Next Article in Journal
Dispatch Optimization Model for Haulage Equipment between Stopes Based on Mine Short-Term Resource Planning
Previous Article in Journal
Causes of Structural Heterogeneity in High-Strength OCTG Tubes and Minimizing Their Impact on Sulfide Stress Corrosion Cracking Resistance
Previous Article in Special Issue
In-Situ Hollow Sample Setup Design for Mechanical Characterisation of Gaseous Hydrogen Embrittlement of Pipeline Steels and Welds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effect of Hydrogen in Mixed Gases on the Mechanical Properties of Steels—Theoretical Background and Review of Test Results

1
Fraunhofer Institute for Mechanics of Materials IWM, Woehlerstrasse 11, 79108 Freiburg, Germany
2
Freiburg Materials Research Center (FMF), University of Freiburg, Stefan-Meier-Straße 21, 79104 Freiburg, Germany
*
Author to whom correspondence should be addressed.
Metals 2021, 11(11), 1847; https://doi.org/10.3390/met11111847
Submission received: 18 October 2021 / Revised: 16 November 2021 / Accepted: 16 November 2021 / Published: 17 November 2021

Abstract

:
This review summarizes the thermodynamics of hydrogen (H2) in mixed gases of nitrogen (N2), methane (CH4) and natural gas, with a special focus on hydrogen fugacity. A compilation and interpretation of literature results for mechanical properties of steels as a function of hydrogen fugacity implies that test results obtained in gas mixtures and in pure hydrogen, both at the same fugacity, are equivalent. However, this needs to be verified experimentally. Among the test methods reviewed here, fatigue crack growth testing is the most sensitive method to measure hydrogen effects in pipeline steels followed by fracture toughness testing and tensile testing.

1. Introduction

Significant efforts are ongoing in the European Union and worldwide to defossilize private and industry sectors by reducing the emission of green house gases. In the energy sector, one important goal is to replace natural gas (NG) with hydrogen. This goal shall be reached stepwise by blending natural gas with increasing amounts of hydrogen (H2). Worldwide, significant research activities are dedicated to investigate, whether the existing natural gas storage and transport infrastructure can be used for NG-H2 blends with up to 30% of hydrogen or even for 100% hydrogen gas, see e.g., [1] for a German case study. The analysis of the material compatibility is one important aspect.
It is well established that the mechanical properties of most metallic alloys including steels deteriorate under the influence of hydrogen. This effect is often referred to as ‘hydrogen embrittlement’. It is further known that the deterioration of mechanical properties of steels increases with increasing hydrogen concentration inside the steel, which increases with increasing hydrogen gas pressure, more precisely hydrogen fugacity [2]. The fugacity is often described as the activity of the real gas, i.e., the gas in states that are not well described by the ideal gas law. It has been found that the fugacity of pure hydrogen is different from the fugacity of hydrogen in gas mixtures [3] and that (small) additions of oxygen even mitigate hydrogen absorption into the steel [4,5].
Several publications report the effect of hydrogen in NG-H2 gas blends upon the mechanical properties of steels, see e.g., [6,7]. In other publications, natural gas is replaced by methane (CH4) [8] or nitrogen (N2) [9] to exclude unwanted secondary effects originating from the complex chemical composition of natural gas. However, in most studies, the degradation of mechanical properties is plotted as a function of the volume fraction of H2 added to NG at a given total gas pressure. To come to more general conclusions, such studies were reviewed, hydrogen fugacities were calculated, and results were compared to tests performed in pure hydrogen gas at comparable fugacities. It will be shown that the hydrogen fugacity can be used as a single parameter to describe hydrogen effects on steels independent of testing these in a gas blend or in pure hydrogen, and that testing in gas blends is not necessary to assess hydrogen effects on materials used in NG-H2 infrastructures.

