Skip to main content
Log in

A macroscopic multi-mechanism based constitutive model for the thermo-mechanical cyclic degeneration of shape memory effect of NiTi shape memory alloy

  • Research Paper
  • Published:
Acta Mechanica Sinica Aims and scope Submit manuscript

Abstract

A macroscopic based multi-mechanism constitutive model is constructed in the framework of irreversible thermodynamics to describe the degeneration of shape memory effect occurring in the thermo-mechanical cyclic deformation of NiTi shape memory alloys (SMAs). Three phases, austenite A, twinned martensite \(M^{\mathrm{t}}\) and detwinned martensite \(M^{\mathrm{d}}\), as well as the phase transitions occurring between each pair of phases (\(A\rightarrow M ^{\mathrm{t}}\), \(M^{\mathrm{t}}\rightarrow A\), \(A\rightarrow M ^{\mathrm{d}}\), \(M^{\mathrm{d}}\rightarrow A\), and \(M^{\mathrm{t}}\rightarrow M ^{\mathrm{d}})\) are considered in the proposed model. Meanwhile, two kinds of inelastic deformation mechanisms, martensite transformation-induced plasticity and reorientation-induced plasticity, are used to explain the degeneration of shape memory effects of NiTi SMAs. The evolution equations of internal variables are proposed by attributing the degeneration of shape memory effect to the interaction between the three phases (A, \(M^{\mathrm{t}}\), and \(M^{\mathrm{d}})\) and plastic deformation. Finally, the capability of the proposed model is verified by comparing the predictions with the experimental results of NiTi SMAs. It is shown that the degeneration of shape memory effect and its dependence on the loading level can be reasonably described by the proposed model.

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

References

  1. Miyazaki, S., Imai, T., Igo, Y., et al.: Effect of cyclic deformation on the pseudoelasticity characteristics of Ti-Ni alloys. Metall. Trans. A 17, 115–120 (1986)

    Article  Google Scholar 

  2. Song, D., Kang, G., Kan, Q., et al.: Non-proportional multiaxial transformation ratchetting of super-elastic NiTi shape memory alloy: experimental observations. Mech. Mater. 70, 94–105 (2014)

    Article  Google Scholar 

  3. Song, D., Kang, G., Kan, Q., et al.: Non-proportional multiaxial whole-life transformation ratchetting and fatigue failure of super-elastic NiTi shape memory alloy micro-tubes. Int. J. Fatigue 80, 372–380 (2015)

    Article  Google Scholar 

  4. Morin, C., Moumni, Z., Zaki, W.: Thermomechanical coupling in shape memory alloys under cyclic loadings: experimental analysis and constitutive modeling. Int. J. Plast. 27, 1959–1980 (2011)

    Article  MATH  Google Scholar 

  5. Kan, Q., Yu, C., Kang, G., et al.: Experimental observations on rate-dependent cyclic deformation of super-elastic NiTi shape memory alloy. Mech. Mater. 97, 48–58 (2016)

    Article  Google Scholar 

  6. Delville, R., Malard, B., Pilch, J., et al.: Transmission electron microscopy investigation of dislocation slip during superelastic cycling of Ni-Ti wires. Int. J. Plast. 27, 282–297 (2011)

    Article  Google Scholar 

  7. Kan, Q., Jiang, H., Xu, X., et al.: Experimental observation on thermo-mechanically cyclic deformation of NiTi shape memory alloy. Submitted for publication (2016)

  8. Huo, Y., Muller, I.: Nonequilibrium thermodynamics of pseudoelasticity. Continuum Mech. Thermodyn. 5, 163–204 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  9. Sun, Q.P., He, Y.J.: A multiscale continuum model of the grain-size dependence of the stress hysteresis in shape memory alloy polycrystals. Int. J. Solids Struct. 45, 3868–3896 (2008)

    Article  MATH  Google Scholar 

  10. Dong, L., Sun, Q.P.: On stability of elastic domain during isothermal solid-solid phase transformation in a tube configuration. Acta Mech. Sinica 28, 683–694 (2012)

