Вы находитесь на странице: 1из 9

Journal of Alloys and Compounds 720 (2017) 451e459

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: http://www.elsevier.com/locate/jalcom

Mechanisms of hydrogen trapping in austenitic, duplex, and super


martensitic stainless steels
R. Silverstein*, D. Eliezer
Department of Material Engineering, Ben-Gurion University of the Negev, Beer-Sheva, Israel

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

Article history: Hydrogen interaction with structural materials, especially stainless steels, is of great importance due to
Received 23 February 2017 the acute effect that it can have on them. Stainless steels have become very common in many applica-
Received in revised form tions, and in water and high pressure environments in particular, due to their high levels of corrosion
11 May 2017
resistance and broad range of strength. Steel's durability is very much dependent on its microstructure
Accepted 27 May 2017
and interaction with hydrogen. The action of hydrogen can lead to changes in mechanical properties,
Available online 29 May 2017
phase transformation and eventually to environmentally-assisted failure, which is known as hydrogen
embrittlement (fracture). The susceptibility of steels to this hydrogen fracture mechanism is directly
Keywords:
Hydrogen trapping
related to the interaction between traps (defects) and hydrogen. In this research, we study hydrogen
Duplex stainless steel fracture mechanisms through hydrogen interaction with trapping sites by thermal desorption spec-
Super martensitic stainless steel trometry (TDS), and the calculation of hydrogen trapping energies states. Microstructure effects on
Austenitic stainless steel hydrogen were investigated by exploring different stainless steels, including: austenitic stainless steel
Thermal desorption spectrometry (TDS) (AUSS), ferritic-austenitic (duplex) stainless steel (DSS), and super martensitic stainless steel (SMSS). The
objective of this study is to determine the inuences of thermal desorption analysis on the crystal
structure of different stainless steels in order to better understand the trapping mechanisms of hydrogen
in a variety of structure materials. It was found that the AUSS has the greatest stability of austenitic (g)
phasee ~22% higher than DSS and ~45% higher than SMSS. Moreover, the AUSS presented the lowest
hydrogen trapping values of ~31% compared with DSS and ~25% compared with SMSS.
Hydrogen fracture mechanism was found to be highly dependent on the hydrogen trapping states and
even more on the g-phase stability. The hydrogen trapping mechanisms are discussed in detail.
2017 Elsevier B.V. All rights reserved.

1. Introduction Understanding the hydrogen embrittlement phenomenon in


the variety of stainless steels requires true insight of hydrogen
Stainless steels are the most common structural material for a solubility, diffusivity and distribution within their different micro-
variety of technologies and applications, particularly for oil, gas, structures, in particular inside phases and defects [6e8]. Since
pipeline and pressure applications e all of these industries combine steel's defects (trapping sites for hydrogen) affect the steel's diffu-
hydrogen and mechanical load [1,2]. These materials are very sivity, the crack initiation is highly dependent on the trapping en-
attractive due to their high level of corrosion resistance and varied ergy states. It is known from earlier publications [9e14] that traps
mechanical properties. Problems in maximizing those steels' full with activation energy  60 kJ/mol (also characterized as reversible
potential may arise from their interaction with hydrogen. In many traps) will have a major inuence on the material's susceptibility to
applications these steels are exposed to hydrogen, which may lead hydrogen embrittlement. In this work we study hydrogen trapping
to a deleterious effect known as hydrogen embrittlement. The base behavior in three different stainless steels: austenitic stainless steel
condition of this fracture mechanism is highly dependent on the (316), ferritic-austenitic (duplex) stainless steel (SAF 2205), and
dislocation process and is determined by hydrogen distribution super martensitic stainless steel (SMSS).
combined with the local stress state in the material [3e5]. The purpose of this work is to study hydrogen thermal
desorption behavior in order to better understand it's trapping
mechanisms in a variety of structural materials.
* Corresponding author. We study the hydrogen trapping behavior with thermal
E-mail address: barrav@post.bgu.ac.il (R. Silverstein). desorption spectrometry (TDS), and hydrogen trapping energies

http://dx.doi.org/10.1016/j.jallcom.2017.05.286
0925-8388/ 2017 Elsevier B.V. All rights reserved.
452 R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459

are analyzed by the Lee and Lee model [15]. The analyzed data was
conrmed by X-ray diffraction (XRD) and microstructural
observations.

