Skip to main content

Advertisement

Log in

Modeling of the stress–strain responses and deformation patterns of superelastic NiTi tubes subjected to biaxial loadings

  • Original Paper
  • Published:
Aerospace Systems Aims and scope Submit manuscript

Abstract

Nearly equiatomic NiTi shape memory alloy exhibits superelasticity, i.e., it can be strained up to ~ 7% and recover completely upon unloading, and consequently, the stress–strain response forms a closed hysteresis. The mechanical behavior of superelastic NiTi is characterized by significant tension–compression asymmetry, which leads to complexity in the stress–strain responses and deformation patterns of thin-walled superelastic NiTi tubes loaded by axial force and internal pressure simultaneously. In the reported biaxial experiments, the NiTi tube exhibits hardening responses and essentially homogeneous deformation in a neighborhood of equibiaxiality. In other cases, its stress–strain responses trace stress plateaus associated with localized deformation patterns, and the level of plateaus, magnitude of transformation strains, and orientation of the localization bands are strongly dependent on the axial-to-hoop stress ratio. In this paper, finite element modeling is performed to analyze numerically the mechanical response of biaxially loaded superelastic NiTi tube. A numerical feedback control scheme is developed to maintain the stress ratio to follow the target value. The simulations reproduce successfully the observed phenomena in the experiments, such as the localization of helical bands, the variation of band angles with stress ratio, as well as the hardening and uniform deformation near the state of equibiaxial stress. In addition, the variation of axial and hoop stress–strain responses with different stress ratios are also studied, which are reasonably close to the experimental ones. The presented work demonstrates the validity of the developed finite element analysis framework and paves the way for analysis of superelastic shape memory alloy structures under multiaxial loading.

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
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Data availability

The datasets analyzed or generated during the present study are available from the corresponding author upon reasonable requests.

References

  1. Mohd Jani J, Leary M, Subic A, Gibson MA (2014) A review of shape memory alloy research, applications and opportunities. Mater Des 56:1078–1113. https://doi.org/10.1016/J.MATDES.2013.11.084

    Article  Google Scholar 

  2. Fähler S, Rößler UK, Kastner O et al (2012) Caloric effects in ferroic materials: new concepts for cooling. Adv Eng Mater 14:10–19. https://doi.org/10.1002/ADEM.201100178

    Article  Google Scholar 

  3. Kwon S-C, Jeon S-H, Oh H-U (2015) Performance evaluation of spaceborne cryocooler micro-vibration isolation system employing pseudoelastic SMA mesh washer. Cryogenics (Guildf) 67:19–27. https://doi.org/10.1016/j.cryogenics.2015.01.002

    Article  Google Scholar 

  4. Padula S, Creager C (2018) Shape memory alloy tires: a new paradigm in tire performance

  5. Petrini L, Migliavacca F (2011) Biomedical applications of shape memory alloys. J Metall 2011:501483. https://doi.org/10.1155/2011/501483

    Article  Google Scholar 

  6. Engelbrecht K (2019) Future prospects for elastocaloric devices. J Phys Energy 1:021001. https://doi.org/10.1088/2515-7655/ab1573

    Article  Google Scholar 

  7. Tušek J, Engelbrecht K, Mikkelsen LP, Pryds N (2015) Elastocaloric effect of Ni-Ti wire for application in a cooling device. J Appl Phys 117:124901. https://doi.org/10.1063/1.4913878

    Article  Google Scholar 

  8. Takeuchi I, Sandeman K (2015) Solid-state cooling with caloric materials. Phys Today 68:48. https://doi.org/10.1063/PT.3.3022

    Article  Google Scholar 

  9. Shaw JA, Kyriakides S (1995) Thermomechanical aspects of NiTi. J Mech Phys Solids 43:1243–1281. https://doi.org/10.1016/0022-5096(95)00024-D

    Article  Google Scholar 

  10. Zhang X, Feng P, He Y et al (2010) Experimental study on rate dependence of macroscopic domain and stress hysteresis in NiTi shape memory alloy strips. Int J Mech Sci 52:1660–1670. https://doi.org/10.1016/J.IJMECSCI.2010.08.007

    Article  Google Scholar 

  11. Shaw JA, Kyriakides S (1997) On the nucleation and propagation of phase transformation fronts in a NiTi alloy. Acta Mater 45:683–700. https://doi.org/10.1016/S1359-6454(96)00189-9