    Article  Google Scholar 

  11. Xuan, C., Ding, S., Huo, Y.: Multiple bifurcations and local energy minimizers in thermoelastic martensitic transformations. Acta Mech. Sinica 31, 660–671 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  12. Li, J.Y., Lei, C.H., Li, L.J., et al.: Unconventional phase field simulations of transforming materials with evolving microstructures. Acta Mech. Sinica 28, 915–927 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  13. Yang, S.Y., Dui, G.S.: Temperature analysis of one-dimensional NiTi shape memory alloys under different loading rates and boundary conditions. Int. J. Solids Struct. 50, 3254–3265 (2013)

    Article  Google Scholar 

  14. Peng, X., Pi, W., Fan, J.: A microstructure-based constitutive model for the pseudoelastic behavior of NiTi SMAs. Int. J. Plast. 24, 966–990 (2008)

    Article  MATH  Google Scholar 

  15. Lagoudas, D.C., Entchev, P.B.: Modeling of transformation-induced plasticity and its effect on the behavior of porous shape memory alloys. Part I: constitutive model for fully dense SMAs. Mech. Mater. 36, 865–892 (2004)

    Article  Google Scholar 

  16. Auricchio, F., Reali, A., Stefanelli, U.: A three-dimensional model describing stress-induced solid phase transformation with permanent inelasticity. Int. J. Plast. 23, 207–226 (2007)

    Article  MATH  Google Scholar 

  17. Zaki, W., Moumni, Z.: A 3D model of the cyclic thermomechanical behavior of shape memory alloys. J. Mech. Phys. Solids 55, 2427–2454 (2007)

    Article  MATH  Google Scholar 

  18. Kan, Q., Kang, G.: Constitutive model for uniaxial transformation ratchetting of super-elastic NiTi shape memory alloy at room temperature. Int. J. Plast. 26, 441–465 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  19. Yu, C., Kang, G., Kan, Q.: A physical mechanism based constitutive model for temperature-dependent transformation ratchetting of NiTi shape memory alloy: One-dimensional model. Mech. Mater. 78, 1–10 (2014)

    Article  Google Scholar 

  20. Yu, C., Kang, G., Kan, Q., et al.: Rate-dependent cyclic deformation of super-elastic NiTi shape memory alloy: thermo-mechanical coupled and physical mechanism-based constitutive model. Int. J. Plast. 72, 60–90 (2015)

    Article  Google Scholar 

  21. Zhang, X., Huang, D., Yan, X., et al.: Modeling functional fatigue of SMA using a more accurate subdivision of martensite volume fractions. Mech. Mater. 96, 12–29 (2016)

    Article  Google Scholar 

  22. Yu, C., Kang, G., Kan, Q., et al.: Physical mechanism based crystal plasticity model of NiTi shape memory alloy addressing the thermo-mechanical cyclic degeneration of shape memory effect. Under review

  23. Popov, P., Lagoudas, D.C.: A 3-D constitutive model for shape memory alloys incorporating pseudoelasticity and detwinning of self-accommodated martensite. Inter. J. Plast. 23, 1679–1720 (2007)

    Article  MATH  Google Scholar 

  24. Yu, C., Kang, G., Song, D.: Effect of martensite reorientation and reorientation-induced plasticity on multiaxial transformation ratchetting of super-elastic NiTi shape memory alloy: new consideration in constitutive model. Int. J. Plast. 67, 69–101 (2015)

    Article  Google Scholar 

  25. Norfleet, D.M., Sarosi, P.M., Manchiraju, S., et al.: Transformation-induced plasticity during pseudoelastic deformation in Ni-Ti microcrystals. Acta Mater. 57, 3549–3561 (2009)

    Article  Google Scholar 

  26. Simon, T., Kröger, A., Somsen, C., et al.: On the multiplication of dislocations during martensitic transformations in NiTi shape memory alloys. Acta Mater. 58, 1850–1860 (2010)