2. Experimental procedure

2.1. Microstructure analysis

In this research we used three different stainless steels: duplex


stainless steels (DSS), SAF 2205, austenitic stainless steel (AUSS),
316L, and super martensitic stainless steel (SMSS). These steels
were received in the fully annealed condition, as a 2 mm thick plate.
The chemical compositions of these steels are presented in Table 1.
These chemical compositions were conrmed to belong to the
aforementioned structures according to the attached Scheler di-
agram, Fig. 1.
The hydrogenation technique was electrochemical (cathodic)
charging. The charging was performed at room temperature (RT) in
a 0.5 N H2SO4 (sulfuric acid) water solution and 0.25 g l1 of NaAsO2
(sodium arsenide) with a constant current density of 50 mA cm2 Fig. 1. Scheler diagram for different steel compositions.
for 72 h (h). The microstructure, phase composition, and lattice
parameter of DSS were examined both before and after hydroge-
nation, by means of an X-ray diffractometer (XRD). XRD patterns Fig. 2. This pattern is compared to an uncharged sample. The un-
were measured at RT using a Philips PW 1050/70 diffractometer, charged DSS (black line) displays ferritic-austenitic phases (a/g) in
operating at a power of 40 kV and 30 mA, with Cu anode, and Ni- equal amounts of about 50% for each phase.
ltered generating Cu-Ka radiation (1.54 ). Data collection was The hydrogenation process reveals the signicant changes in the
performed by step scanning of the sample in a 2q range between g-phase compared with those of the a-phase. The charged samples
30 and 100 at steps of 0.02 with 4e10 s per step depending on showed a signicant decrease in g-phase intensity and shifts to
peak intensity, and a scan rate of ~0.3 /min. small 2q values; the latter change indicates a larger lattice param-
The charged specimens were aged at RT and reexamined after eter, which can be explained by hydrogen solubility in that phase.
one month, in order to observe the stability of hydrogen-induced The signicant changes between both phases are related to the
phase transformations and lattice parameters after hydrogen lower diffusion coefcient and the greater solubility of hydrogen in
desorption. the g-phase [16,17]. These changes were seen simultaneously to the
formation of an additional -martensite phase, an hcp solid solution
phase, and g*-phase, an additional fcc phase with a 2% higher
2.2. Thermal desorption spectrometry (TDS)

The characteristics of hydrogen desorption, and trapping states


were investigated by means of TDS. This technique involves accu-
rate measurement of the desorption rate of hydrogen atoms, as
solute or trapped in the material, while treating the sample by non-
isothermal heating at a known rate under UHV, ~10 mPa. In this
work, the samples were heated from RT to 500  C at constant
heating rates of 2  C/min, 4  C/min and 6  C/min. The mass spec-
trometer was operated under the fast multiple mode detection; the
measured intensity channel was set to 2 amu in order to detect
hydrogen desorption. The working procedure, as described else-
where [15], allowed for the identication of different types of traps
that coexist in the specimen.

3. Results and discussion

3.1. Microstructure changes in the presence of hydrogen

3.1.1. Duplex stainless steel


The XRD diffraction pattern obtained from 72 h cathodic Fig. 2. Phase quantities of 72 h cathodic charged: DSS, AUSS and SMSS, after one
hydrogen charged DSS and aged for one month at RT is shown in month at RT.

Table 1
Chemical compositions of the studied stainless steels (DSS, AUSS, SMSS) in wt%.

Sample C S P Mn Si Ni Cr Mo N Cu

SAF 2205 0.025 0.004 0.022 0.411 0.288 6.745 24.012 4.63 0.3 0.072
316L 0.04 0.006 0.04 1.164 0.375 10.693 15.612 2.748 0.039 0.274
SMSS 0.063 0.002 0.021 1.859 0.364 7.002 10.513 3.315 0.009 0.467
R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459 453

Fig. 3. SEM observation of DSS samples (a) uncharged at lower magnication, (b) uncharged at higher magnication, (c) 72 h cathodic charged, showing the formation of needle
shape -martensite phase after one month at RT, (d) surface cracking in 72 h cathodic charged DSS- non-etched, and (e) etched-red arrows pointing at cracking sites. (For
interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

