Next Article in Journal / Special Issue
Solid-State Hydrogen Storage for a Decarbonized Society
Previous Article in Journal / Special Issue
Hydrogen Sorption Properties of a Novel Refractory Ti-V-Zr-Nb-Mo High Entropy Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogen Diffusivity in Different Microstructures of 42CrMo4 Steel

University of Oviedo, Polytechnic School of Engineering of Gijón, East Building, Department of Materials Science and Metallurgical Engineering, 33203 Asturias, Spain
*
Author to whom correspondence should be addressed.
Hydrogen 2021, 2(4), 414-427; https://doi.org/10.3390/hydrogen2040023
Submission received: 10 October 2021 / Revised: 26 October 2021 / Accepted: 31 October 2021 / Published: 3 November 2021
(This article belongs to the Special Issue Feature Papers in Hydrogen)

Abstract

:
It is well known that the presence of hydrogen decreases the mechanical properties of ferritic steels, giving rise to the phenomenon known as hydrogen embrittlement (HE). The sensitivity to HE increases with the strength of the steel due to the increase of its microstructural defects (hydrogen traps), which eventually increase hydrogen solubility and decrease hydrogen diffusivity in the steel. The aim of this work is to study hydrogen diffusivity in a 42CrMo4 steel submitted to different heat treatments—annealing, normalizing and quench and tempering—to obtain different microstructures, with a broad range of hardness levels. Electrochemical hydrogen permeation tests were performed in a modified Devanathan and Stachursky double-cell. The build-up transient methodology allowed the determination of the apparent hydrogen diffusion coefficient, Dapp, and assessment of its evolution during the progressive filling of the microstructural hydrogen traps. Consequently, the lattice hydrogen diffusion coefficient, DL, was determined. Optical and scanning electron microscopy (SEM) were employed to examine the steel microstructures in order to understand their interaction with hydrogen atoms. In general, the results show that the permeation parameters are strongly related to the steel hardness, being less affected by the type of microstructure.

1. Introduction

In a global context of environmental awareness, the utilization of hydrogen as an energy source has been the subject of intensive study in the last few years [1]. Due to its versatility and zero CO2 emissions, hydrogen is expected to be a good choice for future energy systems [2,3]. Therefore, the study of economical materials that can safely store and transport hydrogen has become a topic of general interest. Specifically, medium and high-strength low-alloy steels have been used in numerous investigations in order to elucidate the influence of hydrogen on their mechanical properties [4].
Hydrogen can be absorbed by steel during fabrication (casting, electroplating, electrochemical machining, pickling, welding, etc.) and during its service (cathodic protection of offshore or buried structures, pipes, and vessels containing high pressure hydrogen gas or hydrogen containing gases, etc.) giving rise to the HE phenomenon [5]. A large number of catastrophic structural failures have been attributed to HE [6,7]. These kinds of failure occur under stress levels lower than the design stress; they are brittle in appearance—with minimum previous plastic deformation—and hence very difficult to predict. Accordingly, steel structures that work in the presence of this element must be properly designed with steels able to withstand HE. In particular, it must be borne in mind that the deterioration of mechanical properties—reduction in strength, fracture toughness and ductility, and crack growth rate enhancement [8,9]—due to the presence of internal hydrogen, increases with the strength of the steel [10].
Hydrogen diffusion and trapping in steels have been studied in order to understand the HE process. Hydrogen atoms have a high mobility in BCC steel lattices, which implies that they can easily interact with their microstructural defects, where they can be temporarily or permanently trapped [11]. Typical hydrogen traps in steels are vacancies, microvoids, dislocations, grain boundaries, matrix-carbides interfaces, and other internal interfaces [12]. These traps can be classified as reversible and irreversible depending on their ability to retain hydrogen at room temperature (RT). Irreversible traps have a high trap activation energy, therefore the hydrogen trapped in them behaves as non-diffusible at RT. In contrast, sites with lower trap activation energy constitute reversible traps, which simply delay hydrogen diffusion at RT [13]. Many studies have employed temperature-programmed desorption to characterize hydrogen trapping in steels [12,14,15,16,17].
Microstructural traps delay hydrogen diffusivity, increasing the residence time of this element in the steel and so the risk of HE. Among all the available techniques to characterize hydrogen diffusivity in steels, electrochemical hydrogen permeation has proven to be one of the most reliable methods [18,19,20,21].
For instance, Parvathavarthini et al. [22] studied hydrogen diffusion in a 2.25Cr1Mo steel submitted to a variety of heat treatments, reporting an inverse correlation between steel hardness and hydrogen diffusivity. This is in line with the findings of other authors, such as Moli-Sanchez et al. [23] and Depover et al. [15]. In general, they agree that the dislocation density reduction that takes place during certain heat treatments—high temperature tempering or annealing—is responsible for decreasing the density of hydrogen traps (mainly dislocations) and thus increasing hydrogen diffusivity. Galindo-Nava et al. [24] also demonstrated that dislocations are the main factor in controlling the mobility of hydrogen in quenched and tempered martensitic steels.
Other researchers have studied the influence of microstructure and hardness on hydrogen transport and trapping in steels [25,26,27,28]. For example, L.B. Peral et al. [29] studied the interaction between hydrogen atoms and the microstructure of quenched and tempered 2.25Cr1Mo and 2.25Cr1MoV steels, underlaying the influence of the steel hardness and of the addition of small amounts of vanadium in hydrogen diffusivity. Chan [30] studied hydrogen diffusivity and trapping in steels with different microstructures. They found that hydrogen pick up decreases with the austenite transformation temperature and, furthermore, not only grain size but also the nature of the grain boundary determines the hydrogen trapping ability and diffusivity. Thomas and Szpunar [31] observed a decrease of the diffusion coefficient and hydrogen trapping with the increase of the grain size in an X-70 ferrite-pearlite pipeline steel. Finally, according to the work of Nanninga et al. [32] on steels with different carbon contents and microstructures, hardness superseded the effects of microstructure and alloying in governing the susceptibility to HE.
Regarding the 42CrMo4 steel used in this study, although Zafra et al. [17,33] already assessed the influence of the grain size, the plastic deformation, and the tempering temperature in the hydrogen permeation parameters, other aspects, such as the influence of the tempering time on quenching and tempering treatments or the austenitization temperature in normalizing and annealing treatments, still need to be thoroughly examined.
Consequently, electrochemical hydrogen permeation tests have been performed in this work in order to study hydrogen diffusivity in a 42CrMo4 steel. Different microstructures were produced by means of annealing, normalizing, and quench and tempering heat treatments. The influence of the hardness, the austenitization temperature and the tempering time on both the apparent and the lattice hydrogen diffusion coefficients was assessed.

