Skip to main content
Log in

Effect of Strain Rate on the Formation of Strain-Induced Martensite in AISI 304L Stainless Steel

  • Published:
Metallurgical and Materials Transactions A Aims and scope Submit manuscript

Abstract

The effect of strain rate on the stress–strain response and the austenite to strain-induced α′-martensite transformation of austenitic steel 304L was studied. Compression tests were carried out at room temperature in the strain rate range of 10−3 to 103 s−1 and the evolution of martensite was quantified using a ferritoscope. Higher strain rates resulted in lower strain-induced α′-martensite. Strain incremental tests were carried out at 1 s−1 to simulate isothermal tests and to delineate effect of adiabatic heating on the α′-martensite transformation. Strain rate change tests were also carried out to determine the effect of prior strain rate history on the strain-induced α′-martensite content. These experiments showed that apart from the effect of adiabatic heating at high strain rates on the α′-martensite content, there is an additional effect of strain rate. While the adiabatic heating effect could be based on the increase in stacking fault energy (reduction of stacking fault width) with temperature, the additional effect due to strain rate was explained based on the expected reduction in stacking fault width with increasing strain rate.

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
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Notes

  1. The ferritoscope is calibrated for the measurement of ferrite content based on magnetic measurements for ferrite with very low alloying elements. The factor of 1.71 arises because the magnetic property of martensite in austenitic stainless steels is different from that of pure ferrite.

References

  1. T. Angel: Journal of Iron and Steel Institute, 1954, vol. 177, pp. 165-174.

    CAS  Google Scholar 

  2. G.H. Eichelmann, F.C. Hull: Trans. ASM, 1953, vol. 45, pp. 45.

    Google Scholar 

  3. G.B. Olson, M. Cohen: Metallurgical Transactions A, 1975, vol. 6, pp. 791.

    Article  CAS  Google Scholar 

  4. J.R. Patel, M. Cohen: Acta Metallurgica, 1953, vol. 1, pp. 531-538.

    Article  CAS  Google Scholar 

  5. R. Lagneborgj: Acta Metallurgica, 1964, vol. 12, pp. 823-843.

    Article  Google Scholar 

  6. P.L. Mangonon, G. Thomas: Metallurgical Transactions, 1970, vol. 1, pp. 1577-1586.

    Article  Google Scholar 

  7. J.A. Venables: Philosophical Magazine, 1962, vol. 7, pp. 35-44.

    Article  Google Scholar 

  8. S.S. Hecker, M.G. Stout, K.P. Staudhammer, J.L. Smith: Metallurgical Transactions A, 1982, vol. 13, pp. 619-626.

    Article  Google Scholar 

  9. G.L. Huang, D.K. Matlock, G. Krauss: Metallurgical Transactions A, 1989, vol. 20, pp. 1239-1246.

    Article  CAS  Google Scholar 

  10. M. Isakov, S. Hiermaier, V.-T. Kuokkala: Metallurgical and Materials Transactions A, 2015, vol. 46, pp. 2352-2355.

    Article  Google Scholar 

  11. J.A. Lichtenfeld, C.J. Van Tyne, M.C. Mataya: Metallurgical and Materials Transactions A, 2006, vol. 37, pp. 147-161.

    Article  Google Scholar 

  12. L.E. Murr, K.P. Staudhammer, S.S. Hecker: Metallurgical Transactions A, 1982, vol. 13, pp. 627-635.

    Article  Google Scholar 

  13. J. Talonen, H. Hänninen, P. Nenonen, G. Pape: Metallurgical and Materials Transactions A, 2005, vol. 36, pp. 421-432.

    Article  CAS  Google Scholar 

  14. W.-S. Lee, C.-F. Lin: Materials Science and Engineering A, 2001, vol. 308, pp. 124-135.

    Article  Google Scholar 

  15. N. Vázquez-Fernández, T. Nyyssönen, M. Isakov, M. Hokka, V.-T. Kuokkala: Acta Materialia, 2019, vol. 176, pp. 134-144.

    Article  Google Scholar 

  16. S. Nemat-Nasser: ASM International, West Conshohocken, 1985, PA, p. 190.

  17. S. Sharma, R. Kapoor, V.M. Chavan, R.G. Agrawal, R.J. Patel, and J.K. Chakravartty: BARC External Report, 2011, BARC/2011/E/013.

  18. J. Talonen, P. Aspegren, H. Hänninen: Materials Science and Technology, 2004, vol. 20, pp. 1506-1512.

    Article  CAS  Google Scholar 

  19. Y.F. Shen, X.X. Li, X. Sun, Y.D. Wang, L. Zuo: Materials Science and Engineering A, 2012, vol. 552, pp. 514-522.

    Article  CAS  Google Scholar 

  20. G. Olson, M. Cohen: Journal of the Less Common Metals, 1972, vol. 28, pp. 107-118.

    Article  CAS  Google Scholar 

  21. L. Rémy, A. Pineau, B. Thomas: Materials Science and Engineering, 1978, vol. 36, pp. 47-63.

    Article  Google Scholar 

  22. J. Talonen, H. Hänninen: Acta Materialia, 2007, vol. 55, pp. 6108-6118.

    Article  CAS  Google Scholar 

  23. D. Hull and D.J. Bacon: Introduction to Dislocations, Elsevier Science & Technology Books, Amsterdam, 1984.

  24. Y.N. Petrov: Zeitschrift für Metallkunde, 2003, vol. 94, pp. 1012-1016.

    Article  CAS  Google Scholar 

  25. D. Kaoumi, J. Liu: Materials Science and Engineering A, 2018, vol. 715, pp. 73-82.

    Article  CAS  Google Scholar 

  26. A. Miodownik: Calphad, 1978, vol. 2, pp. 207-226.

    Article  CAS  Google Scholar 

  27. E.I. Galindo-Nava, P. Rivera-Díaz-del-Castillo: Acta Materialia, 2017, vol. 128, pp. 120-134.

    Article  CAS  Google Scholar 

  28. G. Taylor: Progress in Materials Science, 1992, vol. 36, pp. 29-61.

    Article  CAS  Google Scholar 

  29. U.F. Kocks, A.S. Argon and M.F. Ashby: Progress in Materials Science, 1975, vol. 19, pp. 1-291.

    Article  Google Scholar 

Download references

Acknowledgments

This work was supported under the BARC 12th plan project XII-N-R&D-25 ‘Experimental Studies for Ageing and Life Extension of Nuclear Components’.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rajeev Kapoor.