lattice parameter, Fig. 2. Regarding, a-phase reections, they gradient implies large tensile stress on the surface layer, and has a
exhibited peaks' broadening, which indicates the formation of direct effect on the formation of -martensite, in addition to surface
smaller grain size. These observations were supported by previous cracking [24].
works on cathodic charged DSS alloys [12,18e20]. XRD results are
supported by the SEM micrographs, Fig. 3. The uncharged sample, 3.1.2. Austenitic stainless steel
Fig. 3a and b, presents two phases with elongated grains and an In order to compare the microstructural changes of hydrogen on
average grain size of ~10 mm. Supported by energy dispersive austenitic stainless steel (AUSS), an XRD analysis was performed on
diffraction (EDS) measurements, it was shown that the high weight AUSS with the same charging condition as DSS (72 h cathodic
percent (wt%) areas of chromium and molybdenum, also known as charging and aged for one month at RT, Fig. 4). These results were
a-phase stabilizers [2], were ascribed to a phase regions, and high compared to an uncharged sample, indicated by the black line in
wt% of nickel and nitrogen, also known as g-phase stabilizers [2], Fig. 4. The uncharged sample indicates the appearance of small
were ascribed to g-phase regions. traces of a phase, even though this phase is not seen in the SEM
Supported by the XRD pattern in Fig. 2, the 72 h hydrogen micrograph (Fig. 5a). The SEM micrograph of uncharged austenitic
charged DSS and aged for one month at RT (Fig. 3 c), showed a'- stainless steel (AUSS, Fig. 5a), shows a single austenitic fcc g-phase.
martensite phase, pointed out with a red arrow. During aging of the The AUSS consists of equiaxed grains with an average grain size of
hydrogenated specimens, the -martensite became stable, while ~35 mm. The appearance of annealing twins present in each
the aphase reections became broadened. This change can be austenitic grain are common and are usually found in annealed fcc
0
explained by the next / g / a phase's transition, as described in crystals [25].
our previous work [7,12], as well as in other works [21,22]. This The comparison between XRD spectra of DSS to AUSS, Figs. 2 and
means that part of the -martensite phase changes into a'- 4, respectively, indicates that g-phase experiences the same phe-
martensite with a body centered tetragonal structure (bct). The nomenon: peaks' broadening and a shift to smaller angles (can be
formation of a'-martensite is probably due to elastic stresses during seen at 0 h aging, the green line in Fig. 4). This behavior indicates a
hydrogen desorption [22,23]. ~2% increment in lattice parameter compared with the uncharged
The surface cracking observation of 72 h charged DSS and aged sample. In addition, the formation of -martensite phase can also be
for one month at RT, Fig. 3d, shows severe cracking. Most cracking seen. This appearance is also seen in the SEM micrographs, Fig. 5b,
seems to be intergranular inside g-phase and some of them seem to pointed out with red arrows, which can also indicate the formation
be in the interface of a-g phases, Fig. 3e. These crackings are of a'-martensite due to elastic stresses during hydrogen desorption.
probably due to the lower hydrogen diffusivity inside the austenitic The -martensite phase's reections in DSS demonstrated higher
phase, Dgz3.3  1016 m2/s [16], thus causing the hydrogen con- intensity (~70% higher than AUSS's peaks), meaning a greater sta-
centration gradient beneath the surface along with higher bility of this phase in DSS. We estimated the -martensite quantities
hydrogen concentration on the surface itself. This concentration using the Rietveld method [26,27], whose analysis is based on the
454 R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459

Fig. 4. XRD pattern of the uncharged and cathodic hydrogen charged 316L for 72 h and aged for different time intervals.

Fig. 5. SEM observations of AUSS samples (a) uncharged, (b) 72 h cathodic charged, showing the formation of needle-shaped -martensite phase after one month at RT, and (c)
surface cracking in 72 h cathodic charged AUSS.

fact that intensity of different peaks in an XRD depends on the a steel containing a larger content of -martensite phase will
concentration of this phase in the material. The Rietveld method contribute to less severe damage. The description of g-phase sta-
revealed -martensite quantities of ~7 wt% and ~35 wt% in AUSS bility can be seen in Table 2. This table shows that the phase sta-
and DSS, respectively. This difference is probably related to the bility of g-phase in AUSS is the greatest, leading to a lower
alloying elements in DSS, as we have already shown in our previous hydrogen-induced phase transition (g / ) and, therefore, a
work [7]. higher severity of surface cracking.
Surface cracking observation of 72 h hydrogen charged AUSS
and aged for one month at RT, Fig. 5c, shows severe cracking, due to Table 2
the lower hydrogen diffusivity inside the g-phase. A comparison Calculated activation energies for hydrogenated for 96 h SAF 2205.
with DSS reveals much more severe intergranular cracking in AUSS.
Sample Phase concentration (wt%) S stability factor
The reason belongs to the content of g and phases in those steels:
A lower diffusivity for hydrogen will cause a larger hydrogen con- g a /a

centration gradient inside the steel. The -martensite phase pre- SMSS 30 70 8 17
sents ~7 orders higher diffusion constant, D,Hz5  1010 m2/sec SAF 2205 (DSS) 50 50 35 24
AUSS (316L) 95 5 7 32
[16], compared with g-phase, Dg z 3.3  1016 m2/s [16]. Therefore,
R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459 455

highest -martensite phase content, or with the lowest hydrogen-


induced g-phase stability, will show the least damage.
In order to study the g-phase stability, we used the following
equation, supported by the works of Rozenak and Eliezer
[22,28,29], and others [30]. The g-phase stability of stainless steels
is given by the value of S:

S Ni 0:68Cr 0:55Mn 0:45Si 27C N ; (1.1)

where S is the austenite stability factor and Ni, Cr, Mn, Si, C and N are
the stabilizing elements in wt%.
A lower value of S is related to a greater transformation of g/,
and therefore, to a higher amount of /a'martensite phase. Ac-
cording to a 'Scheler diagram', Fig. 1, in order to stabilize a steel
with a specic phase composition, a certain amount of stabilizers
should be added to the steel, i.e. AUSS, DSS and SMSS will have
different alloying elements, Fig. 1.
The data in Table 2 presents phase quantities estimated by the
Rietveld method. It can be seen that AUSS shows the higher S value
Fig. 6. XRD pattern of the uncharged and cathodic hydrogen charged SMSS for 72 h
(22% higher than DSS and 45% higher than SMSS), therefore the
and aged for different time intervals.
phase transition to -martensite is lowere 7 wt% in AUSS compared
with 35 wt% at DSS. Generally, the transition from g to -martensite
3.1.3. Super martensitic stainless steel (SMSS) can be explained by internal stresses which are created during
The XRD diffraction pattern, Fig. 6, shows a comparison between hydrogen absorption. According to our previously published work
uncharged SMSS and 72 h cathodic charged with hydrogen and [12], as well as the work of others [22,28,29,31e33], these stresses
aged for different aging times at RT. It is clearly seen that imme- can provide a signicant driving force for g / martensite
diately after charging, indicated by green line in Fig. 5, the residual transition.
g-phase peaks shift to smaller angles. As for DSS and AUSS, the A very important conclusion from these results is that high g-
SMSS also showed additional reections which were formed by phase stability accelerates hydrogen embrittlement, which estab-
hydrogen-induced g-phase transition. The XRD spectra show lishes the intensive amount of cracks on the steel's surface.
-martensitic phase with hcp crystal structure. The Rietveld
method revealed -martensite quantities of ~8 wt% in SMSS. No 3.2. Thermal desorption spectrometry (TDS) of hydrogenated
changes were seen in the original a'-martensitic peaks. Two very stainless steels
important conclusions can be deduced. First, the g/ hydrogen-
induced phase transition will always appear in steels containing 3.2.1. Duplex stainless steel (DSS)
hydrogen and g-phase, regardless of the g-phase content in a steel. The hydrogen trapping mechanisms were examined by thermal
Second, the stability of g-phase plays an important role in the phase desorption spectrometry (TDS). An example of a 72 h charged DSS
transition; higher stability will lead to a lower amount of sample after the examination of thermal desorption measurements
-martensitic phase. is presented in Fig. 8 a. All samples were heated from RT to 450  C,
In line with the XRD results of uncharged SMSS, the micro- at three non-isothermal heating rates: 2  C/sec, 4  C/sec and 6  C/
structure of SMSS presented in the SEM micrograph, Fig. 7a, con- sec. In order to solve hydrogen trapping behavior and interaction,
sists mainly of a'- martensite with bct crystal structure and a small we used Lee and Lee's model [15,34]. This model predicts the
quantity of retained g-austenite ~30 wt%, analyzed by the Rietveld hydrogen evolution rate from trap sites; hydrogen desorption rate
method [26,27]. The martensite morphology is lath with an average from trapping sites is given by:
martensitic grain size of ~10 mm. The surface cracking of 72 h
 
hydrogen charged SMSS, can be seen in Fig. 6b. The SEM micro- dx Ea
A1  Xexp  ; (1.2)
graph indicates on that there were few micro-cracks on the surface dt RT
of the sample. Supported by the surface cracking of DSS and AUSS,
Figs. 3d and 5c, it may be concluded that the sample with the where X is the hydrogen content that escapes a trap, A is the

Fig. 7. SEM observations of SMSS: (a) uncharged, and (b) surface cracking at 72 h cathodic charged.
456 R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459

Fig. 8. Experimental results at various ramp rates for 72 h cathodic hydrogen charged DSS samples: (a) hydrogen desorption rate as a function of temperature, and (b) deter-
mination of the activation energies for the three tted hydrogen desorption peaks.

reaction rate constant, Ea is the activation energy for releasing phase. Therefore, the last irreversible trapping value 62.0 0.5 kJ/
hydrogen from its trapping site, R is the gas constant and T is the mol is probably related to a'-martensite. The role of irreversible
absolute temperature. trapping sites is very crucial in preventing the hydrogen embrit-
By referring to the shift of the peak temperature with different tlement phenomenon. This evidence was also seen in previous
heating rates (Fig. 8a), the activation energy for hydrogen desorp- work of us on DSS alloys [7,12,13]. Since trapping sites control the
tion from a certain trap can be calculated. By imposing a linear availability of hydrogen to potential trapping sites, it is very
change of temperature, 4, the activation energy of the rate deter- important that the activation energy with hydrogen will be as high
mining desorption process (Ea) is obtained [15,34]: as possible [39].