2. Materials and Methods

2.1. Steel and Heat Treatments

The material studied in this work was a 42CrMo4 steel. Table 1 shows the chemical composition of the steel.
In order to analyze the influence of the steel microstructure on its hydrogen permeation behavior, different heat treatments were carried out onto hot rolled 250 mm × 125 mm × 12 mm plates. The nomenclature of the obtained 42CrMo4 steel grades and their correspondent heat treatments are shown in Table 2.

2.2. Hardness and Microstructural Characterization

Small, handy specimens machined from the heat-treated plates were firstly ground with SiC papers of different grit sizes (from #60 to #1200) and then successively polished in synthetic cloths with 6 µm and 1 µm diamond paste. After completing the polishing process, samples were etched with Nital-2% and their microstructures observed using a Nikon optical microscope ECLIPSE MA200 (Material Science Laboratory, University of Oviedo, Gijón, Spain) and a JEOL scanning electron microscope JSM5600 (Scientific and Technical Services, University of Oviedo, Gijón, Spain). Vickers hardness (HV30) measurements were performed applying a load of 30 kg for 15 s.

2.3. Hydrogen Permeation Tests

Hydrogen permeation tests were performed on all the heat-treated grades shown in Table 2 in order to determine the hydrogen diffusion and trapping behavior. Specimens of 30 mm × 25 mm were ground up to #1200 SiC paper until a final thickness between 0.7 and 1 mm. They were cleaned with water and acetone before starting the test.
The permeation tests were carried out in a double electrolytic cell based on the one developed by Devanathan and Stachurski [34], which is schematically shown in Figure 1a.
Hydrogen generation occurs in the cathodic cell, where hydrogen reduction takes place (2H+ + 2e → 2Hads, Figure 1b). Hydrogen atoms are adsorbed (Hads) onto the steel surface and then absorbed into the steel (Habs). This cell was filled with 300 mL of an acid solution (pH ≈ 1) composed of 1 M H2SO4 and 0.25 g/L As2O3. In this case, H2SO4 is the hydrogen donor and As2O3 is added to limit the hydrogen recombination reaction (Hads + Hads → H2) [35], which considerably decreases the efficiency of the permeation process [33]. The anodic cell, where hydrogen is desorbed (Hdes) from the specimen and the oxidation reaction occurs (Hdes → H+ + e, Figure 1b), was filled with 300 mL of a basic solution (pH ≈ 12) of 0.1 M NaOH. The cells are separated by the specimen, which represents the working electrode (WE) in each cell. A circular exposed area of 1.25 cm2 was always used. Hydrogen oxidation was enhanced in the anodic cell, ensuring a virtually zero hydrogen concentration on the exit side of the specimen (Figure 1b) via the electrodeposition of a thin palladium layer. It also ensures that only hydrogen oxidation is taking place in the anodic cell [36]. This coating, with a thickness of approximately 50 nm, measured using SEM, was electrochemically deposited from a commercial palladium solution of 2 g/L of Pd, applying a current density of 1 mA/cm2 for 5 min [37].
Thin platinum plates with a total surface area of 1 cm2 were used as counter electrodes (CE) in each cell. A reference silver/silver chloride electrode (Ag/AgCl, RE) was also employed in the anodic cell to keep the sample at a constant anodic potential of ≈−50 mV (open-circuit potential) along all the permeation test. Before starting the tests, the background current density in the anodic cell may be below 0.1 μA/cm2, which in any case was subtracted from the measured oxidation current prior to analysis. The oxidation or permeation current density, Jp, was continuously recorded in the anodic surface of the specimen by means of a pocketSTAT Ivium potentiostat with a current operation range of ±10 mA. All tests were performed at RT.

Successive Build-Up Permeation Transients

The permeation method employed in this study consisted in sequentially increasing the applied cathodic current density, Jc, and recording the associated build-up permeation transients, as can be observed in Figure 2a. Operating this way, it is possible to determine a value of Dapp for each transient and establish the relationship between Dapp and Jc. During these tests, the microstructural traps are progressively filled with hydrogen. Once saturation is reached, the hydrogen diffusion coefficient stabilizes, which means that diffusivity is no longer affected by hydrogen trapping. This is known as the lattice diffusion coefficient of the steel, DL. Some authors have successfully applied this methodology to characterize hydrogen diffusivity in CrMo steels [33,38,39,40].
The cathodic current density was increased in steps of 0.5 mA/cm2 for the first two transients, and of 1 mA/cm2 for the following ones (see Figure 2a). As shown in Figure 2b, Dapp was calculated at each transient following the lag time method [41]. Equation (1), derived from Fick’s diffusion solution under the appropriate boundary conditions, was used for Dapp determination:
D a p p = L 2 6 · t l a g
where L is the specimen thickness, and tlag is the time needed to reach the 63% of the steady-state permeation current state, Jss. The first permeation transient performed under a cathodic current density of 0.5 mA/cm2 allows us to determine Dapp under a very low hydrogen trap occupancy condition.