2. Fugacity of Hydrogen in Gas Mixtures

The state of a real gas is described by the van der Waals equation of state (EOS) which takes into account finite sizes of molecules (co-volume parameter (b) and attractive interactions between molecules (molecular-attraction parameter (a) as follows:
P + a V m 2 V m b = R T
with
  • P: total pressure,
  • R: universal gas constant, 8.314 J/(mol K),
  • T: absolute temperature, K,
  • Vm: molar volume.
The van der Waals EOS as well as other empirical relationships for non-ideal gases with two or more parameters, or virial expansions, are complicated to use for enginering applications because of non-linearities. For applications to hydrogen systems, the Abel-Noble EOS, which is a simplified one-parameter variant of the van der Waals EOS, setting the molecular-interaction parameter a to zero, provides a reasonably good description of hydrogen-gas data with a sufficient accuracy at relevant engineering conditions (T > 223 K and p < 200 MPa) [10].
The Estimation of the amount of hydrogen dissolved in the crystal lattice of a metal from a gas requires the knowledge of the fugacity. For hydrogen as a single-component gas, the Abel-Noble equation of state provides a sufficient prediction of it’s real gas behaviour [10]. For a more general approach see e.g., [11].
The estimation of the hydrogen fugacity in gas mixtures of multiple components is presented in [3,12,13]. In the following the relevant equations are concisely summarized. For convenience, the same nomenclature as in [12] is used here. The hydrogen fugacity fHH in a gas mixture of H2 and another gas can be calculated as [12].
f H H = x H H P e P b R T
with
  • xHH: molar fraction of H2,
  • P: total pressure of the gas mixture,
  • b: co-volume constant of the gas mixture.
The molar fraction of hydrogen as well as the compressibility factors of H2 and another gas in a mixture of two gases, (ZHH and ZGas) can be calculated as [12]:
x H H = p H H Z H H p H H Z H H + p G a s Z G a s
Z H H = 1 + p H H b H H R T   a n d   Z G a s = 1 + p G a s b G a s R T
with
  • pHH, pGas: partial pressures of H2 and the other gas, respectively,
  • bHH, bGas: co-volume constants of H2 and the other gas, respectively.
The co-volume constant of the gas mixture is defined as [13]:
b = x H H b H H + 1 x H H b G a s + x H H 1 x H H b H H b G a s 2 b H H + b G a s
Using this set of equations requires both bHH and bGas for the calculation of fHH in a gas mixture according Equation (2). Within the given temperature and pressure range for most engineering applications, the co-volume constant of hydrogen can be assumed as constant (bHH = 15.84 cm³/mol [10]) whereas the co-volume constant of the other gas (bGas) can be derived from Equation (4) as a function of pressure using experimentally measured compressibility factors.
It shall be emphasized here that the application of the Abel-Noble EOS is restricted to such gases as helium, neon or hydrogen where the kinetic interaction of the molecules can be neglected [13,14]. This is typically not the case for nitrogen (N2), methane (CH4) or natural gas (NG) [13,14]. However, it will be shown in the following that the error is acceptable using the Abel-Noble EOS to assess the fugacity of hydrogen in N2-H2 and CH4-H2 gas mixtures at relevant engineering conditions, e.g., room temperature and pressures up to 20 MPa.
Experimentally measured compressibility factors of N2, CH4 and NG are shown in Figure 1. The general trends of N2, CH4 and NG are similar. For N2, Z drops slightly below unity with increasing pressure, reaches a minimum of about 0.994 at about 6 MPa and then strongly increases at pressures higher than 10 MPa to Z values significantly higher than unity. Since the decrease of Z below unity is very slight, it can fairly be assumed (at least for engineering purposes) that N2 behaves like an ideal gas up to pressures of 10 MPa at room temperature. For CH4 (and NG), Z drops significantly below unity with increasing pressure, reaches a minimum of about 0.8 at a pressure of about 16 MPa and then slightly increases with increasing pressure. Z < 1 means that the movement of the molecules in not hindered, i.e., the attractive forces dominate, which is captured by the molecular attraction parameter a in the Van der Waals equation (Equation (1)) and the co-volume parameter b can be neglected. On the other hand, Z > 1 means that repulsive forces between molecules dominate which is captured by the co-volume parameter b and the molecular attraction parameter a can be neglected. The results for N2 and CH4 from the different references appear very consistent (Figure 1a,b), while the results for NG (Figure 1c) scatter significantly presumably due to the different compositions of the natural gas qualities investigated in the individual studies. It can be seen that in terms of compressibility, natural gas is better represented by CH4 than by N2 because methane is the main constituent of natural gas, typically more than 80 vol%. However, both, the results for N2 and CH4 can be fitted by a polynomial function of the total pressure P as follows
Z = A P 5 + B P 4 + C P 3 + D P 2 + E P + F
with the coefficients given in Table 1.
Equation (6) and the fit coefficients from Table 1 can now be used to calculate ZGas for N2 or CH4, respectively. Now, the co-volume constant bGas can be calculated according to Equation (4). The results are shown in Figure 2 together with the corresponding compressibility factors.
Using the data from Figure 2, the co-volume constant of the respective gas mixture b (N2-H2 or CH4-H2) can be calculated according to Equation (5), and finally the fugacity of hydrogen in a gas mixture fHH can be calculated according to Equation (2).
The error of this method can be assessed by comparing calculated compressibility factors of a gas mixture (Zmix) using Equation (7) with experimentally measured compressibility factors.
Z m i x = 1 + P b R T
Examples for a 75%N2–25%H2 and a 78%CH4–22%H2 gas mixture are shown in Figure 3. It can be seen that the absolute error between calculated and measured compressibility factors is less than 0.03 for the N2-H2 mixture and less than 0.08 for the CH4-H2 mixture at room temperature and the given pressure range. Such errors appear tolerable for engineering applications assessing hydrogen fugacities in mixed gases.
As an example, the evolution of fHH as a function of pHH in N2-H2 and CH4-H2 gas mixtures at a total pressure of 20 MPa is shown in Figure 4a. For N2-H2 gas mixtures, fHHpHH up to hydrogen partial pressures of about 10 MPa and fHH > pHH for higher hydrogen partial pressures with fHH up to about 22 MPa at pHH = 20 MPa. For CH4-H2 gas mixtures, fHH < pHH up to pHH of about 15 MPa. The highest deviation is at pHH = 9 MPa, where fHH is calculated as low as about 7.2 MPa. In Figure 4b, the same data is plotted as a function of the hydrogen fugacity in pure hydrogen, fH2. For N2-H2 gas mixtures, fHHfH2 and the deviation appears negligible for engineering applications. For CH4-H2 gas mixtures, fHH < fH2. The highest deviation is at fH2 = 12 MPa, where fHH is calculated as low as about 9.4 MPa.