 . 
dln 4 Tc2 Ea 3.2.2. Austenitic stainless steel (AUSS)
 ; (1.3) As for the DSS, the activation energies calculations of AUSS,
d1=Tc RT
Fig. 9 and Table 3, show two rst values, 19.0 0.5 kJ/mol and
where, Tc is the peak temperature and 4 is the heating rate. 32.0 2.0 kJ/mol, which are lower than 60 kJ/mol and are classied
Thus by plotting ln(4/Tc2) versus 1/Tc,Fig. 8b, where Tc is the as reversible traps. The higher value in AUSS's trapping sites,
critical temperature for hydrogen evaluation from a trap site, this 43.0 0.5 kJ/mol, is also ascribed to reversible trapping sites. The
should yield straight lines with the slope -Ea/R (R is the gas con- reason for the change, compared with DSS, in the last value is
stant) [15]. The slope of ln4 versus 1/T was tted to experimental probably related to g-phase's stability or hydrogen-induced phase
data with a 95% condence limit on the slope and multiplied by R to transition. Lower gphase stability, such as in the case of DSS, will
obtain Ea values. result in higher hydrogen absorption content and higher trap
The calculated activation energies from tted linear curves, activation energy. In the case of AUSS, the content of hydrogen-
Fig. 8b and Table 3, reveal the rst two values to be lower than
60 kJ/mol and one value which is highere 62 0.5 kJ/mol. The rst
two values can be classied as reversible traps, due to their lower
value [9,35,36]. The last value is classied as an irreversible trap-
ping site. According to different works on steels [11,16,37], revers-
ible trapping sites can be attributed to either an elastic stress eld
of dislocation with a range of 0e20 kJ/mol, a screw's dislocation
core or grain boundaries with a range of 20e30 kJ/mol and high
angle boundaries, or austenitic-ferritic interface with a range of
40e55 kJ/mol. In accordance with some of our previously published
work [7,8,12,13,38], a trapping value close to 60 kJ/mol can be
ascribed to second phases, such as -martensite or a'-martensite

Table 3
Calculated activation energies for cathodic charged for 72 h samples: DSS, AUSS, and
SMSS.

Sample Ea [kJ/mol] Ea [kJ/mol] Ea [kJ/mol]

DSS 20.0 0.5 34.0 0.5 62.0 0.5


AUSS 19.0 0.5 32.0 2.0 43.0 0.5
Fig. 9. Activation energies for the three tted hydrogen desorption peaks of 72 h
SMSS 21.0 0.5 32.0 0.5 58.0 0.5
cathodic hydrogen charged AUSS.
R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459 457

induced -martensite phase is 80% lower compared with DSS. This


difference explains the change in the last value of AUSS's trapping
energy. It can be said that phase transformation and steel's stabi-
lizers can signicantly affect steel trapping states. This data is also
supported by our previously published work [12].
The activation trapping states completely support the hydrogen
embrittlement mechanism, Figs. 3d, 5c, and 7b. It can be seen from
Table 3, that the steel with the lower trapping states levels, such as
AUSS, will mostly promote cracking. This phenomenon can be
explained as follows: At lower trapping states levels, hydrogen will
be easily available to potential cracking sites, establishing the
fracture and pathway.