3. Results

3.1. Microstructural Characterization

3.1.1. Quenched and Tempered 42CrMo4 Steel

The SEM microstructures of the six quenched and tempered 42CrMo steel grades are shown in Figure 3 under a magnification of 5000×. In addition, the HV30 (average ± standard deviation) measured on each steel grade is provided in Table 3.
The microstructure of all these 42CrMo4 grades was tempered martensite except for QT600-3min grade (Figure 3a), in which, due to the extremely short duration of the tempering treatment, it was basically untempered martensite, with a high hardness of 484 HV30. As for the rest of the treatments performed at 600 °C, it is worth noting that tempering time is inversely related to the microstructure acicularity and thus to the distortion of the martensitic structure (dislocation density) [42]. It is also observed that carbide morphology and size are strongly related to the tempering time. Elongated carbides precipitated first along grain and martensitic lath/pack/block boundaries (QT600-30min), but as the tempering time increases, these carbides break up, grow, globulize, and distribute more homogeneously (see for example QT600-7d). In line with these microstructural changes, the hardness of the steel decreases considerably with the tempering duration, as recorded in Table 3. Furthermore, the greatest effect of tempering was produced in the steel grade tempered at 725 °C for 4 h, as observed in Figure 3f. Dislocation density and internal stress levels are certainly the lowest of all the Q + T steel grades, as it presents the lowest hardness, 206 HV30.

3.1.2. Annealed and Normalized 42CrMo4 Steel

Figure 4 shows the optical and SEM micrographs of annealed (furnace cooled, FC) 42CrMo4 steel grades under different magnifications. A ferrite-pearlite microstructure is observed in both annealed grades. However, the grade annealed at lower temperature (845 °C) has a banded ferrite-pearlite microstructure, which was lost when the annealing was performed at 1050 °C. In the latter, a larger prior austenitic grain size is observed with ferrite precipitated along prior austenite grain boundaries. The slightly lower hardness measured in the steel austenitized at 845 °C (Table 4) is justified by the presence of a higher fraction of ferrite.
The SEM microstructures of the 42CrMo4 steel after the normalizing treatments (air cooled, AC) are presented in Figure 5 under 1000× and 3000×. Both grades have complex microstructures, composed by bainite with some fractions of fine pearlite and only traces of ferrite, which conferred these grades a significantly greater hardness in comparison to the annealed ones (see Table 4). As can be clearly observed at 3000× (Figure 5d) the AC-1050 grade also presents a coarser microstructure due to the higher austenitization temperature. The presence of fewer internal interfaces justifies its lower hardness values.

3.2. Hydrogen Diffusion

3.2.1. Quenched and Tempered 42CrMo4 Steel

As an example, Figure 6 and Table 5 show the results of the permeation tests performed on the QT700-7d steel.
Figure 6 presents the Jp registered on all the recorded stepped permeation transients. The different transients were produced by sequentially increasing Jc from 0.5 mA/cm2 up to a final value of 6–7 mA/cm2. Table 5 shows the values of Jc, Jss, tlag, and the Dapp value calculated in each transient by means of Equation (1).
It was observed that Jss increased proportionally with the increase of Jc, as more hydrogen was introduced in the steel microstructure. It is also worth noting that the Dapp value of the first transient was always the lowest (1.27 × 10−10 m2/s) as all hydrogen microstructural traps are initially empty. In the following transients, hydrogen was progressively retained in the traps, increasing Dapp until a maximum and approximately constant value was attained in the last transients, DL (6.19–6.13 × 10−10 m2/s). All the hydrogen microstructural traps present in the steel were saturated with hydrogen at this point [39]. These results are also in accordance with the theoretical work performed by Raina et al., which distinguished three different diffusion regimes depending on trap occupancy (traps empty, traps progressively filled, and saturated traps) [20].
The results obtained on all the Q + T grades are summarized in Figure 7, which represents the evolution of Dapp with Jc. Table 6 shows the values of Dapp, calculated in the first transient (most traps empty) and DL, obtained in the last transients (filled traps).
It is clearly appreciated that the shorter the tempering time (higher hardness), the lower the value of Dapp for any given Jc level, which means that the density of hydrogen traps increases with the hardness of the steel. DL also increases with the tempering time, as the lattice distortion is progressively reduced during tempering [42]. Zafra et al. observed the same behavior in 42CrMo4 steel when increasing the tempering temperature while keeping the treatment time constant [17].
The grade tempered at 725 °C for 4 h after quenching has the highest Dapp and DL values. This behavior indicates that this grade has the lowest lattice distortion and dislocation density and thus the lowest density of hydrogen traps. This is in line with the microstructural recovery (Figure 3f) and low hardness level (Table 3) already discussed.