3. Compilation and Interpretation of Literature Results for Mechanical Properties as a Function of Hydrogen Fugacity

In the following, literature results from tests performed in pure hydrogen gas as well as in gas mixtures are plotted as a function of hydrogen fugacity f. The designation “fugacity f“ is used for both, hydrogen fugacity in pure H2 and hydrogen fugacity in gas mixtures. The fugacities were calculated as described in chapter 0.
It is well known that the degradation of mechanical properties of steels tested in a gaseous hydrogen atmosphere increases with increasing hydrogen fugacity following a power law
H E I   ~   m f n
where HEI means any hydrogen embrittlement index, m is a factor and n is an exponent [2]. Typical hydrogen embrittlement indices use the ratio of the mechanical property measured in H2 and in air, in percent. In this review, the relative reduction of area (RRA = RAH2/RAair) of tensile specimens (smooth and notched), the relative notched ultimate tensile strength (UTSH2/UTSair), the relative fracture toughness (KH2/Kair), and the relative crack growth rate (da/dNH2/da/dNair) were calculated based on published experimental data. Using such indices, the degree of embrittlement increases as the index decreases except for the crack growth tests where the degree of embrittlement increases as the index increases because crack growth is accelerated in hydrogen compared to air. In the context of this study, literature results on the effect of pure gaseous hydrogen as well as gas mixtures upon the mechanical properties of API 5L X42, X52, X60, X70 and X80 grades were reviewed. The data discussed in the following focuses on X52, X70 and X80 grades where a comparatively large set of data is available to allow justified conclusions. The compilation of data from various sources revealed a large scatter so that the power law dependency described above is not always obvious. However, clear trends were observed and will be discussed despite the large scatter.

3.1. Tensile Tests

Tensile RRA of smooth specimens as a function of hydrogen fugacity for X70 and X80 steels is displayed in Figure 5a. It can be seen as a clear trend that RRA decreases with increasing hydrogen fugacity. A significant amount of results show RRA values around 100% up to a fugacity of about 1.2 MPa (dashed square in Figure 5a) whereas a single result reports a RRA value as low as about 75% at a fugacity as low as about 0.2 MPa [6] (arrow in Figure 5a). Since a large amount of data suggests negligible hydrogen effects at a low fugacity below 1.2 MPa, this review implies that tensile RRA of smooth specimens is a comparably insensitive HEI compared to other indices, as will be shown in the following.
Tensile RRA of notched specimens (range of stress intensity factors kt between 2.4 and 6.3) as a function of hydrogen fugacity for X70 and X80 steels is depicted in Figure 5b. Notched RRA decreases rapidly from 98% at 0.1 MPa to about 80% at about 0.7 MPa (dashed square in Figure 5b). However, a single result reports a RRA value as low as about 45% at a fugacity as low as about 0.07 MPa [8] (arrow in Figure 5b). That is, it appears that tensile RRA of notched specimens is more sensitive to assess hydrogen effects in pipeline steels compared to tensile RRA of smooth specimens. This trend was not found for the corresponding values of the relative notched ultimate tensile strength, UTS (Figure 5c) where no significant degradation was reported up to a fugacity of 10.6 MPa.

3.2. Fracture Toughness Tests

The relative fracture toughness (K) as a function of hydrogen fugacity for X52 and X80 steels is plotted in Figure 5d. To increase the data set, elastic-plastic fracture toughness data obtained from J-integral tests were converted to K [39,40,41,44,45] using K = (JE/(1-ν2))0.5. Also here, the relative fracture toughness decreases with increasing fugacity from nearly 100% at 0.6 MPa down to 30% at 10 MPa. Relative fracture toughness values between 45% and 50% are reported for low fugacities between 0.7 MPa and 2.0 MPa (dashed square in Figure 5d). From this review it appears that the sensitivity of fracture toughness results to hydrogen effects in pipeline steels is comparable to notched RRA results (Figure 5b) and significantly higher compared to notched UTS results (Figure 5c). The latter is surprising because a correlation between notched UTS and fracture toughness was reported for stress concentration factors kt greater than 6 [55]. However, the stress concentration factors of the specimens tested in the referenced studies was less than 6, with one exception (kt ≈ 6.3 [38]), which might be one reason for the lack of correlation between the two material properties.

3.3. Fatigue Crack Growth Tests

The relative crack growth rate as a function of hydrogen fugacity for X52, X70 and X80 steels is displayed in Figure 5e,f. For grade X52 an increase in crack growth rate by a factor of about 2 at a fugacity of 1.6 MPa was reported [36] (arrow in Figure 5e) whereas for X70/X80 grades an increase in crack growth rate by a factor of about 10 to 15 at a fugacity less than 0.5 MPa (dashed square in Figure 5f) was measured [9,44,53]. The results from grades X70 and X80 clearly indicate that the growth of an initial crack or flaw is greatly accelerated under the influence of gaseous hydrogen even at a hydrogen fugacity well below 1 MPa and it is worth to mention that no result was found which reports no increase in crack growth rate at a fugacity below 1 MPa.