3.2.3. Super martensitic stainless steel (SMSS)


The activation measurements of 72 h cathodic hydrogen
charged SMSS is presented in Fig. 10 and Table 3. These calculations
predicted values of reversible trapping sites, and one trapping site
with a close value to 60 kJ/mol. These values were 21.0 0.5 kJ/mol,
Fig. 10. Activation energies for the three tted hydrogen desorption peaks of 72 h 32.0 0.5 kJ/mol and 58.0 0.5 kJ/mol. It can be seen that the great
cathodic hydrogen charged SMSS. differences between AUSS and SMSS are in the last trapping site's
value. The SMSS presents a 25% larger value compared with AUSS.
These differences can be explained with s stability factor, Table 2,
which indicates ~47% higher values of s-stability factor in AUSS
compared with SMSS. Changes can also be seen between DSS and
SMSS, both samples presented high activation value ascribed to
martensite phase, and small differences (~6%) were seen between
trapping states. These differences can be explained by difference in
the amount of -martensite phase, and its distribution. This phase
probably had a higher effect on trapping state than a'- martensite.
Even though, a'-martensite content was signicantly higher in
SMSS ~70 wt%, its -martensite content was signicantly lower 8 wt
% compared with 35 wt% DSS. Supported by the above indication
SMSS with the lowest stability of g-phase and among the highest
trapping states' levels will show less severe cracking.
The results in Table 3 are well manifested by the changes applied
in the TDS spectra at 6  C/min of cathodic charged for 72 h SMSS,
DSS and AUSS samples.
It can be seen that each peak position, Fig. 11, is in line with its
suitable activation energy, Table 3; higher temperature's peak
levels indicate on higher activation energy. Moreover, the
Fig. 11. (a) TDS spectra at a 6  C/min heating rate of 72 h cathodic hydrogen charged
-martensite phase trapping site appears only in SMSS and DSS
samples: SMSS, DSS, and AUSS. (b) Integration of TDS spectra (a) resulting in the where the g-phase stability is lower and the amount of martensite
hydrogen desorbed content for each sample. is higher. An additional interesting observation is the spectra in-
tensity, which indicates desorbed hydrogen content, Fig. 11b. It can

Fig. 12. Inuence of heating rate on hydrogen evaluation content vs. temperature in cathodic hydrogen charged for 72 h samples: (a) DSS, (b) AUSS, and (c) SMSS.
458 R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459

