Skip to main content

Advertisement

Log in

Stacking Fault Energy Based Alloy Screening for Hydrogen Compatibility

  • Hydrogen Effects on Material Performance
  • Published:
JOM Aims and scope Submit manuscript

Abstract

The selection of austenitic stainless steels for hydrogen service is challenging since there are few intrinsic metrics that relate alloy composition to hydrogen degradation. One such metric, presented here, is intrinsic stacking fault energy (SFE). This work reviews the exiting literature to use estimated intrinsic SFE values, calculated with a sub-regular solution thermodynamic model, to compare the retention of tensile ductility of γ-austenitic stainless steels in the presence of hydrogen. The goal is to demonstrate SFE as a metric to screen γ-austenitic stainless steels that use diverse alloying strategies for hydrogen compatibility. A transition in the tensile reduction of area of both 300-series and manganese-stabilized stainless steels is observed at a calculated stacking fault energy of approximately 39 mJ m−2, below which pronounced hydrogen degradation on tensile ductility is observed. Calculated intrinsic stacking fault energy is demonstrated as a high-throughput screening metric for a diverse range of austenitic stainless steel compositions with regard to hydrogen compatibility.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. It is worth noting that the influence of Si on the SFE of γ-austenite stainless steels is not resolved in the literature, and several authors report non-monotonic changes in SFE with increasing Si content.27,99,100 The present values for calculation were selected based on the generally good agreement between calculated and measured SFE seen in Fig. 1 due to the limited availability of detailed comparisons between Si content and measured SFE.

References

  1. SAE International, Standard for fuel systems in fuel cell and other hydrogen vehicles (Warrendale, PA: SAE International, 2013).

    Google Scholar 

  2. T. Yamada and H. Kobayashi, J. High Press. Gas Saf. Inst. Jpn. 49, 885 (2013).

    Google Scholar 

  3. M.B. Whiteman and A.R. Troiano, Corrosion 21, 53 (1965).

    Google Scholar 

  4. R.M. Vennett and G.S. Ansell, ASM Trans. Q. 20, 242 (1967).

    Google Scholar 

  5. M.L. Holzworth, Corrosion 25, 107 (1969).

    Google Scholar 

  6. S. Matsuoka, J. Yamabe, and H. Matsunaga, Eng. Fract. Mech. 153, 103 (2016).

    Google Scholar 

  7. H.K. Birnbaum, Scr. Metall. Mater. 31, 149 (1994).

    Google Scholar 

  8. H.K. Birnbaum and P. Sofronis, Mater. Sci. Eng. A 176, 191 (1994).

    Google Scholar 

  9. I.M. Robertson, Eng. Fract. Mech. 64, 649 (1999).

    Google Scholar 

  10. D. Delafosse, in Gaseous Hydrog. Embrittlement Mater. Energy Technol., edited by R. P. Gangloff and B. P. Somerday, 1st ed. (Woodhead Publishing, Philadelphia, PA, 2012), pp. 247–285.

  11. M.L. Martin, P. Sofronis, I.M. Robertson, T. Awane, and Y. Murakami, Int. J. Fatigue 57, 28 (2013).

    Google Scholar 

  12. B.C. Odegard, J.A. Brooks, and A.J. West, in Effects of Hydrogen. Behaviour Materials, edited by A.W. Thomspon and I.M. Bernstein (The Metallurgical Society of AIME, Moran, WY, 1975), pp. 116–125