3.4. General Comments

It was shown in the previous sections that all reviewed HEI follow the known trends as a function of hydrogen fugacity, i.e., all the HEI decrease with increasing fugacity except the relative crack growth rate which increases with increasing fugacity. Although this study only includes results where the test conditions were similar enough to allow a direct comparison of the results, the scatter of the reviewed data is high. Plausible reasons are the different chemical compositions and microstructures allowed within the respective steel specifications, slightly different test parameters (e.g., strain rate, frequency, R ratio) or test conditions (e.g., purity of the test gas especially oxygen residues) as well as differences in sample preparation. This clearly emphasizes the urgent necessity for the development of international test standards.
Furthermore, the results in gas mixtures (N2-H2, CH4-H2, NG-H2) and in pure H2 overlap, which indicates that hydrogen fugacity is the governing parameter for both, tests in pure hydrogen and tests in gas mixtures. It appears that the influence of the other gas (N2, CH4 or natural gas) upon hydrogen-surface interactions, i.e., transport of hydrogen to the crack tip, physical adsorption, dissociative chemical adsorption and absorption [2] is small and that their effect on the mechanical response is smaller than the scatter of the data. If this assumption is true, then test results obtained in gas mixtures and in pure hydrogen, both at the same fugacity, are equivalent. However, this conclusion must be verified since a study supporting this assumption by a direct comparison of results measured in gas mixtures and in pure hydrogen could not be found.

4. Conclusions

The aforementioned results allow the following conclusions:
  • For materials testing purposes requiring a defined atmosphere, testing in CH4-H2 mixtures is preferred compared to N2-H2 mixtures to simulate the effect of H2 additions to NG.
  • The reviewed results imply no significant difference between tests in pure H2 gas and tests in gas mixtures at the same hydrogen fugacity. This needs to be verified experimentally.
  • Among the test methods reviewed here, fatigue crack growth testing is the most sensitive method to measure hydrogen effects in pipeline steels even at a very low fugacity (less than 0.5 MPa). Fracture toughness testing appears less sensitive followed by tensile testing, especially with smooth specimens.