    Article  Google Scholar 

  12. Li ZQ, Sun QP (2002) The initiation and growth of macroscopic martensite band in nano-grained NiTi microtube under tension. Int J Plast 18:1481–1498. https://doi.org/10.1016/S0749-6419(02)00026-8

    Article  Google Scholar 

  13. Watkins RT, Reedlunn B, Daly S, Shaw JA (2018) Uniaxial, pure bending, and column buckling experiments on superelastic NiTi rods and tubes. Int J Solids Struct 146:1–28. https://doi.org/10.1016/J.IJSOLSTR.2018.01.037

    Article  Google Scholar 

  14. Jiang D, Kyriakides S, Bechle NJ, Landis CM (2017) Bending of pseudoelastic NiTi tubes. Int J Solids Struct 124:192–214. https://doi.org/10.1016/j.ijsolstr.2017.06.032

    Article  Google Scholar 

  15. Hallai JF, Kyriakides S (2013) Underlying material response for Lüders-like instabilities. Int J Plast 47:1–12. https://doi.org/10.1016/j.ijplas.2012.12.002

    Article  Google Scholar 

  16. Xiao Y, Zeng P, Lei L (2016) Experimental investigation on the mechanical instability of superelastic NiTi shape memory alloy. J Mater Eng Perform 25:3551–3557. https://doi.org/10.1007/s11665-016-2170-9

    Article  Google Scholar 

  17. Zhao Z, Jiang D, Xiao Y et al (2023) Intrinsic response of nanocrystalline superelastic NiTi shape memory alloy. Extrem Mech Lett 60:101988. https://doi.org/10.1016/J.EML.2023.101988

    Article  Google Scholar 

  18. Greenly JL, Kyriakides S, Tsimpoukis S (2023) On the underlying material response of pseudoelastic NiTi. Eur J Mech A/Solids. https://doi.org/10.1016/J.EUROMECHSOL.2023.105023

    Article  Google Scholar 

  19. Sun QP, Li ZQ (2002) Phase transformation in superelastic NiTi polycrystalline micro-tubes under tension and torsion–from localization to homogeneous deformation. Int J Solids Struct 39:3797–3809. https://doi.org/10.1016/S0020-7683(02)00182-8

    Article  Google Scholar 

  20. Lexcellent C, Blanc P (2004) Phase transformation yield surface determination for some shape memory alloys. Acta Mater 52:2317–2324. https://doi.org/10.1016/J.ACTAMAT.2004.01.022

    Article  Google Scholar 

  21. McNaney JM, Imbeni V, Jung Y et al (2003) An experimental study of the superelastic effect in a shape-memory Nitinol alloy under biaxial loading. Mech Mater 35:969–986. https://doi.org/10.1016/S0167-6636(02)00310-1

    Article  Google Scholar 

  22. Grabe C, Bruhns OT (2009) Path dependence and multiaxial behavior of a polycrystalline NiTi alloy within the pseudoelastic and pseudoplastic temperature regimes. Int J Plast 25:513–545. https://doi.org/10.1016/J.IJPLAS.2008.03.002

    Article  Google Scholar 

  23. Mehrabi R, Andani MT, Kadkhodaei M, Elahinia M (2015) Experimental study of NiTi thin-walled tubes under uniaxial tension, torsion, proportional and non-proportional loadings. Exp Mech 55:1151–1164. https://doi.org/10.1007/s11340-015-0016-2

    Article  Google Scholar 

  24. Reedlunn B, LePage WS, Daly SH, Shaw JA (2020) Axial-torsion behavior of superelastic tubes: part I, proportional isothermal experiments. Int J Solids Struct 199:1–35. https://doi.org/10.1016/J.IJSOLSTR.2020.03.018

    Article  Google Scholar 

  25. Rezaee-Hajidehi M, Stupkiewicz S (2021) Modelling of propagating instabilities in pseudoelastic NiTi tubes under combined tension–torsion: helical bands and apparent yield locus. Int J Solids Struct 221:130–149. https://doi.org/10.1016/J.IJSOLSTR.2020.09.011