  13. G.R. Caskey Jr, Hydrogen Compatibility Handbook for Stainless Steels (Aiken: Savannah River Lab, 1983), p. 156.

    Google Scholar 

  14. S. Weber, M. Martin, and W. Theisen, Mater. Sci. Forum 706–709, 1041 (2012).

    Google Scholar 

  15. C.G. Rhodes and A.W. Thompson, Metall. Trans. A 8, 1901 (1977).

    Google Scholar 

  16. R.E. Schramm and R.P. Reed, Metall. Trans. A 6A, 1345 (1975).

    Google Scholar 

  17. P.J. Brofman and G.S. Ansell, Metall. Mater. Trans. A 9, 879 (1978).

    Google Scholar 

  18. A. Das, Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 47, 748 (2016).

    Google Scholar 

  19. L. Vitos, J.O. Nilsson, and B. Johansson, Acta Mater. 54, 3821 (2006).

    Google Scholar 

  20. L. Vitos, P.A. Korzhavyi, and B. Johansson, Phys. Rev. Lett. 96, 1 (2006).

    Google Scholar 

  21. L. Vitos, P.A. Korzhavyi, J.-O. Nilsson, and B. Johansson, Phys. Scr. 77, 65703 (2008).

    Google Scholar 

  22. P.J. Ferreira and P. Müllner, Acta Metall. 46, 4479 (1998).

    Google Scholar 

  23. L. Mosecker and A. Saeed-Akbari, Sci. Technol. Adv. Mater. 14, 033001 (2013).

    Google Scholar 

  24. A.P. Miodownik, Calphad 2, 207 (1978).

    Google Scholar 

  25. J.F. Breedis and L. Kaufman, Metall. Trans. 2, 2359 (1971).

    Google Scholar 

  26. S. Cotes, M. Sade, and A.F. Guillermet, Metall. Mater. Trans. A 26, 1957 (1995).

    Google Scholar 

  27. A. Dumay, J. Chateau, S. Allain, S. Migot, and O. Bouaziz, Mater. Sci. Eng. A 484, 184 (2008).

    Google Scholar 

  28. A. Saeed-Akbari, J. Imalu, U. Prahl, and W. Bleck, Metall. Mater. Trans. A 40A, 3076 (2009).

    Google Scholar 

  29. S. Curtze and V.-T. Kuokkala, Acta Mater. 58, 5129 (2010).

    Google Scholar 

  30. S. Curtze, V.-T. Kuokkala, A. Oikari, J. Talonen, and H. Hänninen, Acta Mater. 59, 1068 (2011).

    Google Scholar 

  31. S.T. Pisarik and D.C. Van Aken, Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 47, 1009 (2016).

    Google Scholar 

  32. P.H. Adler, G.B. Olson, and W.S. Owen, Metall. Mater. Trans. A 17, 1725 (1986).

    Google Scholar 

  33. G.B. Olson and M. Cohen, Metall. Trans. A 7, 1897 (1976).

    Google Scholar 

  34. M.J. Whalan, Proc. R. Soc. A 249, 114 (1959).

    Google Scholar 

  35. P.R. Swann, Corrosion 19, 102t (1963).

    Google Scholar 

  36. J.F. Breedis, Trans. Metall. Soc. AIME 230, 1583 (1964).

    Google Scholar 

  37. J.M. Silcock, R.W. Rookes, and J. Barford, Iron Steel Inst. J. 204, 623 (1966).

    Google Scholar 

  38. R.M. Latanision and A.W. Ruff Jr, J. Appl. Phys. 40, 2716 (1969).

    Google Scholar 

  39. L.E. Murr, Thin Solid Films 4, 389 (1969).

    Google Scholar 

  40. F. LeCroisey and B. Thomas, Phys. Status Solidi 2, K217 (1970).

    Google Scholar 

  41. R. Latanision and A. Ruff Jr, Metall. Trans. 2, 505 (1971).

    Google Scholar 

  42. C.C. Bampton, I.P. Jones, and M.H. Loretto, Acta Metall. 26, 39 (1978).

    Google Scholar 

  43. V. Gavriljuk, Y. Petrov, and B. Shanina, Scr. Mater. 55, 537 (2006).

    Google Scholar 

  44. J. Kim, S. Lee, and B.C. De Cooman, Scr. Mater. 65, 363 (2011).

    Google Scholar 

  45. J.S. Jeong, W. Woo, K.H. Oh, S.K. Kwon, and Y.M. Koo, Acta Mater. 60, 2290 (2012).

    Google Scholar 

  46. M. Kang, W. Woo, Y.-K. Lee, and B.-S. Seong, Mater. Lett. 76, 93 (2012).

    Google Scholar 

  47. T. Yonezawa, K. Suzuki, S. Ooki, and A. Hashimoto, Metall. Mater. Trans. A 44, 5884 (2013).

    Google Scholar 

  48. G.R. Lehnhoff, K.O. Findley, and B.C. De Cooman, Scr. Mater. 92, 19 (2014).

    Google Scholar 

  49. G.R. Lehnhoff and K.O. Findley, JOM 66, 756 (2014).

    Google Scholar 

  50. R.E. Stoltz and J.B. Vander Sande, Metall. Trans. A 11A, 1033 (1980).

    Google Scholar 

  51. M. Whiteman and A. Troiano, Phys. Status Solidi 109, 109 (1964).

    Google Scholar 

  52. P.J. Ferreira, I.M. Robertson, and H.K. Birnbaum, Mater. Sci. Forum 207–209, 93 (1996).

    Google Scholar 

  53. A.E. Pontini and J.D. Hermida, Scr. Mater. 37, 1831 (1997).

    Google Scholar 

  54. D. M. Bromley, Hydrogen Embrittlement Testing of Austenitic Stainless Steels SUS 316 and 316L, University of British Columbia, 2005.