Author Contributions

Conceptualization, T.M. and C.E.; methodology, T.M. and C.E.; formal analysis, T.M. and C.E.; investigation, T.M.; data curation, T.M. writing—original draft preparation, T.M. and C.E.; writing—review and editing, T.M., C.E., K.W. and F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by Fraunhofer through project “H2 D–a hydrogen economy for Germany”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Christopher San Marchi from Sandia National Laboratories, Livermore, CA, USA for very helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cerniauskas, S.; Junco, A.J.C.; Grube, T.; Robinius, M.; Stolten, D. Options of natural gas pipeline reassignment for hydrogen: Cost assessment for a Germany case study. Int. J. Hydrog. Energy 2020, 45, 12095–12107. [Google Scholar] [CrossRef] [Green Version]
  2. Michler, T.; Wackermann, K.; Schweizer, F. Review and Assessment of the Effect of Hydrogen Gas Pressure on the Embrittlement of Steels in Gaseous Hydrogen Environment. Metals 2021, 11, 637. [Google Scholar] [CrossRef]
  3. Marchi, C.S.; Somerday, B.P.; Larson, R.S.; Rice, S.F. Solubility of hydrogen and its isotopes in metals from mixed gases. J. Nucl. Mater. 2008, 372, 421–425. [Google Scholar] [CrossRef]
  4. Somerday, B.P.; Sofronis, P.; Nibur, K.A.; Marchi, C.S.; Kirchheim, R. Elucidating the variables affecting accelerated fatigue crack growth of steels in hydrogen gas with low oxygen concentrations. Acta Mater. 2013, 61, 6153–6170. [Google Scholar] [CrossRef]
  5. Michler, T.; Boitsov, I.E.; Malkov, I.L.; Yukhimchuk, A.A.; Naumann, J. Assessing the effect of low oxygen concentrations in gaseous hydrogen embrittlement of DIN 1.4301 and 1.1200 steels at high gas pressures. Corros. Sci. 2012, 65, 169–177. [Google Scholar] [CrossRef]
  6. Ez-Zaki, H.; Christien, F.; Bosch, C.; Briottet, L.; Bertin, M.; Levasseur, O.; Leriverain, A. Effect of hydrogen content in natural gas blend on the mechanical properties of a L485-MB low-alloyed steel. In Proceedings of the ASME 2020 Pressure Vessels & Piping Conference, Viral Conference, 19–24 July 2020; ASME Press: Minneapolis, MN, USA, 2020; p. PVP2020–21228. [Google Scholar] [CrossRef]
  7. Shang, J.; Zheng, J.; Hua, Z.; Li, Y.; Gu, C.; Cui, T.; Meng, B. Effects of stress concentration on the mechanical properties of X70 in high-pressure hydrogen-containing gas mixtures. Int. J. Hydrog. Energy 2020, 45, 28204–28215. [Google Scholar] [CrossRef]
  8. Baek, U.; Nahm, S.; Lee, H.; Lee, Y. Mechanical Properties of X70 Steel in Gaseous Hydrogen. In Proceedings of the International Hydrogen Conference (IHC 2012): Hydrogen-Materials Interactions, Moran, WY, USA, 9–12 September 2012; pp. 219–226. [Google Scholar] [CrossRef]
  9. Meng, B.; Gu, C.; Zhang, L.; Zhou, C.; Li, X.; Zhao, Y.; Zheng, J.; Chen, X.; Han, Y. Hydrogen effects on X80 pipeline steel in high-pressure natural gas/hydrogen mixtures. Int. J. Hydrog. Energy 2017, 42, 7404–7412. [Google Scholar] [CrossRef]
  10. Marchi, C.S.; Somerday, B.P.; Robinson, S.L. Permeability, solubility and diffusivity of hydrogen isotopes in stainless steels at high gas pressures. Int. J. Hydrog. Energy 2007, 32, 100–116. [Google Scholar] [CrossRef]
  11. Sugimoto, H.; Fukai, Y. Solubility of hydrogen in metals under high hydrogen pressures: Thermodynamical calculations. Acta Met. Mater. 1992, 40, 2327–2336. [Google Scholar] [CrossRef]
  12. Marchi, C.S.; Somerday, B.P. Thermodynamics of Gaseous Hydrogen and Hydrogen Transport in Metals. In Proceedings of the MRS Spring 2008 Meeting, Session HH: “The Hydrogen Economy”, San Francisco, CA, USA, 24–28 March 2008; pp. 1–12. [Google Scholar]
  13. Chenoweth, D.R. Gas-Transfer Analysis. Section H-Real Gas Results Via the Van der Waals Equation of State and Virial Expansion Extension of its Limiting Abel-Noble Form; U.S. Department of Energy, Office of Scientific and Technical Information: Washington, DC, USA, 1983; p. 87. [Google Scholar]
  14. Chenoweth, D.R.; Paolucci, S. Compressible flow of a two-phase fluid between finite vessels-II. Int. J. Multiph. Flow. 1992, 18, 669–689. [Google Scholar] [CrossRef]
  15. Michels, A.; Wouters, H.; de Boer, J. Isotherms of nitrogen between 0° and 150° and at pressures from 20 to 80 atm. Physica 1934, 1, 587–594. [Google Scholar] [CrossRef]
  16. Michels, A.; Lunbeck, R.J.; Wolkers, G.J. Thermodynamical properties of nitrogen as functions of density and temperature between −125° and +150 °C and densities up to 760 Amagat. Physica 1951, 17, 801–816. [Google Scholar] [CrossRef]
  17. Isotherms of hydrogen, of nitrogen, and of hydrogen-nitrogen mixtures, at 0° and 20 °C, up to a pressure of 200 atmospheres. Proc. R. Soc. Lond. Ser. A 1926, 111, 552–576. [CrossRef] [Green Version]