    Article  Google Scholar 

  27. Gall, K., Maier, H.J.: Cyclic deformation mechanisms in precipitated NiTi shape memory alloys. Acta Mater. 50, 4643–4657 (2002)

    Article  Google Scholar 

  28. Wagner, M.F., Nayan, N., Ramamurty, U.: Healing of fatigue damage in NiTi shape memory alloys. J. Phys. D: Appl. Phys. 41, 185408 (2008)

    Article  Google Scholar 

  29. Grabe, C., Bruhns, O.T.: On the viscous and strain rate dependent behavior of polycrystalline NiTi. Int. J. Solids Struct. 45, 1876–1895 (2008)

    Article  MATH  Google Scholar 

  30. He, Y.J., Sun, Q.P.: Frequency-dependent temperature evolution in NiTi shape memory alloy under cyclic loading. Smart Mater. Struct. 19, 115014 (2010)

    Article  Google Scholar 

  31. Yin, H., He, Y., Sun, Q.: Effect of deformation frequency on temperature and stress oscillations in cyclic phase transition of NiTi shape memory alloy. J. Mech. Phys. Solids 67, 100–128 (2014)

    Article  Google Scholar 

  32. Mecking, H., Kocks, U.F.: Kinetics of flow and strain-hardening. Acta Metall. 29, 1865–1875 (1981)

  33. Mareau, C., Favier, V., Weber, B., et al.: Micromechanical modeling of the interactions between the microstructure and the dissipative deformation mechanisms in steels under cyclic loading. Int. J. Plast. 32, 106–120 (2012)

    Article  Google Scholar 

  34. Otsuka, K., Ren, X., et al.: Physical metallurgy of Ti-Ni-based shape memory alloys. Prog. Mater. Sci 50, 511–678 (2005)

    Article  Google Scholar 

Download references

Acknowledgements

Financial supports by the National Natural Science Foundation of China (Grant 11532010) and the project for Sichuan Provincial Youth Science and Technology Innovation Team, China (Grant 2013TD0004) are appreciated.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Guozheng Kang.

Appendix

Appendix

Integrating Eq. (23) yields:

$$\begin{aligned} \sqrt{\rho }=\frac{k_1 }{k_2 }\left[ 1-\hbox {e}^{-\frac{k_2 }{2}\left( {\gamma _{\mathrm{tp}} +\gamma _{\mathrm{rp}} } \right) }\right] . \end{aligned}$$
(A1)

Equation (A1) is the explicit relationship between the slippage amount \(\gamma _{\mathrm{tp}} +\gamma _{\mathrm{rp}} \) and the dislocation density \(\rho \). Thus, the dislocation densities after the first and tenth cycles can be obtained as

$$\begin{aligned} \sqrt{\rho _1 }= & {} \frac{k_1 }{k_2 }\left( 1-\hbox {e}^{-\frac{k_2 \varepsilon _\mathrm{p}^\mathrm{1} }{2}}\right) , \end{aligned}$$
(A2a)
$$\begin{aligned} \sqrt{\rho _{10} }= & {} \frac{k_1 }{k_2 }\left( 1-\hbox {e}^{-\frac{k_2 \varepsilon _\mathrm{p}^{{10}} }{2}}\right) , \end{aligned}$$
(A2b)

where \(\rho _1 \) and \(\rho _{10} \) are the dislocation densities after the first and tenth cycles, respectively.

In the first loading cycle, the reorientation condition (Eq. 18b) at point \(\sigma _{\mathrm{reo}}^1 \) can be written as (note that \({\varvec{B}}_{\mathrm{reo}} \) and \(f_3 \) are both zero)

$$\begin{aligned} g\sigma _{\mathrm{reo}}^1 =Y_{M_\mathrm{t} \rightarrow M_\mathrm{d} } . \end{aligned}$$
(A3)

The parameter \(Y_{M_\mathrm{t} \rightarrow M_\mathrm{d} } \) can be determined by Eq. (A3), as shown in Eq. (34a).