    Article  Google Scholar 

  26. Grolleau V, Louche H, Delobelle V et al (2011) Assessment of tension–compression asymmetry of NiTi using circular bulge testing of thin plates. Scr Mater 65:347–350. https://doi.org/10.1016/J.SCRIPTAMAT.2011.05.003

    Article  Google Scholar 

  27. Meng JK, Liu L, Jiang JT et al (2023) Fracture behaviors of commercially pure titanium under biaxial tension: experiment and modeling. J Mater Sci Technol 140:176–186. https://doi.org/10.1016/J.JMST.2022.08.035

    Article  Google Scholar 

  28. Bechle NJ, Kyriakides S (2016) Evolution of localization in pseudoelastic NiTi tubes under biaxial stress states. Int J Plast 82:1–31. https://doi.org/10.1016/J.IJPLAS.2016.01.017

    Article  Google Scholar 

  29. Raniecki B, Lexcellent C (1998) Thermodynamics of isotropic pseudoelasticity in shape memory alloys. Eur J Mech A/Solids 17:185–205. https://doi.org/10.1016/S0997-7538(98)80082-X

    Article  Google Scholar 

  30. Qidwai MA, Lagoudas DC (2000) On thermomechanics and transformation surfaces of polycrystalline NiTi shape memory alloy material. Int J Plast 16:1309–1343. https://doi.org/10.1016/S0749-6419(00)00012-7

    Article  Google Scholar 

  31. Aleong D, Dumont C, Chirani SA, Patoor E, McDowell DL (2002) Transformation surfaces of a textured pseudoelastic polycrystalline Cu-Zn-Al shape memory alloy. J Intell Mater Syst Struct 13(12):783–793. https://doi.org/10.1177/1045389X02013012004

    Article  Google Scholar 

  32. Auricchio F, Petrini L (2004) A three-dimensional model describing stress-temperature induced solid phase transformations: solution algorithm and boundary value problems. Int J Num Meth Eng 61:807–836. https://doi.org/10.1002/nme.1086

    Article  MathSciNet  Google Scholar 

  33. Sedlak P, Frost M, Benešova B, Ben Zineb T, Šittner P (2012) Thermomechanical model for NiTi-based shape memory alloys including R-phase and material anisotropy under multi-axial loadings. Intl J Plast 39:132–151. https://doi.org/10.1016/j.ijplas.2012.06.008

    Article  Google Scholar 

  34. Karamooz Ravari MR, Kadkhodaei M, Ghaei A (2015) A microplane constitutive model for shape memory alloys considering tension-compression asymmetry. Smart Mater Struct 24:075016. https://doi.org/10.1088/0964-1726/24/7/075016

    Article  Google Scholar 

  35. Kan Q, Zhang Y, Xu Y, Kang G, Yu C (2023) Tension-compression asymmetric functional degeneration of super-elastic NiTi shape memory alloy: experimental observation and multiscale constitutive model. Int J Solids Struct 280:112384. https://doi.org/10.1016/j.ijsolstr.2023.112384

    Article  Google Scholar 

  36. Jiang D, Landis CM, Kyriakides S (2016) Effects of tension/compression asymmetry on the buckling and recovery of NiTi tubes under axial compression. Int J Solids Struct 100–101:41–53. https://doi.org/10.1016/j.ijsolstr.2016.07.003

    Article  Google Scholar 

  37. Jiang D, Kyriakides S, Landis CM (2017) Propagation of phase transformation fronts in pseudoelastic NiTi tubes under uniaxial tension. Extrem Mech Lett 15:113–121. https://doi.org/10.1016/j.eml.2017.06.006

    Article  Google Scholar 

  38. Jiang D, Kyriakides S, Landis CM, Kazinakis K (2017) Modeling of propagation of phase transformation fronts in NiTi under uniaxial tension. Eur J Mech A/Solids 64:131–142. https://doi.org/10.1016/j.euromechsol.2017.02.004

    Article  Google Scholar 

  39. Kazinakis K, Kyriakides S, Jiang D et al (2021) Buckling and collapse of pseudoelastic NiTi tubes under bending. Int J Solids Struct 221:2–17. https://doi.org/10.1016/J.IJSOLSTR.2019.12.017

    Article  Google Scholar 

  40. Xiao Y, Jiang D (2020) Constitutive modelling of transformation pattern in superelastic NiTi shape memory alloy under cyclic loading. Int J Mech Sci 182:105743. https://doi.org/10.1016/j.ijmecsci.2020.105743