  18. Trappeniers, N.J.; Wassenaar, T.; Abels, J.C. Isotherms and thermodynamic properties of methane at temperatures between 0° and 150°C and at densities up to 570 amagat. Phys. A 1979, 98, 289–297. [Google Scholar] [CrossRef]
  19. Kvalnes, H.M.; Gaddy, V.L. The compressibility isotherms of methane at pressures to 1000 atmospheres and at temperatures from −70 to 200°. J. Am. Chem. Soc. 1931, 53, 394–399. [Google Scholar] [CrossRef]
  20. Schley, P.; Jaeschke, M.; Küchenmeister, C.; Vogel, E. Viscosity measurements and predictions for natural gas. Int. J. Thermophys. 2004, 25, 1623–1652. [Google Scholar] [CrossRef]
  21. Mihara, S.; Sagara, H.; Arai, Y.; Saito, S. The Compressibility Factors of Hydrogen Methane, Hydrogen Ethane and Hydrogen Propane Gaseous Mixtures. J. Chem. Eng. 1977, 10, 395–399. [Google Scholar] [CrossRef] [Green Version]
  22. Azizi, N.; Behbahani, R.; Isazadeh, M.A. An efficient correlation for calculating compressibility factor of natural gases. J. Nat. Gas Chem. 2010, 19, 642–645. [Google Scholar] [CrossRef]
  23. Čapla, L.; Buryan, P.; Jedelský, J.; Rottner, M.; Linek, J. Isothermal pVT measurements on gas hydrocarbon mixtures using a vibrating-tube apparatus. J. Chem. Thermodyn. 2002, 34, 657–667. [Google Scholar] [CrossRef]
  24. Zhou, J.; Patil, P.; Ejaz, S.; Atilhan, M.; Holste, J.C.; Hall, K.R. (p, Vm, T) and phase equilibrium measurements for a natural gas-like mixture using an automated isochoric apparatus. J. Chem. Thermodyn. 2006, 38, 1489–1494. [Google Scholar] [CrossRef]
  25. Langelandsvik, L.I.; Solvang, S.; Rousselet, M.; Metaxa, I.N.; Assael, M.J. Dynamic viscosity measurements of three natural gas mixtures-comparison against prediction models. Int. J. Thermophys. 2007, 28, 1120–1130. [Google Scholar] [CrossRef]
  26. Duncan, A.; Lam, P.-S.; Adams, T. Tensile testing of carbon steel in high pressure hydrogen. In Proceedings of the ASME 2007 Pressure Vessels and Piping Conference, San Antonio, TX, USA, 22–26 July 2007; p. PVP2007–PVP26736. [Google Scholar]
  27. Kussmaul, K.; Deimel, P.; Sattler, E. Tensile properties of the steel X70TM in high pressure hydrogen gas with admixtures of oxygen at different strain rates. In Proceedings of the 10th World Energy Conference Cocoa Hydrogen Energy Progress X, Boca Beach, FL, USA, 20–24 June 1994; pp. 285–293. [Google Scholar]
  28. Michler, T.; Naumann, J. Microstructural aspects upon hydrogen environment embrittlement of various bcc steels. Int. J. Hydrog. Energy 2010, 35, 821–832. [Google Scholar] [CrossRef]
  29. Bae, D.S.; Sung, C.E.; Bang, H.J.; Lee, S.P.; Lee, J.K.; Son, I.S.; Cho, Y.R.; Baek, U.B.; Nahm, S.H. Effect of highly pressurized hydrogen gas charging on the hydrogen embrittlement of API X70 steel. Met. Mater. Int. 2014, 20, 653–658. [Google Scholar] [CrossRef]
  30. Hejazi, D.; Calka, A.; Dunne, D.; Pereloma, E. Effect of gaseous hydrogen charging on the tensile properties of standard and medium Mn X70 pipeline steels. Mater. Sci. Technol. 2016, 32, 675–683. [Google Scholar] [CrossRef] [Green Version]
  31. Baek, U.B.; Lee, H.M.; Baek, S.W.; Nahm, S.H. Hydrogen Embrittlement for X-70 Pipeline Steel in High Pressure Hydrogen Gas. In Proceedings of the ASME 2015 Pressure Vessels and Piping Conference, Boston, MA, USA, 19–23 July 2015; ASME Press: Boston, MA, USA, 2015; p. PVP2015–PVP45475. [Google Scholar] [CrossRef]
  32. Zhou, D.; Li, T.; Huang, D.; Wu, Y.; Huang, Z.; Xiao, W.; Wang, Q.; Wang, X. The experiment study to assess the impact of hydrogen blended natural gas on the tensile properties and damage mechanism of X80 pipeline steel. Int. J. Hydrog. Energy 2021, 46, 7402–7414. [Google Scholar] [CrossRef]
  33. Moro, I.; Briottet, L.; Lemoine, P.; Andrieu, E.; Blanc, C.; Odemer, G. Hydrogen embrittlement susceptibility of a high strength steel X80. Mater. Sci. Eng. A 2010, 527, 7252–7260. [Google Scholar] [CrossRef] [Green Version]
  34. Zhang, T.; Zhao, W.; Zhao, Y.; Ouyang, K.; Deng, Q.; Wang, Y.; Jiang, W. Effects of surface oxide films on hydrogen permeation and susceptibility to embrittlement of X80 steel under hydrogen atmosphere. Int. J. Hydrog. Energy 2018, 43, 3353–3365. [Google Scholar] [CrossRef]
  35. Batisse, R.; Cuni, A.; Wastiaux, S.; Briottet, L.; Lemoine, P.; de Dinechin, G.; Chagnot, C.; Castilan, F.; Klosek, V.; Langlois, P.; et al. Investigation of X80-steel grade for hydrogen gas transmission pipelines. In Proceedings of the 2008 International Gas Research Conference, Paris, France, 8–10 October 2008; Gas Technology Institute: Paris, France, 2008; pp. 1572–1588. [Google Scholar]
  36. Van Wortel, H.; Gomes, M.; Demofoni, G.; Capelle, J.; Alliat, J.; Chatzidouros, E. Final report “Preparing for the hydrogen economy by using the existing natural gas system as a catalyst (NATURALHY) WP3.2: Transmission pipelines, 2009” Funded by the EU within the 6th framework programme.
