L. Martin, B.P. Somerday, R.O. Ritchie, P. Sofronis, I.M. Robertson, Hydrogen-
Atomic-level heterogeneity and defect dynamics in concentrated solid- induced intergranular failure in nickel revisited, Acta Mater. 60 (6–7) (2012)
solution alloys, Curr. Opin. Solid State Mater. Sci. 21 (5) (2017) 221–237. 2739–2745.
[12] C. Lu, L. Niu, N. Chen, K. Jin, T. Yang, P. Xiu, Y. Zhang, F. Gao, H. Bei, S. Shi, M.-R. [23] J. Kacher, B. Eftink, B. Cui, I.M. Robertson, Dislocation interactions with grain
He, I.M. Robertson, W.J. Weber, L. Wang, Enhancing radiation tolerance by boundary, Curr. Opin. Solid State Mater. Sci. 18 (2014) 227–243.
controlling defect mobility and migration pathways in multicomponent [24] I.M. Robertson, P. Sofronis, A. Nagao, M.L. Martin, S. Wang, D.W. Gross, K.E.
single-phase alloys, Nat. Commun. 7 (2016) 13564. Nygren, Hydrogen embrittlement understood, Metall. Mater. Trans. A 46
[13] M.-R. He, S. Wang, S. Shi, K. Jin, H. Bei, K. Yasuda, S. Matsumura, K. Higashida, I. (2015) 1–19.
M. Robertson, Mechanisms of radiation-induced segregation in CrFeCoNi- [25] A. Nagao, M.L. Martin, M. Dadfarnia, P. Sofronis, I.M. Robertson, The effect of
based single-phase concentrated solid solution alloys, Acta Mater. 126 (2017) nanosized (Ti, Mo)C precipitates on hydrogen embrittlement of tempered lath
182–193. martensitic steel, Acta Mater. 74 (2014) 244–254.
[14] J.-W. Yeh, Recent progress in high-entropy alloys, Annales De Chimie – Science [26] S. Wang, M.L. Martin, P. Sofronis, S. Ohnuki, N. Hashimoto, I.M. Robertson,
des Materiaux 31 (2006) 633–648. Hydrogen-induced intergranular failure of iron, Acta Mater. 69 (2014) 275–
[15] Y. Zhao, D.-H. Lee, M.-Y. Seok, J.-A. Lee, M.P. Phaniraj, J.-Y. Suh, H.-Y. Ha, J.-Y. 282.
Kim, U. Ramamurty, J.-I. Jang, Resistance of CoCrFeMnNi high-entropy alloy to [27] K. Bertsch, Hydrogen Effects on the Evolution of Plastic Deformation in
gaseous hydrogen embrittlement, Scripta Mater. 135 (2017) 54–58. Polycrystalline Nickel: A Mechanism for Intergranular Failure, Materials
[16] H. Luo, Z. Li, D. Raabe, Hydrogen enhances strength and ductility of an Science and Engineering, University of Illinois Urbana-Champaign, 2017.
equiatomic high-entropy alloy, Scient. Rep. 7 (1) (2017) 9892. [28] T. Boniszewski, G.C. Smith, The influence of hydrogen on the plastic
[17] A. Macadre, N. Nakada, T. Tsuchiyama, S. Takaki, Critical grain size to limit the deformation ductility and fracture of nickel in tension, Acta Metall. 11 (3)
hydrogen-induced ductility drop in a metastable austenitic steel, Int. J. (1963) 165–178.
Hydrogen Energy 40 (33) (2015) 10697–10703. [29] D.H. Lassila, H.K. Birnbaum, The effect of diffusive segregation on the fracture
[18] I.M. Robertson, H.K. Birnbaum, P. Sofronis, Hydrogen Effects on Plasticity, in: J. of hydrogen charged nickel, Acta Metall. 36 (10) (1988) 2821–2825.
P. Hirth, L. Kubin (Eds.), Dislocations in Solids, Elsevier, 2009, pp. 249–293. [30] P.J. Ferreira, I.M. Robertson, H.K. Birnbaum, Influence of hydrogen on the
[19] H.K. Birnbaum, P. Sofronis, Hydrogen-enhanced localized plasticity-a stacking-fault energy of an austenitic stainless steel, Mater. Sci. For. 207–209
mechanism for hydrogen-related fracture, Mater. Sci. Eng. A A176 (1–2) (1996) 93–96.
(1994) 191–202. [31] J.M. Rigsbee, R.B. Benson, A TEM investigation of hydrogen-induced
[20] M. Dadfarnia, A. Nagao, S. Wang, M.L. Martin, B.P. Somerday, P. Sofronis, deformation twinning and associated martensitic phases in 304-type
Recent advances on hydrogen embrittlement of structural materials, Int. J. stainless steel, J. Mater. Sci. 12 (2) (1977) 406–409.
Fract. 196 (1) (2015) 223–243. [32] K. Nygren, The Influence of Hydrogen on the Evolving Microstructure During
[21] S. Bechtle, M. Kumar, B.P. Somerday, M.E. Launey, R.O. Ritchie, Grain-boundary Fatigue Crack Growth in Metastable and Stable Austenitic Stainless Steels,
engineering markedly reduces susceptibility to intergranular hydrogen Materials Science and Engineering, University of Illinois Urbana-Champaign,
embrittlement in metallic materials, Acta Mater. 57 (14) (2009) 4148–4157. 2016.

You might also like