be clearly seen that -martensite phase signicantly affects the severe damage and promote the fracture mechanism, due to a
amount of hydrogen absorption/desorption, due to its higher lower diffusion constant in that phase. Regarding these insights, it
hydrogen diffusion constant. SMSS presents the highest desorbed can be said that susceptibility to the hydrogen fracture mechanism
hydrogen content, ~50% higher than DSS and ~90% higher than will be the highest in AUSS and lowest in SMSS, where g-phase
AUSS, due to higher martensite content. stability is lower and trapping states are higher. Moreover, it can be
In order to better understand the desorption process and trap- concluded that the main mechanism affecting the hydrogen frac-
ping mechanisms, the inuence of heating rate on hydrogen eval- ture mechanism is g-phase stability rather than trapping states.
uation content in DSS, AUSS, and SMSS are presented in Fig. 12.
Signicant changes can be seen between DSS and AUSS steels. References
Fig. 12a and b, respectively. These changes are related to the
appearance of -martensite as a trapping site at DSS. The DSS [1] J.R. Davis, Stainless Steels, third ed., Ohaio: ASM international, 1999.
presents a certain jump at ~200  C, whereas the AUSS does not [2] P. Lacombe, B. Baroux, G. Beranger, Stainless steels, Phys. Les Ulis (1993)
16e55.
show any jump. The reason is probably due to the -martensite [3] I.M. Robertson, P. Sofronis, A. Nagao, M.L. Martin, S. Wang, D.W. Gross,
trapping site. Since AUSS does not show any sign for an K.E. Nygren, Hydrogen embrittlement understood, Metall. Mater. Trans. B 46
-martensite trapping site, the jump, which is related to hydrogen (2015) 1085e1103.
[4] R.K. Dayal, N. Parvathavarthini, Hydrogen embrittlement in power plant
desorption from the -martensite trap site, does not appear. steels, in: R. Baldev, R. Bhanu Sankara (Eds.), Sadhana Academy Proceedings in
Changes related to heating rates are the same in all steels. The Engineering Sciences, vol. 28, Universities press, 2003, pp. 431e451.
2  C/min heating rate presents the lowest desorption rate while the [5] R.A. Oriani, A decohesion theory for hydrogen-induced crack propagation, in:
Stress Corrosion Cracking and Hydrogen Embrittlement of Iron Base Alloys,
6  C/min presents the highest desorption rate, Fig. 12a and b. These 1977, pp. 351e358.
results are in line with the TDS spectra of DSS (Fig. 8a). [6] V. Olden, C. Thaulow, R. Johnsen, Modelling of hydrogen diffusion and
Regarding SMSS, Fig. 12c, the big difference between DSS and hydrogen induced cracking in supermartensitic and duplex stainless steels,
Mater. Des. 29 (2008) 1934e1948.
SMSS is related to the amount of martensite content compared with [7] R. Silverstein, D. Eliezer, B. Glam, S. Eliezer, D. Moreno, Evaluation of hydrogen
this of g (70 wt % in SMSS and 35 wt % in DSS). It is known from trapping mechanisms during performance of different hydrogen fugacity in a
previously published works that hydrogen's solubility in a-phase is lean duplex stainless steel, J. Alloys Compd. 648 (2015) 601e608.
[8] R. Silverstein, D. Eliezer, Hydrogen trapping energy levels and hydrogen
signicantly higher than in g-phase [16,17]. Therefore, SMSS will diffusion at high and low strain rates (~105s1 and 107s1) in lean duplex
show greater hydrogen content evaluation. In addition, it can be stainless steel, Mater. Sci. Eng. A 674 (2016) 419e427.
seen that SMSS and AUSS show lower desorption temperatures [9] G.M. Presouyre, Trap theory of hydrogen embrittlement, Acta Metall. 28
compared with DSS; ~220  C for SMSS, ~160  C for AUSS and (1980) 895e911.
[10] A. Mcnabb, P.K. Foster, A new analysis of the diffusion of hydrogen in iron and
~300  C for DSS. These differences were ascribed to the activation ferritic steels, Trans. Metall. Soc. AIME 227 (1963) 618e627.
energy values which were the highest in DSS. Therefore, the [11] I. Maroef, D.L. Olson, M. Eberhart, G.R. Edwards, Hydrogen trapping in ferritic
following order of maximum temperature desorption can be steel weld metal, Int. Mater. Rev. 47 (2002) 191e223.
[12] R. Silverstein, D. Eliezer, Hydrogen trapping mechanism of different duplex
written as follows: DSS>SMSS>AUSS. This order is also valid for the stainless steels alloys, J. Alloys Compd. 644 (2015) 280e286.
activation energy values. [13] R. Silverstein, O. Sobol, T. Boellinghaus, W. Unger, D. Eliezer, Hydrogen
behavior in SAF 2205 duplex stainless steel, J. Alloys Compd. 659 (2016)
2689e2695.
4. Summary and conclusions [14] Y. Yagodzinskyy, O. Todoshchenko, S. Papula, H. Ha, Hydrogen solubility and
diffusion in austenitic stainless steels studied with thermal desorption spec-
Hydrogen trapping mechanisms of SAF 2205 duplex stainless troscopy, Steel Res. Int. 82 (1) (2011) 20e25.
[15] S. Lee, J. Lee, The trapping and transport phenomena of hydrogen in nickel,
steel, 316L austenitic stainless steel and super martensitic stainless Metall. Trans. A 17 (1986) 181e187.
steel were investigated using thermal desorption spectrometry, [16] A. Turnbull, R.B. Hutchings, Analysis of hydrogen atom transport in a two-
and microstructural observations. phase alloy, Mater. Sci. Eng. A 177 (1994) 161e171.
[17] A. Turnbull, E. Beylegaard, R. Hutchings, Hydrogen transport in SAF2205 and
In all of the above mentioned steels, the dominant phase
0 SAF2507 duplex stainless steels, in: Hydrogen Transport and Cracking in
transformation was the hydrogen induced g / / a or g / . Metals, 1994, pp. 268e279.
These indications were supported by XRD and SEM results. We [18] R. Silverstein, D. Eliezer, B. Glam, D. Moreno, S. Eliezer, Inuence of hydrogen
concluded that the next transformation g / will always appear in on microstructure and dynamic strength of lean duplex stainless steel,
J. Mater. Sci. 49 (2014) 4025e4031.
steels containing hydrogen and g-phase, regardless of the g-phase [19] R. Bar, E. Dabah, D. Eliezer, T. Kannengiesser, T. Boellinghaus, The inuence of
content in the steel. hydrogen on thermal desorption processes in structural materials, Procedia
It was shown using Lee and Lee's model that trapping sites were Eng. 10 (2011) 3668e3676.
[20] R. Silverstein, D. Eliezer, B. Glam, D. Moreno, Dynamic strength of duplex steel
attributed to stress dislocation elds at 0e20 kJ/mol and to higher in the presence of hydrogen, in: SteelyHydrogen2014 Conference Pro-
activation energies at 30e40 kJ/mol to grain boundaries. The DSS ceedings, 2014, pp. 662e666.
and SMSS presented additional trapping site close to 60 kJ/mol at [21] S. Teus, V. Shyvanyuk, V. Gavriljuk, Hydrogen-induced g/3 transformation
and the role of 3-martensite in hydrogen embrittlement of austenitic steels,
~58 kJ/mol for DSS and ~62 kJ/mol for SMSS, which was related to Mater. Sci. Eng. A 497 (1e2) (Dec. 2008) 290e294.
the massive formation of ~35 %wt and ~70 %wt -martensite phase [22] P. Rozenak, L. Zevin, D. Eliezer, Hydrogen effects on phase transformations in
contents in DSS and SMSS, respectively. In this paper, we have austenitic stainless steels, Mater. Sci. 19 (1984) 567e573.
[23] N. Narita, C.J. Altstetter, H.K. Birnbaum, Hydrogen-related phase trans-
shown that the stability degree of g-phase signicantly affects the formations in austenitic stainless steels, Metall. Trans. A 13A (1982)
trapping mechanisms and, therefore, affects the hydrogen embrit- 1355e1365.
tlement mechanism. The hydrogen charged AUSS showed the [24] P. Rozenak, D. Eliezer, Quantitative X-ray phase analysis of sensitized type 316
stainless steel after cathodic hydrogen charging, Mater. Sci. Eng. 67 (1984)
greatest stability factor (S) among the rest of the steels: ~22% higher 1e4.
than DSS and ~45% higher than SMSS. This determination affects [25] S. Mahajan, C.S. Pande, M.A. Imam, B.B. Rath, Formation of annealing twins in
the amount of formed -martensite phase, which was 80% lower in fcc crystals, Acta Matter 45 (1997) 2633e2638.
[26] H.M. Rietveld, A prole renement method for nuclear and magnetic struc-
AUSS compared with DSS and 12% lower compared with SMSS. This
tures, J. Appl. Crystallogr. 2 (1969) 65e71.
determination is in line with hydrogen trapping behavior. It was [27] L.B. McCusker, R.B. Von Dreele, D.E. Cox, D. Loue r, P. Scardi, Rietveld rene-
concluded that the activation energy levels were the highest in DSS ment guidelines, J. Appl. Crystallogr. 32 (1999) 36e50.
and the lowest in AUSS. [28] P. Rozenak, D. Eliezer, Phase changes related to hydrogen- induced cracking in
austenitic stainless steel, Acta Met. 35 (1987) 2329e2340.
Hydrogen embrittlement was proved to be signicantly affected [29] P. Rozenak, D. Eliezer, Nature of the g and g*phases in austenitic stainless
by the g-phase stability; higher stability of g-phase will produce steels cathodically charged with hydrogen, Metall. Trans. A 19 (1988)
R. Silverstein, D. Eliezer / Journal of Alloys and Compounds 720 (2017) 451e459 459