    Article  Google Scholar 

  41. Kazinakis K, Kyriakides S, Landis CM (2022) Simulation of the response and evolution of localization in pseudoelastic NiTi tubes under biaxial stress states. Int J Plast 151:103179. https://doi.org/10.1016/J.IJPLAS.2021.103179

    Article  Google Scholar 

  42. Hill R (1952) On discontinuous plastic states, with special reference to localized necking in thin sheets. J Mech Phys Solids 1:19–30. https://doi.org/10.1016/0022-5096(52)90003-3

    Article  MathSciNet  Google Scholar 

  43. Xiao Y, Jiang D (2020) Rate dependence of transformation pattern in superelastic NiTi tube. Extrem Mech Lett 39:100819. https://doi.org/10.1016/J.EML.2020.100819

    Article  Google Scholar 

  44. Shaw JA, Kyriakides S (1997) Initiation and propagation of localized deformation in elasto-plastic strips under uniaxial tension. Int J Plast 13:837–871. https://doi.org/10.1016/S0749-6419(97)00062-4

    Article  Google Scholar 

  45. Jiang D, Landis CM (2016) A constitutive model for isothermal pseudoelasticity coupled with plasticity. Shape Mem Superelasticity 2:360–370. https://doi.org/10.1007/s40830-016-0078-8

    Article  Google Scholar 

  46. Needleman A (1988) Material rate dependence and mesh sensitivity in localization problems. Comput Methods Appl Mech Eng 67:69–85. https://doi.org/10.1016/0045-7825(88)90069-2

    Article  Google Scholar 

  47. Hallai JF, Kyriakides S (2011) On the effect of Lüders bands on the bending of steel tubes. Part II: analysis. Int J Solids Struct 48:3285–3298. https://doi.org/10.1016/J.IJSOLSTR.2011.07.012

    Article  Google Scholar 

  48. He YJ, Sun QP (2010) Macroscopic equilibrium domain structure and geometric compatibility in elastic phase transition of thin plates. Int J Mech Sci 52:198–211. https://doi.org/10.1016/J.IJMECSCI.2009.09.008

    Article  Google Scholar 

  49. Duval A, Haboussi M, Ben Zineb T (2011) Modelling of localization and propagation of phase transformation in superelastic SMA by a gradient nonlocal approach. Int J Solids Struct 48:1879–1893. https://doi.org/10.1016/J.IJSOLSTR.2011.02.019

    Article  Google Scholar 

  50. Stupkiewicz S, Rezaee-Hajidehi M, Petryk H (2021) Multiscale analysis of the effect of interfacial energy on non-monotonic stress–strain response in shape memory alloys. Int J Solids Struct 221:77–91. https://doi.org/10.1016/J.IJSOLSTR.2020.04.006

    Article  Google Scholar 

Download references

Acknowledgements

The authors acknowledge the financial support from National Natural Science Foundation of China (Grant No. 11902195).

Funding

This work was supported by National Natural Science Foundation of China (Grant No. 11902195).

Author information

Authors and Affiliations

Authors

Contributions

DJ contributed to conceptualization, methodology, funding acquisition, and supervision and provided resources; MW performed formal analysis and investigation; DJ and MW were involved in writing—original draft preparation and writing—review and editing.

Corresponding author

Correspondence to Dongjie Jiang.

Ethics declarations

Conflict of interest

The authors have no competing interests to declare that are relevant to the content of this article.

Appendix

Appendix

1.1 Constitutive model

The constitutive model is based on the classical kinematic hardening model of plasticity. The total strain is decomposed into elastic and transformation strains:

$$\varepsilon_{ij} = \varepsilon_{ij}^{e} + \varepsilon_{ij}^{t} .$$
(4)

The elastic strain is related to the stress tensor by elasticity:

$$\sigma_{ij} = C_{ijkl} \varepsilon_{kl}^{e} ,$$
(5)

where

$$C_{ijkl} = \frac{E}{{2\left( {1 + \nu } \right)}}\left( {\delta_{ik} \delta_{jl} + \delta_{il} \delta_{jk} + \frac{2\nu }{{1 - 2\nu }}\delta_{ij} \delta_{kl} } \right)$$
(6)

is the elastic stiffness tensor and E and ν are the Young’s modulus and Poisson’s ratio of the material, respectively.