  37. Zhou, Z.; Zhang, K.; Hong, Y.; Zhu, H.; Zhang, W.; He, Y.; Zhou, C.; Zheng, J.; Zhang, L. The dependence of hydrogen embrittlement on hydrogen transport in selective laser melted 304L stainless steel. Int. J. Hydrog. Energy 2021, 46, 16153–16163. [Google Scholar] [CrossRef]
  38. Shi, H.; Xing, Y.; Wang, X. Influence law of hydrogen content in coal gas system on hydrogen embrittlement sensitivity of X80 pipeline steel. Corros. Prot. 2018, 39, 336–339. [Google Scholar]
  39. Shang, J.; Wang, J.Z.; Chen, W.F.; Wei, H.T.; Zheng, J.Y.; Hua, Z.L.; Zhang, L.; Gu, C.H. Different effects of pure hydrogen vs. hydrogen/natural gas mixture on fracture toughness degradation of two carbon steels. Mater. Lett. 2021, 296, 129924. [Google Scholar] [CrossRef]
  40. Marchi, C.S.; Somerday, B.P.; Nibur, K.A.; Stalheim, D.G.; Boggess, T.; Jansto, S. Fracture and fatigue of commercial grade API pipeline steels in gaseous hydroge. In Proceedings of the ASME 2010 Pressure Vessels and Piping Division Conference (PVP 2010), Washington, DC, USA, 18–22 July 2010; p. PVP2010–2PVP5825. [Google Scholar]
  41. Cialone, H.J.; Holbrook, J.H. Sensitivity of Steels to Degradation in Gaseous Hydrogen. In ASTM STP 962; Raymond, E.L., Ed.; ASTM International: Philadelphia, PA, USA, 1988; pp. 134–152. [Google Scholar]
  42. Ronevich, J.A.; Marchi, C.S. Materials compatibility concerns for hydrogen blended into natural gas. In Proceedings of the ASME 2021 Pressure Vessels and Piping Conference (PVP2021), Virtual Online Conference, 12–16 July 2021; ASME Press: New York, NY, USA, 2021; p. PVP2021–PVP62045. [Google Scholar]
  43. Stalheim, D.; Jansto, S.G.; Boggess, T.; Ningileri, S.; Bromley, D. Microstructure and mechanical property performance evaluation of commercial grade API pipeline steels in high pressure gaseous hydrogen. In Proceedings of the 2012 International Hydrogen Conference Hydrogen-Materials Interactions, Moran, WY, USA, 9–12 September 2012; Somerday, B.P., Sofronis, P., Eds.; ASME Press: Jackson Lake Lodge, WY, USA, 2012; pp. 209–218. [Google Scholar]
  44. An, T.; Zhang, S.; Feng, M.; Luo, B.; Zheng, S.; Chen, L.; Zhang, L. Synergistic action of hydrogen gas and weld defects on fracture toughness of X80 pipeline steel. Int. J. Fatigue 2019, 120, 23–32. [Google Scholar] [CrossRef]
  45. Zhang, S.; Li, J.; An, T.; Zheng, S.; Yang, K.; Lv, L.; Xie, C.; Chen, L.; Zhang, L. Investigating the influence mechanism of hydrogen partial pressure on fracture toughness and fatigue life by in-situ hydrogen permeation. Int. J. Hydrog. Energy 2021, 46, 20621–20629. [Google Scholar] [CrossRef]
  46. Drexler, E.S.; Slifka, A.J.; Amaro, R.L.; Sowards, J.W.; Connolly, M.J.; Martin, M.L.; Lauria, D.S. Fatigue Testing of Pipeline Welds and Heat-Affected Zones in Pressurized Hydrogen Gas. J. Res. Natl. Inst. Stand. Technol. 2019, 124, 124008. [Google Scholar] [CrossRef]
  47. Drexler, E.S.; Slifka, A.J.; Amaro, R.L.; Barbosa, N.; Lauria, D.S.; Hayden, L.E.; Stalheim, D.G. Fatigue crack growth rates of API X70 pipeline steel in a pressurized hydrogen gas environment. Fatigue Fract. Eng. Mater. Struct. 2014, 37, 517–525. [Google Scholar] [CrossRef]
  48. Slifka, A.J.; Drexler, E.S.; Stalheim, D.G.; Amaro, R.L.; Lauria, D.S.; Stevenson, A.E.; Hayden, L.E. The effect of microstructure on the hydrogen-assisted fatigue of pipeline steels. In Proceedings of the ASME 2013 Pressure Vessels and Piping Conference (PVP2013), Paris, France, 14–18 July 2013; ASME Press: Paris, France, 2013; p. PVP2013–PVP97217. [Google Scholar]
  49. Drexler, E.S.; Amaro, R.L.; Slifka, A.J.; Bradley, P.E.; Lauria, D.S. Operating Hydrogen Gas Transmission Pipelines at Pressures Above 21 MPa. J. Press. Vessel Technol. 2018, 140, 61702. [Google Scholar] [CrossRef]
  50. Baek, U.B.; Nahm, S.H.; Kim, W.S.; Ronevich, J.A.; Marchi, C.S. Compatibility and suitability of existing steel pipeline for transport of hydrogen-natural gas blends. In Proceedings of the International Conference on Hydrogen Safety, Hamburg, Germany, 11–13 September 2017. [Google Scholar]
  51. Slifka, A.J.; Drexler, E.S.; Nanninga, N.E.; Levy, Y.S.; McColskey, J.D.; Amaro, R.L.; Stevenson, A.E. Fatigue crack growth of two pipeline steels in a pressurized hydrogen environment. Corros. Sci. 2014, 78, 313–321. [Google Scholar] [CrossRef]
  52. Ronevich, J.A.; Somerday, B.P.; Feng, Z. Hydrogen accelerated fatigue crack growth of friction stir welded X52 steel pipe. Int. J. Hydrog. Energy 2017, 42, 4259–4268. [Google Scholar] [CrossRef] [Green Version]
  53. Chandra, A.; Thodla, R.; Prewitt, T.J.; Matthews, W.; Sosa, S. Fatigue Crack Growth Study of X70 Line Pipe Steel in Hydrogen Containing Natural Gas Blends. In Proceedings of the ASME 2021 Pressure Vessels and Piping Conference (PVP2021), Virtual Conference, 12–16 July 2021; ASME Press: New York, NY, USA, 2021; p. PVP2021–PVP61821. [Google Scholar]