  55. C. San Marchi, B.P. Somerday, X. Tang, and G.H. Schiroky, Int. J. Hydrogen Energy 33, 889 (2008).

    Google Scholar 

  56. M. Martin, S. Weber, W. Theisen, T. Michler, and J. Naumann, Int. J. Hydrogen Energy 36, 15888 (2011).

    Google Scholar 

  57. T. Michler, C. San Marchi, J. Naumann, S. Weber, and M. Martin, Int. J. Hydrogen Energy 37, 16231 (2012).

    Google Scholar 

  58. M. Martin, S. Weber, W. Theisen, T. Michler, and J. Naumann, Int. J. Hydrogen Energy 38, 5989 (2013).

    Google Scholar 

  59. T. Michler, J. Naumann, and E. Sattler, Int. J. Fatigue 51, 1 (2013).

    Google Scholar 

  60. M. Hatano, M. Fujinami, K. Arai, H. Fujii, and M. Nagumo, Acta Mater. 67, 342 (2014).

    Google Scholar 

  61. N. T. Switzner, T. Neidt, J. Hollenbeck, J. Knutson, W. Everhart, R. Hanlin, R. Bergen, D. K. Balch, and C. San Marchi, in Hydrogen interactions, edited by B. P. Somerday and P. Sofronis (ASME Press, Moran, WY, 2012), pp. 273–280

  62. C. San Marchi, T. Michler, K.A. Nibur, and B.P. Somerday, Int. J. Hydrogen Energy 35, 9736 (2010).

    Google Scholar 

  63. P. Deimel and E. Sattler, Corros. Sci. 50, 1598 (2008).

    Google Scholar 

  64. A.J. West Jr and M.R. Louthan Jr, Metall. Trans. A 13, 2049 (1982).

    Google Scholar 

  65. T. Michler, K. Berreth, J. Naumann, and E. Sattler, Int. J. Hydrogen Energy 37, 3567 (2012).

    Google Scholar 

  66. M. Koyama, E. Akiyama, T. Sawaguchi, D. Raabe, and K. Tsuzaki, Scr. Mater. 66, 459 (2012).

    Google Scholar 

  67. M. Koyama, E. Akiyama, and K. Tsuzaki, Corros. Sci. 54, 1 (2012).

    Google Scholar 

  68. T. Michler, J. Naumann, S. Weber, M. Martin, and R. Pargeter, Int. J. Hydrogen Energy 38, 9935 (2013).

    Google Scholar 

  69. S. Allain, J. Chateau, O. Bouaziz, S. Migot, and N. Guelton, Mater. Sci. Eng. A 387–389, 158 (2004).

    Google Scholar 

  70. L. Remy, Acta Metall. 25, 173 (1977).

    Google Scholar 

  71. K. Sato, M. Ichinose, Y. Hirotsu, and Y. Inoue, ISIJ Int. 29, 868 (1989).

    Google Scholar 

  72. H. Schumann, Neie Hutte 38, 647 (1967).

    Google Scholar 

  73. M. Koyama, S. Okazaki, T. Sawaguchi, and K. Tsuzaki, Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 47, 2656 (2016).

    Google Scholar 

  74. C.S. Marchi, D.K. Balch, K. Nibur, and B.P. Somerday, J. Press. Vessel Technol. 130, 041401 (2008).

    Google Scholar 

  75. P. J. Gibbs, C. San Marchi, K. A. Nibur, and X. Tang, in American Society of Mechanical Engineers, Pressure Vessels and Piping Division. PVP (Vancouver, 2016), pp. 1–8.

  76. J.A. Brooks and A.W. Thompson, Metall. Trans. A 24, 1983 (1993).

    Google Scholar 

  77. T. Michler and J. Naumann, Int. J. Hydrogen Energy 35, 821 (2010).

    Google Scholar 

  78. G. Gedge, J. Constr. Steel Res. 64, 1194 (2008).

    Google Scholar 

  79. J.W. Simmons, Mater. Sci. Eng. A 207, 159 (1996).

    Google Scholar 

  80. K.-T. Park, K.G. Jin, S.H. Han, S.W. Hwang, K. Choi, and C.S. Lee, Mater. Sci. Eng. A 527, 3651 (2010).

    Google Scholar 

  81. D. Suh, S. Park, T. Lee, C. Oh, and S. Kim, Metall. Mater. Trans. A 41, 397 (2010).

    Google Scholar 

  82. M.L. Glenn, J. Mater. Eng. 10, 181 (1988).

    Google Scholar 

  83. G.H. Eichelman and F.C. Hull, Trans. Am. Soc. Met. 45, 77 (1953).

    Google Scholar 

  84. S.S. Babu, E.D. Specht, S.A. David, E. Karapetrova, P. Zschack, M. Peet, and H.K.D.H. Bhadeshia, Metall. Mater. Trans. A 36, 3281–3289 (2005).