The stress state is enclosed by a J2-type transformation surface described by

$$\Phi = \frac{3}{2}\left( {s_{ij} - s_{ij}^{B} } \right)\left( {s_{ij} - s_{ij}^{B} } \right) - \sigma_{o}^{2} = 0,$$
(7)

where \(s_{ij} = \sigma_{ij} - \sigma_{kk} \delta_{ij} /3\) and \(s_{ij}^{B} = \sigma_{ij}^{B} - \sigma_{kk}^{B} \delta_{ij} /3\).

The evolution of transformation strain abides by the following flow rule:

$$\dot{\varepsilon }_{ij}^{t} = \Lambda \left( {s_{ij} - s_{ij}^{B} } \right),$$
(8)

where Λ is a non-negative multiplier.

The back stress \(\sigma_{ij}^{B}\) is defined as the derivative of a potential function \(\psi^{t}\) with respect to the transformation strain tensor:

$$\sigma_{ij}^{B} = \frac{{\partial \psi^{t} }}{{\partial \varepsilon_{ij}^{t} }}.$$
(9)

To introduce the tension/compression asymmetry of superelastic NiTi, the following form of the potential function is adopted:

$$\psi^{t} = \xi \psi_{C}^{t} (\overline{\varepsilon }^{t} ) + \left( {1 - \xi } \right)\psi_{T}^{t} (\overline{\varepsilon }^{t} ),$$
(10)

where \(\psi_{C}^{t}\) and \(\psi_{T}^{t}\) are two potential functions calibrated to the compressive and tensile stress–transformation strain responses, respectively. They are assumed to depend on an equivalent transformation strain \(\overline{\varepsilon }^{t}\), which is defined as

$$\overline{\varepsilon }^{t} = J_{2}^{e} f\left( {J_{r} } \right).$$
(11)

In Eq. (11), \(J_{2}^{e}=(2e_{ij}^{t} e_{ij}^{t} /3)^{1/2}\) and \(J_{3}^{e} = \left( {4e_{ij}^{t} e_{jk}^{t} e_{ki}^{t} /3} \right)^{1/3}\) are, respectively, the second and third invariants of the deviatoric transformation strain and \(J_{r} = {{J_{3}^{e} } \mathord{\left/ {\vphantom {{J_{3}^{e} } {J_{2}^{e} }}} \right. \kern-0pt} {J_{2}^{e} }}\). The ratio \(J_{r}\) represents the transformation state, since it reaches − 1 under uniaxial contraction and 1 under uniaxial extension, and always lies in the interval [− 1, 1] in other cases. The function f scales \(J_{2}^{e}\) to the equivalent transformation strain \(\overline{\varepsilon }^{t}\) and it takes the following form:

$$f = \cos \left\{ {\frac{1}{3}\arccos \left[ {1 - a\left( {J_{r}^{3} + 1} \right)} \right]} \right\}.$$
(12)

One should note that \(f( - 1) = 1\) and \(f(1) < 1\). Resultantly, f (1) is equals to the ratio between the maximum transformation strain under compression versus that under tension.

The weight function ξ(Jr) aims to interpolate between the compressive and the tensile responses, by requiring ξ(− 1) = 1 and ξ(1) = 0. It takes the following form:

$$\xi = \frac{{f\left( {J_{r} } \right) - f\left( 1 \right)}}{{f\left( { - 1} \right) - f\left( 1 \right)}}.$$
(13)

The back stress is then derived as

$$\sigma_{ij}^{B} = \xi \frac{{{\text{d}}\psi_{C}^{t} }}{{{\text{d}}\overline{\varepsilon }^{t} }}\frac{{\partial \overline{\varepsilon }^{t} }}{{\partial \varepsilon_{ij}^{t} }} + \left( {1 - \xi } \right)\frac{{{\text{d}}\psi_{T}^{t} }}{{{\text{d}}\overline{\varepsilon }^{t} }}\frac{{\partial \overline{\varepsilon }^{t} }}{{\partial \varepsilon_{ij}^{t} }} + \left( {\psi_{C}^{t} - \psi_{T}^{t} } \right)\frac{\partial \xi }{{\partial \varepsilon_{ij}^{t} }}.$$
(14)