3.2.2. Annealed and Normalized 42CrMo4 Steel

Similarly, the Dapp vs. Jc plot as well as the Dapp (first transient) and DL values obtained with the annealed and normalized 42CrMo4 grades are displayed in Figure 8 and Table 7, respectively. Again, Jss progressively increases with the increase of Jc, the Dapp value of the first transient was always the lowest, increasing until a maximum and approximately constant value was attained in the last transients.
Ferrite-pearlite microstructures (FC) always showed higher hydrogen diffusion coefficients than bainitic ones (AC), due to their lower hardness and thus dislocation density.
The effect of modifying the austenitizing temperature in the annealed microstructures was minimal, being the Dapp and DL values practically the same, even though the steel austenitized at 845 °C had a hardness slightly lower than that austenitized at 1050 °C.
On the other hand, a great variation in the first Dapp was detected between the two normalized grades. The finer microstructure—austenitized at 845 °C—displayed a greater density of internal interfaces and a slightly higher hardness having therefore a greater number of hydrogen traps. However, this difference in diffusivity was progressively reduced as the hydrogen concentration in the microstructure (i.e., cathodic current density) increased and, eventually, a very similar DL was obtained in these two grades.

4. Discussion

Figure 9a,b respectively show the relationship between the steel hardness, HV30, the apparent hydrogen diffusion coefficient for the first transient, Dapp, and the lattice diffusion coefficient, DL. In general, a good correlation between these two hydrogen diffusion coefficients and the steel hardness is noticed.
First, it is observed that the steel microstructure (martensite with different degrees of tempering, ferrite + pearlite or bainite) barely affects Dapp and DL values. On the contrary, there exists a correlation between hardness and hydrogen diffusion coefficients of the steel. It is well known that hardness is directly related to the dislocation density present in the steel, since these linear structural defects are considered the most important microstructural traps present in martensitic CrMo steels [17,21,43]. The same situation is observed in the annealed (ferrite-pearlite) grades, although only two steels of this family were studied. Even though the steel annealed from 845 °C shows a fine and banded ferrite-pearlite microstructure, which was lost when the annealing was performed at 1050 °C (coarse microstructure with small fractions of ferrite precipitated along prior austenite grain boundaries), both grades display very similar Dapp and DL values. It should be mentioned here that Lee and Chan [44] reported similar hydrogen diffusivities (0.65, 1.59, and 1.68 × 10−10 m2/s) for the through-thickness, longitudinal (rolling) and traverse directions of a 30CrMo4 steel with a banded ferrite-pearlite microstructure (achieved by heating to 870 °C for 1h and then furnace cooling). Therefore, the influence of ferrite-pearlite alignment in FC845 steel can be considered negligible in this work.
It is also interesting to notice that the DL obtained in the softest—more recovered—microstructures (QT725-4h, FC845 and FC1050), which lay between 1.01 and 1.16 × 10−9 m2/s, is not far from the diffusivity reported by other authors for pure iron [11,45]. It is clear then that trapping phenomena is minimum in this situation, with lattice diffusion being predominant.
On the other hand, normalized steels with bainitic microstructures do not seem to follow exactly the same trend. In this case Dapp values are certainly affected by the steel microstructure: the coarsest bainite microstructure (AC1050), austenitized at a higher temperature, provides a considerably larger Dapp value (more than 2.4 times). Assuming that hydrogen atoms mainly diffuse through the bainitic-ferrite phase [46,47], a finer bainitic microstructure with higher density of internal interfaces gives rise to more tortuous diffusion paths, in which translates into lower hydrogen diffusion coefficients. However, and regardless of their differences in hardness, the DL of both bainitic microstructures is identical. When most microstructural traps are filled, the possible differences in hydrogen diffusivity may be mainly attributed to internal structural distortion and residual stresses. It is then evident that after air cooling, both characteristics attain similar values in the two bainitic microstructures. In fact, when compared to the quenched and tempered microstructures, which were cooled very fast in water after austenitizing, the normalized steels have slightly higher DL values for similar hardness levels. Air-cooled bainitic microstructures, obtained under a relatively low cooling rate, are more relaxed and less distorted than the corresponding tempered martensite ones.

5. Conclusions

The present work studies the influence of the steel microstructure on hydrogen permeation by means of partial build-up permeation transient tests. Annealing, normalizing, and quench and tempering heat treatments were applied to hot rolled plates of a 42CrMo4 steel in order to obtain ferrite-pearlite, bainite, and tempered martensite microstructures, respectively.
The apparent hydrogen diffusion coefficient, Dapp, associated with the first permeation transient was always the lowest in all the studied grades as most hydrogen microstructural traps were initially empty. As hydrogen atoms permeated the steel, Dapp increased due to the filling of the microstructural traps. A maximum and constant characteristic Dapp value, known as the lattice diffusion coefficient DL, was reached in the last transients, when most hydrogen traps were practically saturated with hydrogen.
Dapp and DL values are strongly influenced by the hardness of the steel, with the kind of microstructure having a much lower influence. Only in the case of bainitic microstructures did the coarsest bainite give a larger Dapp value when most hydrogen microstructural traps were empty, which was justified because the hydrogen diffusion path through the bainitic ferrite was in this case less tortuous. On the other hand, differences in the lattice hydrogen diffusion coefficient, DL, may be mainly attributed to internal structural distortion and residual stresses, as in this case most traps were filled.

Author Contributions

Conceptualization, A.Z. and J.B.; methodology, A.Z.; validation, J.B. and A.Z.; formal analysis, A.I. and A.Z.; investigation, A.I., A.Z. and V.A.; resources, J.B.; data curation, A.I. and A.Z.; writing—original draft preparation, A.I., A.Z. and V.A.; writing—review and editing, J.B. and A.Z.; visualization, A.I., A.Z. and V.A.; supervision, J.B.; project administration, J.B.; funding acquisition, J.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministry of Science, Innovation and Universities, grant number RTI2018-096070-B-C31 (H2steelweld). A. Zafra acknowledges the financial support of the regional government of the Principality of Asturias for the Severo Ochoa grant PA-18-PF-BP17-038. V. Arniella thanks the funding from the Spanish State Research Agency with reference RTI2018-096070-B-C31.

Data Availability Statement

Most of the data is contained within the article. The curves obtained in all the permeation tests can be shared under demand.

Acknowledgments

The authors would like to acknowledge the technical support provided by the Scientific and Technical Service of the University of Oviedo for the use of the SEM JEOLJSM5600 scanning electron microscope.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following symbols and abbreviations are used in this manuscript:
ACAir cooled
BCCBody centered cubic
CECounter electrode
DappApparent hydrogen diffusion coefficient
DLLattice hydrogen diffusion coefficient
FCFurnace cooled
HabsHydrogen absorbed in the steel
HadsHydrogen adsorbed on the cathodic surface of the specimen
HdesHydrogen desorbed from the anodic surface of the specimen
HEHydrogen embrittlement
HV30Vickers hardness
JcCathodic current density
JpOxidation or permeation current density
JssSteady-state permeation current density
LSpecimen thickness
Q + TQuench and tempered
REReference electrode
RTRoom temperature
SEMScanning electron microscope/microscopy
tlagTime lag (time needed to reach the 63% of Jss)
WEWorking electrode

References

  1. Martin, A.; Agnoletti, M.F.; Brangier, E. Users in the design of Hydrogen Energy Systems: A systematic review. Int. J. Hydrogen Energy 2020, 45, 11889–11900. [Google Scholar] [CrossRef]
  2. International Renewable Energy Agency (IRENA). Hydrogen: A Renewable Energy Perspective, 2nd ed.; Hydrogen Energy Ministerial Meeting: Tokyo, Japan, 2019; ISBN 9789292601515. [Google Scholar]
  3. Hosseini, S.E.; Wahid, M.A. Hydrogen production from renewable and sustainable energy resources: Promising green energy carrier for clean development. Renew. Sustain. Energy Rev. 2016, 57, 850–866. [Google Scholar] [CrossRef]
  4. Liu, Q.; Atrens, A. A critical review of the influence of hydrogen on the mechanical properties of medium-strength steels. Corros. Rev. 2013, 31, 85–103. [Google Scholar] [CrossRef] [Green Version]
  5. Trautmann, A.; Mori, G.; Oberndorfer, M.; Bauer, S.; Holzer, C.; Dittmann, C. Hydrogen uptake and embrittlement of carbon steels in various environments. Materials 2020, 13, 3604. [Google Scholar] [CrossRef]
  6. Woodtli, J.; Kieselbach, R. Damage due to hydrogen embrittlement and stress corrosion cracking. Eng. Fail. Anal. 2000, 7, 427–450. [Google Scholar] [CrossRef]
  7. Graça, M.L.A.; Hoo, C.Y.; Silva, O.M.M.; Lourenço, N.J. Failure analysis of a 300M steel pressure vessel. Eng. Fail. Anal. 2009, 16, 182–186. [Google Scholar] [CrossRef]
  8. Bhadeshia, H.K.D.H. Prevention of hydrogen embrittlement in steels. ISIJ Int. 2016, 56, 24–36. [Google Scholar] [CrossRef] [Green Version]
  9. Sun, Z.; Moriconi, C.; Benoit, G.; Halm, D.; Henaff, G. Fatigue crack growth under high pressure of gaseous hydrogen in a 15-5PH martensitic stainless steel: Influence of pressure and loading frequency. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2013, 44, 1320–1330. [Google Scholar] [CrossRef]
  10. Ohaeri, E.; Eduok, U.; Szpunar, J. Hydrogen related degradation in pipeline steel: A review. Int. J. Hydrogen Energy 2018, 43, 14584–14617. [Google Scholar] [CrossRef]
  11. Kumnick, A.J.; Johnson, H.H. Hydrogen Transport Through Annealed and Deformed Armco Iron. Met. Trans 1974, 5, 1199–1206. [Google Scholar] [CrossRef]
  12. Wei, F.G.; Tsuzaki, K. Hydrogen trapping phenomena in martensitic steels. In Gaseous Hydrogen Embrittlement of Materials in Energy Technologies; Woodhead Publishing: Sawston, UK, 2012; pp. 493–525. [Google Scholar] [CrossRef]
  13. Oriani, R.A. The diffusion and trapping of hydrogen in steel. Acta Metall. 1970, 18, 147–157. [Google Scholar] [CrossRef]
  14. Pérez Escobar, D.; Depover, T.; Wallaert, E.; Duprez, L.; Verhaege, M.; Verbeken, K. Thermal desorption spectroscopy study of the interaction between hydrogen and different microstructural constituents in lab cast Fe-C alloys. Corros. Sci. 2012, 65, 199–208. [Google Scholar] [CrossRef]
  15. Depover, T.; Verbeken, K. Thermal desorption spectroscopy study of the hydrogen trapping ability of W based precipitates in a Q&T matrix. Int. J. Hydrogen Energy 2018, 43, 5760–5769. [Google Scholar] [CrossRef]
  16. Frappart, S.; Oudriss, A.; Feaugas, X.; Creus, J.; Bouhattate, J.; Thébault, F.; Delattre, L.; Marchebois, H. Hydrogen trapping in martensitic steel investigated using electrochemical permeation and thermal desorption spectroscopy. Scr. Mater. 2011, 65, 859–862. [Google Scholar] [CrossRef]
  17. Zafra, A.; Peral, L.B.; Belzunce, J. Hydrogen diffusion and trapping in A 42CrMo4 quenched and tempered steel: Influence of tempering temperature. Int. J. Hydrogen Energy 2020, 45, 31225–31242. [Google Scholar] [CrossRef]
  18. Frappart, S.; Feaugas, X.; Creus, J.; Thebault, F.; Delattre, L.; Marchebois, H. Study of the hydrogen diffusion and segregation into FeCMo martensitic HSLA steel using electrochemical permeation test. J. Phys. Chem. Solids 2010, 71, 1467–1479. [Google Scholar] [CrossRef] [Green Version]
  19. Zakroczymski, T. Electrochemical Method for Hydrogen Determination in Steel. Corrosion 1982, 38, 218–223. [Google Scholar] [CrossRef]
  20. Raina, A.; Deshpande, V.S.; Fleck, N.A. Analysis of electro-permeation of hydrogen in metallic alloys. Acta Mater. 2018, 144, 777–785. [Google Scholar] [CrossRef] [Green Version]
  21. Wei, F.G.; Tsuzaki, K. Response of hydrogen trapping capability to microstructural change in tempered Fe-0.2C martensite. Scr. Mater. 2005, 52, 467–472. [Google Scholar] [CrossRef]
  22. Parvathavarthini, N.; Saroja, S.; Dayal, R.K.; Khatak, H.S. Studies on hydrogen permeability of 2.25% Cr-1% Mo ferritic steel: Correlation with microstructure. J. Nucl. Mater. 2001, 288, 187–196. [Google Scholar] [CrossRef]
  23. Moli-Sanchez, L.; Martin, F.; Leunis, E.; Chene, J.; Wery, M. Hydrogen transport in 34CrMo4 martensitic steel: Influence of microstructural defects on H diffusion. Defect Diffus. Forum 2012, 323–325, 485–490. [Google Scholar] [CrossRef]
  24. Galindo-Nava, E.I.; Basha, B.I.Y.; Rivera-Díaz-del-Castillo, P.E.J. Hydrogen transport in metals: Integration of permeation, thermal desorption and degassing. J. Mater. Sci. Technol. 2017, 33, 1433–1447. [Google Scholar] [CrossRef]
  25. Luppo, M.I.; Ovejero-Garcia, J. The influence of microstructure on the trapping and diffusion of hydrogen in a low carbon steel. Corros. Sci. 1991, 32, 1125–1136. [Google Scholar] [CrossRef]
  26. Huang, F.; Liu, J.; Deng, Z.J.; Cheng, J.H.; Lu, Z.H.; Li, X.G. Effect of microstructure and inclusions on hydrogen induced cracking susceptibility and hydrogen trapping efficiency of X120 pipeline steel. Mater. Sci. Eng. A 2010, 527, 6997–7001. [Google Scholar] [CrossRef]
  27. Valentini, R.; Salina, A. Influence of microstructure on hydrogen embrittlement behaviour of 2·25Cr–1 Mo steel. Mater. Sci. Technol. 1994, 10, 908–914. [Google Scholar] [CrossRef]
  28. Yin, C.; Chen, J.; Ye, D.; Xu, Z.; Ge, J.; Zhou, H. Hydrogen concentration distribution in 2.25Cr-1Mo-0.25V steel under the electrochemical hydrogen charging and its influence on the mechanical properties. Materials 2020, 13. [Google Scholar] [CrossRef] [PubMed]
  29. Peral, L.B.; Zafra, A.; Fernández-Pariente, I.; Rodríguez, C.; Belzunce, J. Effect of internal hydrogen on the tensile properties of different CrMo(V) steel grades: Influence of vanadium addition on hydrogen trapping and diffusion. Int. J. Hydrogen Energy 2020, 45, 22054–22079. [Google Scholar] [CrossRef]
  30. Chan, S.L.I. Hydrogen trapping ability of steels with different microstructures. J. Chinese Inst. Eng. Trans. Chinese Inst. Eng. A/Chung-kuo K. Ch’eng Hsuch K’an 1999, 22, 43–53. [Google Scholar] [CrossRef]
  31. Thomas, A.; Szpunar, J.A. Hydrogen diffusion and trapping in X70 pipeline steel. Int. J. Hydrogen Energy 2019, 45, 2390–2404. [Google Scholar] [CrossRef]
  32. Nanninga, N.; Grochowsi, J.; Heldt, L.; Rundman, K. Role of microstructure, composition and hardness in resisting hydrogen embrittlement of fastener grade steels. Corros. Sci. 2010, 52, 1237–1246. [Google Scholar] [CrossRef]
  33. Zafra, A.; Belzunce, J.; Rodríguez, C. Hydrogen diffusion and trapping in 42CrMo4 quenched and tempered steel: Influence of quenching temperature and plastic deformation. Mater. Chem. Phys. 2020, 255, 123599. [Google Scholar] [CrossRef]
  34. Devanathan, M.A.V.; Stachurski, Z. The adsorption and diffusion of electrolytic hydrogen in palladium. Proc. R. Soc. London. Ser. A. Math. Phys. Sci. 1962, 270, 90–102. [Google Scholar] [CrossRef]
  35. Zakroczymski, T.; Szklarska-Śmiałowska, Z.; Smialowski, M. Effect of Arsenic on Permeation of Hydrogen Through Steel Membranes polarized cathodically in aqueous solution. Mater. Corros. 1975, 26, 617–624. [Google Scholar] [CrossRef]
  36. Manolatos, P.; Jerome, M.; Galland, J. Necessity of a palladium coating to ensure hydrogen oxidation during electrochemical permeation measurements on iron. Electrochim. Acta 1995, 40, 867–871. [Google Scholar] [CrossRef]
  37. Zafra, A. Study on Hydrogen Diffusivity and Embrittlement of Quenched and Tempered 42CrMo4 Steel. Ph.D. Thesis, University of Oviedo, Oviedo, Spain, 2021. [Google Scholar]
  38. Rudomilova, D.; Prošek, T.; Salvetr, P.; Knaislová, A.; Novák, P.; Kodým, R.; Schimo-Aichhorn, G.; Muhr, A.; Duchaczek, H.; Luckeneder, G. The effect of microstructure on hydrogen permeability of high strength steels. Mater. Corros. 2019, 71, 909–917. [Google Scholar] [CrossRef]
  39. Fallahmohammadi, E.; Bolzoni, F.; Fumagalli, G.; Re, G.; Benassi, G.; Lazzari, L. Hydrogen diffusion into three metallurgical microstructures of a C-Mn X65 and low alloy F22 sour service steel pipelines. Int. J. Hydrogen Energy 2014, 39, 13300–13313. [Google Scholar] [CrossRef]
  40. Zhou, P.; Li, W.; Zhao, H.; Jin, X. Role of microstructure on electrochemical hydrogen permeation properties in advanced high strength steels. Int. J. Hydrogen Energy 2018, 43, 10905–10914. [Google Scholar] [CrossRef]
  41. ASTM. Standard Practice for Evaluation of Hydrogen Uptake, Permeation, and Transport in Metals by an Electrochemical Technique; G148-97(2018); ASTM International: West Conshohocken, PA, USA, 2018. [Google Scholar]
  42. Krauss, G. Martensite in steel: Strength and structure. Mater. Sci. Eng. A 1999, 273–275, 40–57. [Google Scholar] [CrossRef]
  43. Takebayashl, S.; Kunieda, T.; Yoshinaga, N.; Ushioda, K.; Ogata, S. Comparison of the dislocation density in martensitic steels evaluated by some X-ray diffraction methods. ISIJ Int. 2010, 50, 875–882. [Google Scholar] [CrossRef] [Green Version]
  44. Lee, H.L.; Lap-Ip Chan, S. Hydrogen embrittlement of AISI 4130 steel with an alternate ferrite/pearlite banded structure. Mater. Sci. Eng. A 1991, 142, 193–201. [Google Scholar] [CrossRef]
  45. Díaz, A.; Zafra, A.; Martínez-Pañeda, E.; Alegre, J.M.; Belzunce, J.; Cuesta, I.I. Simulation of hydrogen permeation through pure iron for trapping and surface phenomena characterisation. Theor. Appl. Fract. Mech. 2020, 110, 102818. [Google Scholar] [CrossRef]
  46. Garcia, D.C.S.; Carvalho, R.N.; Lins, V.F.C.; Rezende, D.M.; Dos Santos, D.S. Influence of microstructure in the hydrogen permeation in martensitic-ferritic stainless steel. Int. J. Hydrogen Energy 2015, 40, 17102–17109. [Google Scholar] [CrossRef]
  47. Song, Y.; Han, Z.; Chai, M.; Yang, B.; Liu, Y.; Cheng, G.; Li, Y.; Ai, S. Effect of cementite on the hydrogen diffusion/trap characteristics of 2.25Cr-1Mo-0.25V steel with and without annealing. Materials 2018, 11, 788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (a) Scheme of the permeation double-cell employed in this work and (b) all the steps taking place in the hydrogen permeation process, from hydrogen reduction in the cathodic cell (left) to hydrogen oxidation in the anodic cell (right).
Figure 1. (a) Scheme of the permeation double-cell employed in this work and (b) all the steps taking place in the hydrogen permeation process, from hydrogen reduction in the cathodic cell (left) to hydrogen oxidation in the anodic cell (right).
Hydrogen 02 00023 g001
Figure 2. (a) Characteristic build-up permeation transients produced increasing the cathodic current density (Jc = 0.5 + 0.5 + 1 + 1 + 1 + 1 + 1 mA/cm2) and (b) example of Dapp calculation in a particular transient.
Figure 2. (a) Characteristic build-up permeation transients produced increasing the cathodic current density (Jc = 0.5 + 0.5 + 1 + 1 + 1 + 1 + 1 mA/cm2) and (b) example of Dapp calculation in a particular transient.
Hydrogen 02 00023 g002
Figure 3. Microstructures (SEM, 5000×) of Q + T 42CrMo4 steel grades. (a) QT300-3min, (b) QT600-30min, (c) QT600-2h, (d) QT600-24h, (e) QT600-7d, and (f) QT725-4h.
Figure 3. Microstructures (SEM, 5000×) of Q + T 42CrMo4 steel grades. (a) QT300-3min, (b) QT600-30min, (c) QT600-2h, (d) QT600-24h, (e) QT600-7d, and (f) QT725-4h.
Hydrogen 02 00023 g003
Figure 4. Microstructures of annealed 42CrMo4 steel: FC-845, (a) optical micrograph and (b) SEM micrograph; FC-1050, (c) optical micrograph and (d) SEM micrograph.
Figure 4. Microstructures of annealed 42CrMo4 steel: FC-845, (a) optical micrograph and (b) SEM micrograph; FC-1050, (c) optical micrograph and (d) SEM micrograph.
Hydrogen 02 00023 g004
Figure 5. Microstructures (SEM) of normalized 42CrMo4 steel: AC-845, (a) 1000× and (b) 3000×; AC-1050, (c) 1000× and (d) 3000×.
Figure 5. Microstructures (SEM) of normalized 42CrMo4 steel: AC-845, (a) 1000× and (b) 3000×; AC-1050, (c) 1000× and (d) 3000×.
Hydrogen 02 00023 g005
Figure 6. Evolution of the permeation current density over time for QT700-7 d grade (Jc = 0.5 + 0.5 + 1 + 1 + 1 + 1 + 1 mA/cm2).
Figure 6. Evolution of the permeation current density over time for QT700-7 d grade (Jc = 0.5 + 0.5 + 1 + 1 + 1 + 1 + 1 mA/cm2).
Hydrogen 02 00023 g006
Figure 7. Evolution of Dapp with Jc for the quenched and tempered 42CrMo4 grades.
Figure 7. Evolution of Dapp with Jc for the quenched and tempered 42CrMo4 grades.
Hydrogen 02 00023 g007
Figure 8. Evolution of Dapp with Jc for the annealed and normalized 42CrMo4 grades.
Figure 8. Evolution of Dapp with Jc for the annealed and normalized 42CrMo4 grades.
Hydrogen 02 00023 g008
Figure 9. Relationship between hydrogen diffusivity and steel hardness. (a) Dapp of the first transient vs. HV30 and (b) DL vs. HV30.
Figure 9. Relationship between hydrogen diffusivity and steel hardness. (a) Dapp of the first transient vs. HV30 and (b) DL vs. HV30.
Hydrogen 02 00023 g009
Table 1. Chemical composition (weight %) of the 42CrMo4 steel used in this study.
Table 1. Chemical composition (weight %) of the 42CrMo4 steel used in this study.
%C%Cr%Mo%Mn%Si%S%P
0.420.980.220.620.180.0020.008
Table 2. Heat treatments applied to 42CrMo4 steel and obtained grades.
Table 2. Heat treatments applied to 42CrMo4 steel and obtained grades.
Steel GradeHeat Treatment
QT600-3min845 °C/40 min + water quench + 600 °C/3 min tempering
QT600-30min845 °C/40 min + water quench + 600 °C/30 min tempering
QT600-2h845 °C/40 min + water quench + 600 °C/2 h tempering
QT600-24h845 °C/40 min + water quench + 600 °C/24 h tempering
QT600-7d845 °C/40 min + water quench + 600 °C/7 days tempering
QT725-4h845 °C/40 min + water quench + 725 °C/4 h tempering
845FC845 °C/40 min + furnace cooling
1050FC1050 °C/40 min + furnace cooling
845AC845 °C/40 min + air cooling
1050AC1050 °C/40 min + air cooing
Table 3. Chemical composition (weight %) of the 42CrMo4 steel used in this study.
Table 3. Chemical composition (weight %) of the 42CrMo4 steel used in this study.
Steel GradeMicrostructureHardness, HV30
QT600-3minUntempered martensite484 ± 3
QT600-30minTempered martensite332 ± 4
QT600-2hTempered martensite307 ± 2
QT600-24hTempered martensite280 ± 7
QT600-7dTempered martensite244 ± 5
QT725-4hTempered martensite206 ± 3
Table 4. Chemical composition (weight %) of the 42CrMo4 steel used in this study.
Table 4. Chemical composition (weight %) of the 42CrMo4 steel used in this study.
Steel GradeMicrostructureHardness, HV30
FC845Banded ferrite-pearlite183 ± 6
FC1050Ferrite-pearlite210 ± 3
AC845Bainite-ferrite-pearlite301 ± 3
AC1050Bainite-ferrite-pearlite285 ± 8
Table 5. Results obtained in the stepped permeation tests carried out on the QT700-7d grade.
Table 5. Results obtained in the stepped permeation tests carried out on the QT700-7d grade.
Transient Jc [mA/cm2]Jss [µA/cm2]tlag [s]Dapp [m2/s]
10.540.311351.27 × 10−10
21.068.24952.91 × 10−10
32.0108.03424.21 × 10−10
43.0141.42974.85 × 10−10
54.0168.32755.24 × 10−10
65.0197.42336.19 × 10−10
76.0221.72356.13 × 10−10
Table 6. Dapp (first transient) and DL of quenched and tempered 42CrMo4 grades.
Table 6. Dapp (first transient) and DL of quenched and tempered 42CrMo4 grades.
Steel GradeDapp [m2/s]DL [m2/s]
QT600-3min3.43 × 10−118.78 × 10−11
QT600-30min5.42 × 10−111.61 × 10−10
QT600-2h7.71 × 10−114.27 × 10−10
QT600-24h1.15 × 10−105.22 × 10−10
QT600-7d1.27 × 10−106.19 × 10−10
QT725-4h1.43 × 10−101.16 × 10−9
Table 7. Dapp (first transient) and DL of quenched and tempered 42CrMo4 grades.
Table 7. Dapp (first transient) and DL of quenched and tempered 42CrMo4 grades.
Steel GradeDapp [m2/s]DL [m2/s]
FC8452.24 × 10−101.15 × 10−9
FC10502.16 × 10−101.01 × 10−9
AC8455.63 × 10−116.19 × 10−10
AC10501.36 × 10−106.24 × 10−10
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Imdad, A.; Zafra, A.; Arniella, V.; Belzunce, J. Hydrogen Diffusivity in Different Microstructures of 42CrMo4 Steel. Hydrogen 2021, 2, 414-427. https://doi.org/10.3390/hydrogen2040023

AMA Style

Imdad A, Zafra A, Arniella V, Belzunce J. Hydrogen Diffusivity in Different Microstructures of 42CrMo4 Steel. Hydrogen. 2021; 2(4):414-427. https://doi.org/10.3390/hydrogen2040023

Chicago/Turabian Style

Imdad, Atif, Alfredo Zafra, Victor Arniella, and Javier Belzunce. 2021. "Hydrogen Diffusivity in Different Microstructures of 42CrMo4 Steel" Hydrogen 2, no. 4: 414-427. https://doi.org/10.3390/hydrogen2040023

Article Metrics

Back to TopTop