At the inflection point \(\left( {\sigma ^{*},\varepsilon ^{*}} \right) \), since the applied stress is very low, the plastic deformation can be neglected. In this condition, the total strain consists of the elastic strain and the transformation strain, i.e.,

$$\begin{aligned} \varepsilon ^{*}=\frac{\sigma ^{*}}{E}+g\lambda _{\mathrm{ref}} . \end{aligned}$$
(A4)

The parameter \(\lambda _{\mathrm{ref}} \) can be obtained by Eq. (A4), as shown in Eq. (34b).

The reorientation condition at the inflection point can be written as

$$\begin{aligned} g\sigma ^*-H_{M_\mathrm{t} \rightarrow M_\mathrm{d} }^\mathrm{I} \lambda _{\mathrm{ref}} =Y_{M_\mathrm{t} \rightarrow M_\mathrm{d} } . \end{aligned}$$
(A5)

The parameter \(H_{M_\mathrm{t} \rightarrow M_\mathrm{d} }^\mathrm{I} \) can be obtained by Eq. (A5), as shown in Eq. (34c).

Considering the transition between the austenite and twinned martensite phases under the stress-free conditions, the transformation condition at the finish point of the forward transformation and the start point of the reverse transformation can be written as (by Eqs. (15a), (16b), (16c), and (20a))

$$\begin{aligned}&-\left[ {\beta \left( {M_\mathrm{f} -M_\mathrm{s} } \right) -Y_{A\rightarrow M_\mathrm{t} } +H_{A\rightarrow M_\mathrm{t} } } \right] = Y_{A\rightarrow M_\mathrm{t} } , \end{aligned}$$
(A6a)
$$\begin{aligned}&-\left[ {\beta \left( {A_\mathrm{s} -M_\mathrm{s} } \right) -Y_{A\rightarrow M_\mathrm{t} } +H_{A\rightarrow M_\mathrm{t} } } \right] = -Y_{A\rightarrow M_\mathrm{t} } . \end{aligned}$$
(A6b)

Two parameters \(H_{A\rightarrow M_\mathrm{t} } \) and \(Y_{A\rightarrow M_\mathrm{t} } \) can be obtained by Eq. (A6a) and  (A6b), as shown in Eq. (34d) and (34e).

Rewrite Eqs. (26a), (26b), (29), and (31) in the following form

$$\begin{aligned} \dot{Z}_i =d\left( {c_i \sqrt{\rho }-Z_i } \right) \left( {\sum _{i=1}^3 {\left| {\dot{\lambda }_i } \right| } } \right) ,\quad i=1,2,3,4,5, \end{aligned}$$
(A7)

where \(Z_1, Z_2, Z_3, Z_4\), and \(Z_5 \) represent \(B_{\mathrm{tr}}^\mathrm{n} \), \(B_{\mathrm{reo}}^\mathrm{n} \), \(H_{A\rightarrow M_\mathrm{d} }^{\uprho } \), \(H_{M_\mathrm{t} \rightarrow M_\mathrm{d} }^{\mathrm{I\uprho }} \), and \(Y_{A\rightarrow M_\mathrm{d} }^{\uprho } \), respectively. Note that Eq. (A7) cannot be integrated unless the \(\sqrt{\rho }\) is regarded as a constant in the first cycle. Thus, for simplicity, it is assumed that \(\sqrt{\rho }\) can be replaced by \(\sqrt{\rho _1 }\) in Eq. (A7). Then, integrating Eq. (A7) yields:

$$\begin{aligned} Z_i =c_i \sqrt{\rho _1 }\left[ {1-\exp \left( {-d\sum _{i=1}^3 {\bar{{\lambda }}_i } } \right) } \right] ,\quad i=1,2,3,4,5,\nonumber \\ \end{aligned}$$
(A8)

where \(\dot{\bar{{\lambda }}}_i =\left| {\dot{\lambda }_i } \right| \).