In uniaxial settings, as shown in Appendix A in [29] and [45], Eq. (14) can be reduced to

$$\sigma_{11,C}^{B} = \frac{{{\text{d}}\psi_{C}^{t} }}{{{\text{d}}\overline{\varepsilon }^{t} }}\frac{{\partial \overline{\varepsilon }^{t} }}{{\partial \varepsilon_{11}^{t} }}\;\;{\text{and}}\;\;\sigma_{11,T}^{B} = \frac{{{\text{d}}\psi_{T}^{t} }}{{{\text{d}}\overline{\varepsilon }^{t} }}\frac{{\partial \overline{\varepsilon }^{t} }}{{\partial \varepsilon_{11}^{t} }},$$
(15)

for compression and tension, respectively. Resultantly, the transformation function in Eq. (7) becomes

$$\left| {\sigma_{11,C} } \right| - f( - 1)\frac{{{\text{d}}\psi_{C}^{t} }}{{{\text{d}}\overline{\varepsilon }^{t} }} = \sigma_{oC} \;\;{\text{and}}\;\;\sigma_{11,T} - f(1)\frac{{{\text{d}}\psi_{T}^{t} }}{{{\text{d}}\overline{\varepsilon }^{t} }} = \sigma_{oT} ,$$
(16)

for loading, and a minus sign will appear in the right-hand sides for unloading. Obviously, the back stress under uniaxial compression and tension is governed by the derivative of the potential functions \(\psi_{C}^{t}\) and \(\psi_{T}^{t}\), respectively, which are assigned the same form:

$$\frac{{{\text{d}}\psi^{t} }}{{{\text{d}}\overline{\varepsilon }^{t} }} = h_{0} \overline{\varepsilon }^{t} + \left( {h_{1} - h_{0} } \right)\left[ {\overline{\varepsilon }^{t} - \frac{1}{b}\left( {1 - e^{{ - b\overline{\varepsilon }^{t} }} } \right)} \right] + \left( {h_{2} - h_{1} } \right)\left\{ {\begin{array}{*{20}c} 0 \\ {\left( {\varepsilon_{2} - \varepsilon_{1} } \right)\left( {{{\zeta^{3} } \mathord{\left/ {\vphantom {{\zeta^{3} } 4}} \right. \kern-0pt} 4} + {{\zeta^{2} } \mathord{\left/ {\vphantom {{\zeta^{2} } 8}} \right. \kern-0pt} 8}} \right)} \\ {{{3\left( {\varepsilon_{2} - \varepsilon_{1} } \right)} \mathord{\left/ {\vphantom {{3\left( {\varepsilon_{2} - \varepsilon_{1} } \right)} 8}} \right. \kern-0pt} 8} + \left( {\overline{\varepsilon }^{t} - \varepsilon_{2} } \right)} \\ \end{array} } \right. \, \begin{array}{*{20}c} {0 \le \overline{\varepsilon }^{t} < \varepsilon_{1} } \\ {\varepsilon_{1} \le \overline{\varepsilon }^{t} \le \varepsilon_{2} } \\ {\varepsilon_{1} \le \overline{\varepsilon }^{t} } \\ \end{array} ,$$
(17)

where \(\zeta = {{\left( {\overline{\varepsilon }^{t} - \varepsilon_{1} } \right)} \mathord{\left/ {\vphantom {{\left( {\overline{\varepsilon }^{t} - \varepsilon_{1} } \right)} {\left( {\varepsilon_{2} - \varepsilon_{1} } \right)}}} \right. \kern-0pt} {\left( {\varepsilon_{2} - \varepsilon_{1} } \right)}}\). The parameters b, h0, h1, h2, ε1, and ε2 for compression and those for tension are calibrated separately.

The size of the transformation strain σo in Eq. (7) also shows asymmetry between tension and compression and the following form is adopted:

$$\sigma_{o} = \left\{ {\begin{array}{*{20}c} {\sigma_{oC} } \\ {\sigma_{oC} + \left( {\sigma_{oT} - \sigma_{oC} } \right)\left( {10\eta^{3} - 15\eta^{4} + 6\eta^{5} } \right)} \\ {\sigma_{oT} } \\ \end{array} } \right. \, \begin{array}{*{20}c} {\eta < 0} \\ {0 \le \eta \le 1} \\ {\eta > 1} \\ \end{array} ,$$
(18)

where \(\eta = \left( {J_{3}^{e} + 0.01} \right)/0.0098\).