    Google Scholar 

  85. C.G. de Andre, F.G. Caballero, C. Capdevila, and H.K.D.H. Bhadeshia, Scripta Mater 39, 791–796 (1998).

    Google Scholar 

  86. K. Ishida, Phys. Stat. Sol. 36, 717 (1976).

    Google Scholar 

  87. M. Hillert and M. Jarl, CALPHAD 2, 227 (1978).

    Google Scholar 

  88. G. Inden, Phys. B+C 103, 82 (1981).

    Google Scholar 

  89. I.A. Yakubtsov, A. Ariapour, and D.D. Perovic, Acta Mater. 47, 1271 (1999).

    Google Scholar 

  90. Y. Lee and C. Choi, Metall. Mater. Trans. A 31, 355 (2000).

    Google Scholar 

  91. S. Takaki, H. Nakatsu, and Y. Tokunaga, Mater. Trans. JIM 34, 489 (1993).

    Google Scholar 

  92. J.-H. Jun and C.-S. Choi, Mater. Sci. Eng. A 257, 353 (1998).

    Google Scholar 

  93. T. Ericsson, Acta Metall. 14, 1037 (1966).

    Google Scholar 

  94. A.T. Dinsdale, CALPHAD 15, 317 (1991).

    Google Scholar 

  95. W.S.S. Yang, C.M. Wan, and C.M. Wam, J. Mater. Sci. 25, 1821 (1990).

    Google Scholar 

  96. O. Grässel, G. Frommeyer, C. Derder, and H. Hofmann, J. Phys. IV Fr. 7, C5 (1997).

    Google Scholar 

  97. C. Ko and R.B. McLeallan, Acta Met. 31, 1821 (1983).

    Google Scholar 

  98. W. Huang, CALPHAD 13, 243 (1989).

    Google Scholar 

  99. Q.X. Dai, X.N. Cheng, Y.T. Zhao, X.M. Luo, and Z.Z. Yuan, Mater. Charact. 52, 349 (2004).

    Google Scholar 

  100. Y.-K. Lee, J.-H. Jun, and C. Choi, Scr. Mater. 35, 825 (1996).

    Google Scholar 

Download references

Acknowledgements

The authors acknowledge J.A. Ronevich, C.D. Spataru, M.E. Foster, D.L. Medlin, and S.K. Lawrence at Sandia National Laboratories for their productive comments contributing to the construction of this manuscript and K.A. Nibur for his thought-provoking suggestions on hydrogen-metal interactions. The Office of Energy Efficiency and Renewable Energy’s Fuel Cell Technologies Office at the US Department of Energy, through the Hydrogen Storage program element, provided funding support for this work under Project ST113. Sandia National Laboratories is a multi-mission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC, a wholly owned subsidiary of Honeywell International, Inc., for the US Department of Energy’s National Nuclear Security Administration under Contract DE-NA-0003525. This paper describes objective technical results and analysis. Any subjective views or opinions that might be expressed in the paper do not necessarily represent the views of the U.S. Department of Energy or the United States Government.

Author information

Authors and Affiliations

Author notes

  1. Paul Gibbs performed this work as a Postdoctoral Appointee at Sandia National Laboratories and is now a Research and Development Engineer at Los Alamos National Laboratories.

    • P. J. Gibbs
Authors

Corresponding author

Correspondence to P. J. Gibbs.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A

Appendix A

To calculate intrinsic SFE for steel alloys of interest, the relevant thermodynamic parameters and model architectures were collected from the available literature; the works of Curtze et al.30 and Saeed-Akbari et al.28 were of great assistance in compiling the present construction. Starting with the construction in Eq. 1, a value of 8 mJ m−2 was used for the interfacial energy (σ) based on the survey of reported values by Saeed-Akbari et al.28 The molar surface density (ρ) was calculated following the work by Allain et al.:69

$$ \rho = \frac{4}{\sqrt 3 }\frac{1}{{\left( {a_{\gamma }^{T} } \right)^{2} N_{\text{A}} }} $$
(2)

where aTγ is the lattice parameter at the temperature of interest and NA is Avogadro’s number. The composition-dependent lattice parameter in angstroms at room temperature was calculated using the expression of Babu et al.:84

$$ a_{\gamma }^{300K} = 3.5780 - 0.002X_{\text{Ni}} + 0.00095X_{\text{Mn}} + 0.0006X_{\text{Cr}} + 0.0056X_{\text{Al}} + 0.0033X_{\text{C}} + 0.0031X_{\text{Mo}} $$
(3)

where Xi is the molar fraction of each of the elements (i). The change in lattice parameter of the γ-austenitic phase due to thermal expansion was considered using the equation presented by García de Andrés et al.:85

$$ a_{\gamma }^{T} = a_{\gamma }^{300K} \left[ {1 + \delta_{\gamma } \left( {T - 300} \right)} \right] $$
(4)

where δγ is the coefficient of thermal expansion for austenite of 2.065 × 10−9 K−1 and T is the temperature in Kelvin.