From Eqs. (21a) and (22a), it can be concluded that the amount of dislocation slippage in the loading part is much larger than that in the unloading and cooling/heating processes. Thus, in the first and tenth cycles, the plastic strains at maximum stress points are approximately equal to \(\varepsilon _\mathrm{p}^1 \) and \(\varepsilon _\mathrm{p}^{10} \), respectively. Thus, the maximum strains in the first and tenth cycles can be given as

$$\begin{aligned} \varepsilon _{\max }^1= & {} \frac{\sigma _{\max }^1 }{E}+g\xi _\mathrm{d}^{\max ,1} +\varepsilon _\mathrm{p}^1 , \end{aligned}$$
(A9a)
$$\begin{aligned} \varepsilon _{\max }^{10}= & {} \frac{\sigma _{\max }^{10} }{E}+g\xi _\mathrm{d}^{\max ,10} +\varepsilon _\mathrm{p}^{10} . \end{aligned}$$
(A9b)

By Eq. (A9), the maximum volume fraction of the detwinned martensite in the first cycle, i.e., \(\xi _\mathrm{d}^{\max ,1} \) can be obtained as

$$\begin{aligned} \xi _\mathrm{d}^{\max ,1} =\frac{\varepsilon _{\max }^1 -\frac{\sigma _{\max }^1 }{E}-\varepsilon _\mathrm{p}^1 }{g}, \end{aligned}$$
(A10a)
$$\begin{aligned} \xi _\mathrm{d}^{\max ,10} =\frac{\varepsilon _{\max }^{10} -\frac{\sigma _{\max }^{10} }{E}-\varepsilon _\mathrm{p}^{10} }{g}. \end{aligned}$$
(A10b)

By Eq. (A8), the variable \(Z_i \) at the start and finish points, respectively, of reverse transformation in the first cycle can be written as

$$\begin{aligned} Z_{i,A_{\hbox {s}}}^1= & {} c_i \sqrt{\rho _1 }\left[ {1-\exp \left( {-d\xi _\mathrm{d}^{\max ,1} } \right) } \right] ,\quad i=1,2,3,4,5, \nonumber \\ \end{aligned}$$
(A11a)
$$\begin{aligned} Z_{i,A_{\hbox {f}}}^1= & {} c_i \sqrt{\rho _1 }\left[ {1-\exp \left( {-2d\xi _\mathrm{d}^{\max ,1} } \right) } \right] ,\quad i=1,2,3,4,5.\nonumber \\ \end{aligned}$$
(A11b)

For the transition between the austenite and detwinned martensite phases, the transformation conditions at the start and finish points of reverse transformation in the first cycle can be written as, respectively (by Eqs. (15b), (17b), (17c), (20b), (26)–(31))

$$\begin{aligned}&gc_1 R-\beta \left( {T_{A_{\hbox {s}}}^1 -M_\mathrm{s} } \right) +Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} -c_5 R\nonumber \\&\quad -\left( {H_{A\rightarrow M_\mathrm{d} }^\mathrm{f} +c_3 R} \right) \xi _\mathrm{d}^{\max ,1} =-Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} +c_5 R, \end{aligned}$$
(A12a)
$$\begin{aligned}&gc_1 P-\beta \left( {T_{A_{\hbox {f}}}^1 -M_\mathrm{s} } \right) +Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} -c_5 P\nonumber \\&\quad =-Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} +c_5 P, \end{aligned}$$
(A12b)

where \(R=\sqrt{\rho _1 }\left[ {1-\exp \left( {-d\xi _\mathrm{d}^{\max ,1} } \right) } \right] \), \(P=\sqrt{\rho _1 }\times \left[ {1-\exp \left( {-2d\xi _\mathrm{d}^{\max ,1} } \right) } \right] \).