The calibrated parameters in the constitutive model are listed in Table 1. The uniaxial compressive and tensile stress–strain responses produced by the calibrated model are shown in Fig. 14.

Table 1 Model parameters for compressive and tensile responses
Fig. 14
figure 14

The resultant tensile and compressive stress–strain responses together with the measured ones

1.2 Mesh sensitivity of the simulation results

Incorporation of softening in the material’s tensile stress–strain response implies mesh dependence of the finite element solution, which can be removed by adding a mild rate dependence into the constitutive model (see [46, 47] for example) or regularizing the model by gradient-related energy terms, like [25, 48, 49]. In this paper, the adopted constitutive model for superelastic NiTi is rate independent; hence, we must examine the mesh sensitivity of the finite element simulations. Therefore, we adopted three mesh densities to compare their performances on simulating the uniaxial tension test. Mesh 1 has 1 element in the radial (through-the-thickness) direction, 90 elements in the circumferential direction, and 225 elements in the length direction. Mesh 2 (the mesh density used in this study) has 1, 112, and 280 elements in the radial, circumferential, and axial directions, respectively. Mesh 3 has 1, 160, and 400 elements in the radial, circumferential, and axial directions, respectively. All three meshes contain only one element through the thickness based on the consensus that one element is enough for successful simulation of the localization phenomena in thin walls since the variation of strains and other quantities through the thickness is nearly negligible [40, 44, 50].

The stress–strain responses calculated using the three meshes are shown in Fig. 15. It is seen that Meshes 2 and 3 produce almost identical stress–strain curves, regarding the levels of the upper and lower plateaus and the strain at completion of martensitic transformation, while the stress plateaus obtained by Mesh 1 are obviously higher than those by the two denser meshes. As the mesh density increases from Mesh 1 to Mesh 3, the calculated stress–strain curve exhibits a trend of convergence. The calculated stress–strain curve using Mesh 2 is much closer to the experimental one compared with Mesh 1, but the improvement brought by further refinement into Mesh 3 is not that significant.

Fig. 15
figure 15

Simulation results of stress–strain responses under uniaxial tension with the three meshes

Figure 16 shows a set of axial strain contours in the FE simulation results using the three meshes. At the onset of localization, all three meshes exhibit similar localization bands with an angle of about 55° with respect to the axial direction, which is consistent with experimental observations. However, the localization band of Mesh 1 quickly turns perpendicular to the axial direction and shows distinct features to Mesh 2 and Mesh 3, both of which produce a finger-like propagating front that travels all the way to the bottom of the tube until the completion of the forward transformation. During unloading, the localization band of Mesh 1 keeps perpendicular to the axis while propagating in the reverse direction to that during loading, while Mesh 2 and Mesh 3 exhibit a spiral band that extends along the axis. The finger-like and spiral localization bands under uniaxial tension and the orientation angle of approximately 55° have been predicted theoretically [42] and reported widely in the literature, see [13, 19, 50] and other publications. Mesh 1 fails but both Meshes 2 and 3 successfully reproduce such important features of localized deformation in superelastic NiTi.

Fig. 16
figure 16

Simulation results of strain contours during uniaxial tension with the three meshes

Based on the results of FE simulations on uniaxial tension of the superelastic NiTi tube using the three mesh densities, we can conclude that Mesh 2 is sufficiently fine to reproduce the experimentally recorded stress–strain response and finer features of localization of deformation, which are very similar to the simulations by the finer mesh. Thus, this paper chose Mesh 2 for finite element modeling in order to reach a tradeoff between accuracy and computational cost, as was done in other works on superelastic shape memory alloy thin-walled tubes [37, 38, 40]. Main effects of mesh sensitivity lie on the width of the transition zones between the high and low strain regions in the localization band, which can be eliminated by introducing rate dependence or gradient energies into the constitutive model in the future.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, M., Jiang, D. Modeling of the stress–strain responses and deformation patterns of superelastic NiTi tubes subjected to biaxial loadings. AS (2024). https://doi.org/10.1007/s42401-023-00266-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s42401-023-00266-x

Keywords

Navigation