2860e2862. [35] G.M. Pressouyre, I.M. Bernstein, A quantitative analysis of hydrogen trapping,
[30] S. Floreen, J.R. Mihalisim, High strength stainless steel by deformation in low Metall. Trans. A 9A (November) (1978) 1571e1580.
temperature, in: Advances in the Technology of Stainless Steels and Related [36] E. Barel, G. Ben Hamu, D. Eliezer, L. Wagner, The effect of heat treatment and
Alloys, American Society for Testing Materials, 1965, pp. 17e25. HCF performance on hydrogen trapping mechanism in timetal LCB alloy,
[31] P. Rozenak, D. Eliezer, Effects of ageing after cathodic charging in austenitic J. Alloy. Compd. 468 (2009) 77e86.
stainless steels, J. Mater. Sci. 19 (1984) 3873e3879. [37] E. Abramov, D. Eliezer, Trapping of hydrogen in helium-implanted metals,
[32] D. Eliezer, D.G. Chakrapani, C.J. Altstetter, E.N. Pugh, The inuence of austenite J. Mater. Sci. Lett. 7 (2) (Feb. 1988) 108e110.
stability on the hydrogen embrittlement and stress- corrosion cracking of [38] R. Silverstein, D. Eliezer, Effects of residual stresses on hydrogen trapping in
stainless steel, Metall. Trans. A 10 (1979) 935e941. duplex stainless steels, Mater. Sci. Eng. A 684 (2017) 64e70.
[33] E. Minkovitz, M. Talianker, D. Eliezer, TEM investigation of hydrogen induced [39] P. Sofronis, M. Dadfarnia, P. Novak, R. Yuan, B. Somerday, I.M. Robertson,
-hcp-martensite in 316L-type stainless steel, J. Mater. Sci. 16 (1981) R.O. Ritchie, T. Kanezaki, Y. Murakami, A combined applied mechanics/ma-
3506e3508. terials science approach toward quantifying the role of hydrogen on material
[34] A. Turnbull, R.B. Hutchings, D.H. Ferriss, Modelling of thermal desorption of degradation, in: Proceedings of the 12th International Conference on Fracture
hydrogen from metals, Mater. Sci. Eng. 238 (January) (1997) 317e328. (ICF-12), 2009, pp. 1e10.

Вам также может понравиться