The driving force for the formation of the ε-martensite phase, \( \Delta G^{\gamma \to \varepsilon } \), may be estimated for a given alloy composition and temperature if sufficient thermodynamic data are available. The overall driving force may be broken into component terms using a sub-regular solution model, resulting in a summation of the individual energetic contributions to the overall transformation potential:

$$ \Delta G^{\gamma \to \varepsilon } = \Delta G^{\gamma \to \varepsilon }_{{{\text{Chem}} .}} + \Delta G^{\gamma \to \varepsilon }_{{{\text{Mag}} .}} + \Delta G^{\gamma \to \varepsilon }_{{{\text{Seg}} .}} + \Delta G^{\gamma \to \varepsilon }_{\text{GS}} + \Delta G^{\gamma \to \varepsilon }_{\text{Strain}} $$
(5)

Here, the chemical potential for the transformation (\( \Delta G^{\gamma \to \varepsilon }_{{{\text{Chem}} .}} \))33,86 is augmented by the excess driving force for the transformation because of magnetism of the respective phases (\( \Delta G^{\gamma \to \varepsilon }_{{{\text{Mag}} .}} \)),87,88 the strain induced by the change in crystallography and lattice parameter (\( \Delta G^{\gamma \to \varepsilon }_{\text{Strain}} \)),33 the contribution of segregated nitrogen at the stacking fault (\( \Delta G^{\gamma \to \varepsilon }_{{{\text{Seg}} .}} \)),89 and the change in driving force due to grain size (\( \Delta G^{\gamma \to \varepsilon }_{\text{GS}} \)).28,90,91 The contribution of \( \Delta G^{\gamma \to \varepsilon }_{\text{Strain}} \) to the overall transformation is typically small and is neglected here.33 Similarly, when the grain size is larger than approximately 30 μm, as is typical for commercial austenitic stainless steels, the excess energy due to grain size is negligible.91,92 Details for the calculation of the other parameters are summarized in the following sections.

Chemical Potential

In a sub-regular solution model the chemical potential for a transformation may be separated into two summations:

$$ \Delta G^{\gamma \to \varepsilon }_{{{\text{Chem}} .}} = \mathop \sum \limits_{i} X_{i} \Delta G^{\gamma \to \varepsilon }_{i} + \mathop \sum \limits_{ij} X_{i} X_{j} \varOmega_{ij}^{\gamma \to \varepsilon } $$
(6)

where \( \Delta G^{\gamma \to \varepsilon }_{i} \) is the contribution of each element (i) to the transformation driving force, \( \varOmega_{ij}^{\gamma \to \varepsilon } \) represents the entropic first-order interactions between elements i and j to modify the transformation potential, and Xi and Xj are the molar fractions of the constituent species. Here we consider alloy systems consisting of Fe-Ni-Mn-Cr-Al-Si-Cu; the relative terms for \( \Delta G^{\gamma \to \varepsilon }_{i} \) and \( \varOmega_{ij}^{\gamma \to \varepsilon } \) are included in Table I.

Table I Chemical contributions to \( \Delta G^{\gamma \to \varepsilon }_{{{\text{Chem}} .}} \)

The calculation of the contribution of bulk nitrogen to the γ to ε phase transition (\( \Delta G^{\gamma \to \varepsilon }_{\text{N,bulk}} \)) requires additional rigor due to the site occupancy of nitrogen in the FCC lattice. This work was initially undertaken by Yakubtsov89 with additional terms for Mn added by Curtze30 based on the work by Cotes.26 A complete description of the derivation of \( \Delta G^{\gamma \to \varepsilon }_{\text{N,bulk}} \) may be found in the sources above; the final form follows:

$$ \Delta G^{\gamma \to \varepsilon }_{\text{N,bulk}} = 6\left( {P_{1} + P_{2} + P_{3} + P_{4} - P_{5} - P_{6} - P_{7} } \right) $$
(7)

where:

$$ P_{1} = X{_{\text{Cr }}}\Delta G^{\gamma \to \varepsilon }_{\text{Cr}} $$
(8)
$$ P_{2} = \frac{{X_{\text{Fe}} \left( {U_{\text{FeN}}^{\varepsilon } - U_{\text{CrN}}^{\varepsilon } } \right)}}{{X_{\text{Fe}} + X_{\text{Cr}}\exp \left( {\frac{{ - U_{\text{FeN}}^{\varepsilon } - U_{\text{CrN}}^{\varepsilon } }}{{RT}}} \right) + X_{\text{Ni}}\exp \left( {\frac{{ - U_{\text{FeN}}^{\varepsilon } - U_{\text{NiN}}^{\varepsilon } }}{{RT}}} \right) + X_{\text{Mn}}\exp \left( {\frac{{ - U_{\text{FeN}}^{\varepsilon } - U_{\text{MnN}}^{\varepsilon } }}{{RT}}} \right)}} $$
(9)
$$ P_{3} = \frac{{{\text{X}}_{\text{Ni}}\left( {U_{\text{NiN}}^{\varepsilon } - U_{\text{CrN}}^{\varepsilon } } \right)}}{{X_{\text{Fe}}{ \exp }\left( {\frac{{ - U_{\text{NiN}}^{\varepsilon } - U_{\text{FeN}}^{\varepsilon} }}{RT}} \right) + X_{\text{Cr}}{ \exp }\left( {\frac{{ - U_{\text{NiN}}^{\varepsilon } - U_{\text{CrN}}^{\varepsilon } }}{RT}} \right) + X_{\text{Ni}} + X_{\text{Mn}}{ \exp }\left( {\frac{{ - U_{\text{NiN}}^{\varepsilon } - U_{\text{MnN}}^{\varepsilon } }}{RT}} \right)}} $$
(10)
$$ P_{4} = \frac{{X_{\text{Ni}} \left( {U_{\text{MnN}}^{\varepsilon } - U_{\text{CrN}}^{\varepsilon } } \right)}}{{X_{\text{Fe}}\exp \left( {\frac{{ - U_{\text{MnN}}^{\varepsilon } - U_{\text{FeN}}^{\varepsilon } }}{{RT}}} \right) + X_{\text{Cr}}\exp \left( {\frac{{ - U_{\text{MnN}}^{\varepsilon } - U_{\text{CrN}}^{\varepsilon } }}{{RT}}} \right) + X_{\text{Ni}}\exp \left( {\frac{{ - U_{\text{MnN}}^{\varepsilon } - U_{\text{NiN}}^{\varepsilon } }}{{RT}}} \right) + X_{\text{Mn}}}} $$
(11)
$$ P_{5} = \frac{{X_{\text{Fe }}\left( {U_{\text{FeN}}^{\gamma } - U_{\text{CrN}}^{\gamma } } \right)}}{{X_{\text{Fe}} + X_{\text{Cr}}\exp \left( {\frac{{ - U_{\text{FeN}}^{\gamma } - U_{\text{CrN}}^{\gamma } }}{{RT}}} \right) + X_{\text{Ni}}\exp \left( {\frac{{ - U_{\text{FeN}}^{\gamma } - U_{\text{NiN}}^{\gamma } }}{{RT}}} \right) + X_{\text{Mn}}\exp \left( {\frac{{ - U_{\text{FeN}}^{\gamma } - U_{\text{MnN}}^{\gamma } }}{{RT}}} \right)}} $$
(12)
$$ P_{6} = \frac{{X_{\text{Ni}} \left( {U_{\text{NiN}}^{\gamma } - U_{\text{CrN}}^{\gamma } } \right)}}{{X_{\text{Fe}}\exp \left( {\frac{{ - U_{\text{NiN}}^{\gamma } - U_{\text{FeN}}^{\gamma } }}{{RT}}} \right) + X_{\text{Cr}}\exp \left( {\frac{{ - U_{\text{NiN}}^{\gamma } - U_{\text{CrN}}^{\gamma } }}{{RT}}} \right) + X_{\text{Ni}} + X_{\text{Mn}}\exp \left( {\frac{{ - U_{\text{NiN}}^{\gamma } - U_{\text{MnN}}^{\gamma } }}{{RT}}} \right)}} $$
(13)
$$ P_{7} = \frac{{X_{\text{Mn}}\left( {U_{\text{MnN}}^{\gamma } - U_{\text{CrN}}^{\gamma } } \right)}}{{X_{\text{Fe}}\exp \left( {\frac{{ - U_{\text{MnN}}^{\gamma } - U_{\text{FeN}}^{\gamma } }}{{RT}}} \right) + X_{\text{Cr}}\exp \left( {\frac{{ - U_{\text{MnN}}^{\gamma } - U_{\text{CrN}}^{\gamma } }}{{RT}}} \right) + X_{\text{Ni}}\exp \left( {\frac{{ - U_{\text{MnN}}^{\gamma } - U_{\text{NiN}}^{\gamma } }}{{RT}}} \right) + X_{\text{Mn}}}} $$
(14)

The values for \( U_{iN}^{\gamma } - U_{jN}^{\gamma } \) are included in Table I while the values of \( U_{iN}^{\varepsilon } - U_{jN}^{\varepsilon } \) may be estimated using:

$$ \left( {U_{iN}^{\varepsilon } - U_{jN}^{\varepsilon } } \right) = \Delta G^{\gamma \to \varepsilon }_{i} + \left( {U_{iN}^{\gamma } - U_{jN}^{\gamma } } \right) $$
(15)

Segregated Nitrogen

Nitrogen, in addition to the bulk contribution to the driving force for γ to ε transformation, can also segregate to stacking fault interfaces. This elevated nitrogen content at the structural defect has a coupled influence on the overall transformation likelihood and must be addressed in addition to the bulk effect. This influence was considered in detail by Ishida86 and Yakubtsov89 and follows the form:

$$ \Delta G^{\gamma \to \varepsilon }_{\text{N,seg}} = \Delta G^{\gamma \to \varepsilon }_{{{\text{N,Chem}} - {\text{seg}}}} + \Delta G^{\gamma \to \varepsilon }_{{{\text{N,Surf}} - {\text{seg}}}} + \Delta G^{\gamma \to \varepsilon }_{{{\text{N,El}} - {\text{seg}}}} $$
(16)

where \( \Delta G^{\gamma \to \varepsilon }_{{{\text{N,Chem}} - {\text{seg}}}} \) is the chemical energetic contribution chemical energy due to Suzuki segregation,\( \Delta G^{\gamma \to \varepsilon }_{{{\text{N,Surf}} - {\text{seg}}}} \) is the surface energy contribution of N at the stacking fault boundary, and \( \Delta G^{\gamma \to \varepsilon }_{{{\text{N,El}} - {\text{seg}}}} \) is the elastic energetic component due to segregation of elements of different size. The elastic component is generally considered to be negligible compared with the chemical components of segregated N at the stacking fault. The chemical contribution may be estimated as follows:

$$ \Delta G^{\gamma \to \varepsilon }_{{N,{\text{Chem}} - {\text{seg}}}} = RT\left[ {\ln \left( {\frac{{X_{{{\text{N}}\left( {\text{seg}} \right)}} }}{{X_{{{\text{N}}\left( {\text{bulk}} \right)}} }}} \right) + \left( {1 - X_{{N\left( {\text{bulk}} \right)}} } \right)\ln \left( {\frac{{1 - X_{{N\left( {\text{seg}} \right)}} }}{{1 - X_{{{\text{N}}\left( {\text{bulk}} \right)}} }}} \right)} \right] $$
(17)

where \( X_{{N\left( {seg} \right)}} \) may be estimated by:

$$ X_{{{\text{N}}\left( {\text{seg}} \right)}} = \left[ {1 + \frac{{\left( {1 - X_{{{\text{N}}\left( {\text{bulk}} \right)}} } \right)}}{{X_{{{\text{N}}\left( {\text{bulk}} \right)}} }}\exp \left( {\frac{ - \Lambda }{{RT}}} \right)} \right] $$
(18)

\( \Lambda \) is the interaction energy of nitrogen with a dislocation in the FCC structure, estimated by

$$ \Lambda = 11848 + 824X_{\text{N}} $$
(19)

The contribution to the total transformation driving force due to the surface energy contribution of N at the stacking fault may be estimated using the methodology of Ericsson:93

$$ \Delta G^{\gamma \to \varepsilon }_{{{\text{N}},{\text{Surf}} - {\text{seg}}}} = \frac{1}{4}\Lambda \left( {X_{{{\text{N}}\left( {\text{seg}} \right)}} - X_{{{\text{N}}\left( {\text{bulk}} \right)}} } \right)^{2} $$
(20)

Magnetic Potential

The excess driving force for the γ-austenite to ε-martensite transformation due to magnetism of the respective phases (\( \Delta G^{\gamma \to \varepsilon }_{{{\text{Mag}} .}} \)) was calculated using the methodology by Inden,88 as modified by Hillert and Jarl,87 following the observations of Saeed-Akbari et al.28 and Curtze.30 The contribution to the change in free energy due to the paramagnetic to antiferromagnetic transition follows:

$$ \Delta G^{\gamma \to \varepsilon }_{{{\text{Mag}} .}} = G^{\varepsilon }_{{{\text{Mag}} .}} - G^{\gamma }_{{{\text{Mag}} .}} $$
(21)

where the magnetic free energy of each phase (i.e., \( G^{\varepsilon }_{{{\text{Mag}} .}} \) and \( G^{\gamma }_{{{\text{Mag}} .}} \)) may be calculated using:

$$ G^{\varphi }_{{{\text{Mag}} .}} = {{RT}}\ln \left( {1 + \frac{{\beta^{\varphi } }}{{\mu_{\text{B}} }}} \right)f\left( {\tau^{\varphi } } \right) $$
(22)

where R is the ideal gas constant, \( \beta^{\varphi } \) is the magnetic moment scaled by the Bohr magnetron (\( \mu_{\text{B}} \)), and \( \tau^{\varphi } \) is the temperature of interest (T) divided by paramagnetic transition or Neél temperature \( \left( T_{{\text{Ne}}\acute{{\text{e}}}{\text{l}}}^{\varphi } \right) \), for each phase (\( \varphi ) \), respectively. The scaled magnetic moment of each phase is calculated by taking the sum of the pure element’s magnetic moments (Table II) weighed by the atomic fraction of each element following Eqs. 23 and 24 for the martensite and austenite phases, respectively.