Additional information

Publisher's Note

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

Manuscript submitted May 23, 2020.

Appendix A

Appendix A

The temperature rise in the sample due to plastic deformation can be estimated by considering the heat dissipation to the environment by the specimen as a one dimension heat transfer problem as shown in Figure A1. The specimen is touching the platens at either end of its length, and it is taken that the ends are always at ambient temperature of 25 °C. There is a constant heat source Q present in the specimen due to the work being done on it that is getting converted to heat. In actual experiments the specimen length decreases with strain, however for sake of simplifying the problem the length of sample is kept fixed. Thus, as the conduction path is slightly longer than actual, the temperature solved for may be a small overestimation.

Fig. A1
figure 9

Schematic of 1-D specimen with the ends held at ambient temperature (25 °C)

The heat transfer equation in 1-dimension is given by

$$ \frac{\partial T}{\partial t} = \frac{K}{{\rho \cdot C_{\text{p}} }} \cdot \frac{{\partial^{2} T}}{{\partial x^{2} }} + \frac{Q}{{\rho \cdot C_{\text{p}} }}, $$
(A1)

where ρ is mass density (kg/m3), Cp is specific heat capacity (J/kg K), K is thermal conductivity (W/K m), and Q is heat generation rate per unit volume (W/m3). In the present case Q is derived from the incremental work done per unit volume dw on the specimen and is given by

$$ {\text{d}}w = \sigma \cdot {\text{d}}\varepsilon $$
(A2)

Of the plastic work increment (Eq. [A2]) some will be consumed in increasing the defect density (internal energy) and the remaining will generate heat (factor β taken here as 0.9). Thus the rate of heat generation is

$$ Q = \beta \frac{{{\text{d}}w}}{{{\text{d}}t}} = \beta {\kern 1pt} \sigma \frac{{{\text{d}}\varepsilon }}{{{\text{d}}t}} = \beta \,{\kern 1pt} \sigma \,\dot{\varepsilon } $$
(A3)

For the sake of estimating an average number for Q, it can be assumed as a first approximation that σ varies about linearly with strain, and thus an average value of σ over the strain range should suffice to get an average value of Q. From Figure 1 of text, σ varies from ~300 to 1200 MPa at a strain of 0.5 for 10−3 to 1 s−1 and varies from ~600 to 1200 MPa for 103 s−1. Hence an average value of 750 MPa for strain rate up to 1 s−1 and 900 MPa for strain rate of 103 s−1 was taken. Thus the values of Q using Eq. [A3] for strain rates of 10−3, 10−1, 1 and 103 s−1 are 6.75×105, 6.75×107, 6.75×108, and 8 × 1011 W/m3, respectively.

Equation [A1] needs to be solved numerically using a finite difference method for which time steps need to be given. Solving for different strain rates would require different time steps. To overcome this, the variable of time t was converted to strain ε, as the strain increments can be kept same for all strain rates used. The variable t is replaced by ε using

$$ \frac{\partial T}{\partial t} = \frac{\partial T}{\partial \varepsilon }\frac{\partial \varepsilon }{\partial t} = \frac{\partial T}{\partial \varepsilon }\dot{\varepsilon }. $$
(A4)

Substituting Eq. [A4] in Eq. [A1], the 1D heat transfer equations can be re-written in terms of ε as

$$ \frac{\partial T}{\partial \varepsilon } = \left( {\frac{K}{{\dot{\varepsilon }\,\rho \,C_{P} }}} \right)\frac{{\partial^{2} T}}{{\partial x^{2} }} + \frac{Q}{{\dot{\varepsilon }\,\rho \,C_{P} }}. $$
(A5)

The initial condition is T(x, ε = 0) = 25 °C, and the boundary conditions are T(0, ε) = 25 °C and T(L, ε) = 25 °C. For SS 304 ρ is 7900 kg/m3, Cp is 500 J/(kg K), K is 16 W/(m K). Figure A2 shows the results of the numerical solution of partial differential equation (Eq. [A5]).

Fig. A2
figure 10

Temperature rise at the center of the specimen as a function of strain (a), and temperature rise at a strain of 0.5 as a function of distance in the specimen from one end at 0 mm to another end at 7 mm (b)

There was no temperature rise at 10−3 s−1 making it a completely isothermal test. The temperature rise at 103 s−1 was the highest of about 100 °C and was near constant throughout its length making it a completely adiabatic test. At 10−1 s−1 there was a small temperature rise of about 20 °C, while that at 1 s−1 there was a temperature rise of 80 °C. The strain rates 10−1 and 1 s−1 are in the transition zone between isothermal and adiabatic conditions, with 10−1 s−1 being a near isothermal test, while 1 s−1 being a near adiabatic test.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sunil, S., Kapoor, R. Effect of Strain Rate on the Formation of Strain-Induced Martensite in AISI 304L Stainless Steel. Metall Mater Trans A 51, 5667–5676 (2020). https://doi.org/10.1007/s11661-020-05968-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11661-020-05968-x

Navigation