In the tenth cycle, the term \(\exp \left( {-d\sum \limits _{i=1}^3 {\bar{{\lambda }}_i } } \right) \) approaches zero, since \(\sum \limits _{i=1}^3 {\bar{{\lambda }}_i } \) is a very large number. Thus, the variable \(Z_i \) at the start and finish points, respectively, of the reverse transformation in the tenth cycle can be written as

$$\begin{aligned} Z_{i,A{\mathrm{s}}}^{10} =c_i \sqrt{\rho _{10} },\quad i=1,2,3,4,5, \end{aligned}$$
(A13a)
$$\begin{aligned} Z_{i,A{\mathrm{f}}}^{10} =c_i \sqrt{\rho _{10} },\quad i=1,2,3,4,5. \end{aligned}$$
(A13b)

Similarly, the transformation conditions at the start and finish points of the reverse transformation can be written as:

$$\begin{aligned}&gc_1 \sqrt{\rho _{10} }-\beta \left( {T_{A\mathrm{s}}^{10} -M_\mathrm{s} } \right) +Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} -c_5 \sqrt{\rho _{10} }\nonumber \\&\qquad -\left( {H_{A\rightarrow M_\mathrm{d} }^\mathrm{f} +c_3 \sqrt{\rho _{10} }} \right) \xi _\mathrm{d}^{\max ,10} \nonumber \\&\quad =-Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} +c_5 \sqrt{\rho _{10} }, \end{aligned}$$
(A14a)
$$\begin{aligned}&gc_1 \sqrt{\rho _{10} }-\beta \left( {T_{A\mathrm{f}}^{10} -M_\mathrm{s} } \right) +Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} -c_5 \sqrt{\rho _{10} }\nonumber \\&\quad =-Y_{A\rightarrow M_\mathrm{d} }^\mathrm{f} +c_5 \sqrt{\rho _{10} }. \end{aligned}$$
(A14b)

After the cyclic deformation, the residual strain \(\varepsilon _\mathrm{r}^{\mathrm{10}} \) consists of two parts, the transformation strain and the plastic strain, i.e.,

$$\begin{aligned} \varepsilon _\mathrm{r}^{10} =g\xi _\mathrm{d}^{M_{\hbox {f}},10} +\varepsilon _\mathrm{p}^{10} . \end{aligned}$$
(A15)

By Eq. (A15), the volume fraction of the detwinned martensite phase at the finish point of the forward transformation in the tenth cycle, i.e., \(\xi _\mathrm{d}^{M_{\hbox {f}},10} \) can be obtained as

$$\begin{aligned} \xi _\mathrm{d}^{M_{\hbox {f}},10} =\frac{\varepsilon _\mathrm{r}^{10} -\varepsilon _\mathrm{p}^{10} }{g}. \end{aligned}$$
(A16)

The transformation condition at this point can be written as

$$\begin{aligned} gc_1 \sqrt{\rho _{10} }-\beta \left\langle {T_{M_{\hbox {f}}}^{10} -M_\mathrm{s} } \right\rangle -\left( {H_{A\rightarrow M_\mathrm{d} }^\mathrm{f} +c_3 \sqrt{\rho _{10} }} \right) \xi _\mathrm{d}^{M_{\hbox {f}},10} =0.\nonumber \\ \end{aligned}$$
(A17)

By Eqs. (A12a), (A12b), (A14a), (A14b), and (A17), the five parameters \(H_{A\rightarrow M_\mathrm{d} }^\mathrm{f} \), \(Y_{A\rightarrow M_\mathrm{d} }^0 \), \(c_1 \), \(c_3 \), and \(c_5 \) can be obtained, as shown in Eq. (34f).

In the tenth cycle, the reorientation condition at point \(\sigma _{\mathrm{reo}}^{10} \) can be written as

$$\begin{aligned} g\left( {\sigma _{\mathrm{reo}}^{10} +c_2 \sqrt{\rho _{10} }} \right) =Y_{M_\mathrm{t} \rightarrow M_\mathrm{d} } . \end{aligned}$$
(A18)

The parameter \(c_2 \) can be determined by Eq. (A18), as shown in Eq. (34g).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yu, C., Kang, G. & Kan, Q. A macroscopic multi-mechanism based constitutive model for the thermo-mechanical cyclic degeneration of shape memory effect of NiTi shape memory alloy. Acta Mech. Sin. 33, 619–634 (2017). https://doi.org/10.1007/s10409-016-0632-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10409-016-0632-9

Keywords

Navigation