$$ \frac{{\beta^{\gamma } }}{{\mu_{\text{B}} }} = X_{\text{Fe}} \beta_{\text{Fe}}^{\gamma } + X_{\text{Ni}} \beta_{\text{Ni}}^{\gamma } + X_{\text{Mn}} \beta_{\text{Mn}}^{\gamma } + X_{\text{Cr}} \beta_{\text{Cr}}^{\gamma } + X_{\text{Fe}} \left( {X_{\text{Mn}} + X_{\text{Ni}} } \right)\beta_{\text{Fe,Mn/Ni}}^{\gamma } + X_{\text{C}} \beta_{\text{C}}^{\gamma } $$
(23)
$$ \frac{{\beta^{\varepsilon } }}{{\mu_{\text{B}} }} = X_{\text{Mn}} \beta_{\text{Mn}}^{\varepsilon } + X_{\text{C}} \beta_{\text{C}}^{\varepsilon } $$
(24)
Table II Scaled elemental magnetic moments

Only Mn and C were considered when estimating the scaled magnetic moment in the ε-martensite phase because of the lack of data on the dependence of the magnetic character of the HCP phase. The Neél temperature of the martensite and the austenite phases were calculated following Eqs. 25 and 26, respectively.

$$ T_{{\text{Ne}}\acute{{\text{e}}}{\text{l}}}^{\gamma } = 251.71 + 681X_{\text{Mn}} - 272X_{\text{Cr}} - 1800X_{\text{Ni}} - 1151X_{\text{Al}} - 1575X_{\text{Si}} - 1740X_{\text{C}} $$
(25)
$$ T_{{\text{Ne}}\acute{{\text{e}}}{\text{l}}}^{\varepsilon } = 580X_{\text{Mn}} $$
(26)

Only the Mn and C fractions were considered when estimating the scaled magnetic moment, and only Mn was included in the estimation of the Neél temperature for the ε-martensite phase because of the lack of data on the dependence of the magnetic characteristics of the HCP phase. Finally, the polynomial function \( f\left( {\tau^{\varphi } } \right) \) was calculated using Eq. 27 for \( \tau^{\varphi } \le 1 \)

$$ f(\tau^{\varphi } ) = 1\frac{{\left\{ {\frac{{79\tau^{{\varphi^{ - 1} }} }}{140p} + \frac{474}{497}\left[ {\frac{1}{p} - 1} \right]\left[ {\frac{{\tau^{{\varphi^{3} }} }}{6} + \frac{{\tau^{\varphi 9} }}{135} + \frac{{\tau^{\varphi 15} }}{600}} \right]} \right\}}}{D} $$
(27)

and Eq. 28 for \( \tau^{\varphi } > 1 \).

$$ f\left( {\tau^{\varphi } } \right) = - \frac{{\left[ {\frac{{\tau^{ - 5} }}{10} + \frac{{\tau^{ - 15} }}{315} + \frac{{\tau^{1 - 25} }}{1500}} \right]}}{D} $$
(28)

In Eqs. 27 and 28, p = 0.28 and D = 2.342457. For negative Neél temperatures the value of \( \tau^{\varphi } \) was assumed to be 300 to avoid unrealistically large contributions to the calculated stacking fault energy because of uncertainties in the estimation of the Neél temperature. Increasing this arbitrarily scaled Neél temperature did not significantly affect the resulting magnetic contribution to the calculated stacking fault energy.

As can be seen in Eqs. 23 and 25, the composition dependence of the scaled magnetic moment and the Néel temperature for the major alloy additions of interest for stainless steels are available from the literature for the austenite phase. However, very limited data exist for these parameters for the ε-martensitic phase because of the difficulty in making experimental measurements on the HCP phase in the composition space of interest; Eqs. 24 and 26 therefore only include deviations in the respective calculated properties due to Mn and C additions.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gibbs, P.J., Hough, P.D., Thürmer, K. et al. Stacking Fault Energy Based Alloy Screening for Hydrogen Compatibility. JOM 72, 1982–1992 (2020). https://doi.org/10.1007/s11837-020-04106-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11837-020-04106-7

Navigation