  54. Engel, V.C.; Marewski, U.; Schnotz, G.; Silcher, H.; Steiner, M.; Zickler, S. Bruchmechanische Prüfungen von Werkstoffen für Gasleitungen zur Bewertung der Wasserstofftauglichkeit: Erste Ergebnisse (In German), 3R Fachzeitschrift Für Sichere Und Effiziente Rohrleitungssysteme. Pipelintechnik 2020, 10–11, 34–41. [Google Scholar]
  55. Lee, J.A. Rapid and low cost method to determine the plane strain fracture toughness (K1C) in hydrogen. In Proceedings of the 2012 International Hydrogen Conference Hydrogen-Materials Interactions, Moran, WY, USA, 9–12 September 2012; Sofronis, P., Ed.; ASME Press: Jackson Lake Lodge, WY, USA, 2012; pp. 461–470. [Google Scholar]
Figure 1. Experimental compressibility factors of (a) N2 data from [15,16,17], (b) CH4 data from [18,19,20,21] and (c) Natural Gas data from [20,22,23,24,25].
Figure 1. Experimental compressibility factors of (a) N2 data from [15,16,17], (b) CH4 data from [18,19,20,21] and (c) Natural Gas data from [20,22,23,24,25].
Metals 11 01847 g001
Figure 2. Calculated co-volume constants and compressibility factors according to Equation (6) of (a) N2, (b) CH4.
Figure 2. Calculated co-volume constants and compressibility factors according to Equation (6) of (a) N2, (b) CH4.
Metals 11 01847 g002
Figure 3. Measured and calculated compressibility factors for (a) a 75%N2–25% H2 gas mixture [17] and (b) a 78% CH4–22% H2 gas mixture data from [21] as a function of total pressure P.
Figure 3. Measured and calculated compressibility factors for (a) a 75%N2–25% H2 gas mixture [17] and (b) a 78% CH4–22% H2 gas mixture data from [21] as a function of total pressure P.
Metals 11 01847 g003
Figure 4. Evolution of fHH in N2-H2 and CH4-H2 gas mixtures at a total pressure of 20 MPa as a function of (a) pHH and (b) fugacity of hydrogen in pure H2 gas, fH2. The dashed line represents the one-by-one ratio.
Figure 4. Evolution of fHH in N2-H2 and CH4-H2 gas mixtures at a total pressure of 20 MPa as a function of (a) pHH and (b) fugacity of hydrogen in pure H2 gas, fH2. The dashed line represents the one-by-one ratio.
Metals 11 01847 g004
Figure 5. HEI as a function of hydrogen fugacity. (a) Tensile RRA of smooth specimens for X70 and X80 steels data from [6,8,9,26,27,28,29,30,31,32,33,34,35,36]. (b) Tensile RRA of notched specimens for X70 and X80 steels data from [7,8,9,37,38]. (c) Relative notched ultimate tensile strength for X70 and X80 steels data from [7,8,9,37,38]. (d) Relative fracture toughness (K) for X52 and X80 steels data from [39,40,41,42,43,44,45]. (e) Relative crack growth rate for X52 and (f) Relative crack growth rate for X70 as well as X80 steels data from [9,36,40,42,44,46,47,48,49,50,51,52,53,54].
Figure 5. HEI as a function of hydrogen fugacity. (a) Tensile RRA of smooth specimens for X70 and X80 steels data from [6,8,9,26,27,28,29,30,31,32,33,34,35,36]. (b) Tensile RRA of notched specimens for X70 and X80 steels data from [7,8,9,37,38]. (c) Relative notched ultimate tensile strength for X70 and X80 steels data from [7,8,9,37,38]. (d) Relative fracture toughness (K) for X52 and X80 steels data from [39,40,41,42,43,44,45]. (e) Relative crack growth rate for X52 and (f) Relative crack growth rate for X70 as well as X80 steels data from [9,36,40,42,44,46,47,48,49,50,51,52,53,54].
Metals 11 01847 g005
Table 1. Fit coefficients for the calculation of the compressibility factor according to Equation (6) and coefficient of determination (R²) for N2 and CH4 according to Figure 1a,b.
Table 1. Fit coefficients for the calculation of the compressibility factor according to Equation (6) and coefficient of determination (R²) for N2 and CH4 according to Figure 1a,b.
CoefficientN2CH4
A, MPa−50−0.00000006
B, MPa−400.00000200
C, MPa−30.00000551−0.00000243
D, MPa−20.000092560.00012034
E, MPa−1−0.00133489−0.01709066
F0.999793101.00027004
0.986919550.99889577
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Michler, T.; Elsässer, C.; Wackermann, K.; Schweizer, F. Effect of Hydrogen in Mixed Gases on the Mechanical Properties of Steels—Theoretical Background and Review of Test Results. Metals 2021, 11, 1847. https://doi.org/10.3390/met11111847

AMA Style

Michler T, Elsässer C, Wackermann K, Schweizer F. Effect of Hydrogen in Mixed Gases on the Mechanical Properties of Steels—Theoretical Background and Review of Test Results. Metals. 2021; 11(11):1847. https://doi.org/10.3390/met11111847

Chicago/Turabian Style

Michler, Thorsten, Christian Elsässer, Ken Wackermann, and Frank Schweizer. 2021. "Effect of Hydrogen in Mixed Gases on the Mechanical Properties of Steels—Theoretical Background and Review of Test Results" Metals 11, no. 11: 1847. https://doi.org/10.3390/met11111847

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop